.wfi .s. u, . _, .fi .. unarmwfimfia .s ./ Gr“ ”My” WW. .mu/r... N n J... . an; Mina. . ..., .an.§.«.,..%; .n H... aw: V... . ,. ,..._._._.»£z.n...,wm s... «we... 2.: , V v Iv. . a .m. w... my.” , . 0 19*». . H . xx. .. I‘U «Inna.§r J . aways” 2a.”? .. :1) « .l.a 9 v (A. o In... v 1.: a f. n‘ a. Run a A no: Vt; # ‘I .. llb A; . J r «L. ._)v.w11 ‘ 2V5! 11.1031: . um?“ 31$: i t 53;; ”"do ‘ , l o I v 'H' .n,‘ I u an {1-..}-.. n 2 . ; 7,: ....u..3\..)&mi.n....XJhlll! Wm: -..lr.r.. . tunic...”th (lo . I'll-‘0 , . III'I - I lo 3.4.: I .| (.ltlu:.fll‘!t..)lovl\.‘lll ‘. -.0Tl OOIJUFII” !| l l O “l'-’ I..l.“|o ‘ '.V.~‘ llflpl tI‘P‘§U‘L 1‘ THESlS 3 lllllllllllllllllllllllllllllllllllllllllllllllllllllIll 1293 01690 0684 This is to certify that the dissertation entitled Biochemical and NMR Studies of Human Cellular Retinoic Acid Binding Proteins presented by Lincong Wang has been accepted towards fulfillment of the requirements for Ph . D . degree in Biochemistry Hm 27% Major protessor a! Date MSU is an Affirmative Action/Equal Opportunity Institution LIBRARY Mlchlgan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINE return on or before date due. DATE DUE MTE DUE DATE DUE 1/96 alClRC/DalaDuopGS—p.“ BIOCHEMICAL AND NMR STUDIES OF HUMAN CELLULAR RETINOIC ACID BINDING PROTEINS By Lincong Wang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Biochemistry 1998 Copyright by LINCONG WANG 1 998 BIOCHEMICAL AND NMR STUDIES OF HUMAN CELLULAR RETINOIC ACID BINDING PROTEINS By Lincong Wang ABSTRACT The structure-fimction relationships of cellular retinoic acid binding proteins (CRABPS) have been investigated by site-directed mutagenesis, biochemical and NMR methods. The major results and conclusions are summarized as follows. Electrostatic interactions have been suggested to be critical for binding of retinoic acid (RA) by CRABPs. However, the roles of two conserved arginine residues (Argl ll and Arg13l in type-I CRABP; Argl 11 and Argl32 in type-II CRABP) that interact with the carboxyl group of RA have not been evaluated. A novel competitive binding assay was developed for measuring the relative dissociation constants of the site-directed mutants including the two mutants R1 11M and R132M. Binding studies of the two mutants showed that Argl 11 and Argl32 contribute to the binding energy by ~2.2 and 1.2 kcal/mol, respectively. The crystal structure of holo-CRABP suggests that RA cannot enter or exit the deep binding pocket without major conformational changes in the protein. Dimerization has been proposed to be the mechanism by which CRABPs open their binding pockets. The solution structure of apo-CRABPII has been determined by NMR spectroscopy. Although the apo solution structure was similar to the holo crystal structure, the ligand entrance was greatly enlarged in the former and readily accessible to RA. The enlargement was mainly due to a concerted conformational change in three structural elements, namely the second helix, the BC-BD loop and the BE-BF loop. The ligand- binding pocket of apo-CRABPII was rather dynamic. CRABPII is predominately monomeric in solution. Although the formation of transient dimers could not be ruled out, dimerization apparently is not a prerequisite for entry of retinoic acid into the binding pocket of CRABPII. Preliminary NMR analysis of holo-CRABPII showed that RA binds to CRABPII in solution in the same manner as determined by crystallography. NMR studies indicated that the ligand entrance of holo-CRABPII is much less flexible than that of apo- CRABPII. The results taken together suggested that RA binding indeed induces major changes in the conformation and dynamics of CRABPII. The solution structure of R11 1M has been determined by NMR spectroscopy. Although apo-R1 11M had a structure similar to that of wild-type apo-CRABPII, there were significant conformational differences between the two proteins, mainly localized to three segments, clustered around the ligand entrance more than 17 A from the site of point mutation. Furthermore, the ligand-binding pocket of apo-R1 1 1M, especially the ligand entrance, was much less flexible than that of apo-CRABPII. The results suggested that Argl 11 play a critical role in determining the structure and dynamics of CRABPII. To my family ACKNOWLEDGMENTS I would like to express my gratitude to my thesis advisor, Dr. Honggao Yan, for his financial support and for providing me the unique opportunity to learn both protein biochemistry and NMR. I have learned a lot from him and become a better researcher. I would like to thank my guidance committee members Dr. David McConnell, Dr. John McCraken, Dr. Robert Hausinger, Dr. John Wilson and Dr. William Wells for their criticism and advice, which truly guided me through the tough journey of graduate school. I would like to thank Mr. Kermit Johnson for teaching me NMR. His kindness and patience made doing NMR experiments much more enjoyable. Thanks also go to Dr. Yue Li for teaching me molecular biology. Special thanks go to Dr. Robert Cukier of Chemistry Department for his advice on my thesis. His comment, “You have summarized a great deal of (complicated) wor ”, encourages me to go further. In experimental aspect I would like to thank Dr. Yue Li for establishing the expression systems for CRABPs and for generating the mutants, Dr. Frits Abildgaard at the National Magnetic Resonance Facility at Madison, Wisconsin for recording some of the triple resonance spectra of apo-CRABPII and Dr. Honggao Yan for acquiring the NMR spectra of uniformly 15N-labeled apo-R11 1M. Finally, I am grateful to my wife for her understanding and support. vi TABLE OF CONTENTS LIST OF TABLES ......................................................................................................... ix LIST OF FIGURES ........................................................................................................ x LIST OF ABBREVIATIONS ....................................................................................... xiv CHAPTER 1 LITERATURE REVIEW .............................................................................................. 1 Part 1. Cellular retinoic acid binding protein-II (CRABPII) .................................... 1 Section 1. Biological functions ofCRABPII 1 Section 2. Biochemical characterizations of CRABPs ............................. 7 Section 3. Structural and dynamical characterizations of CRABPs ....... 13 References ............................................................................. 22 Part 2. Physical basis ofprotein NMR 28 Section 1. Theory of Fourier transform pulse NMR ............................... 28 Section 2. NMR experiments for protein studies .................................... 42 Section 3. Theories of NMR relaxation in liquid .................................... 47 References ............................................................................. 77 CHAPTER 2 STRUCTURE-FUNCTION RELATIONSHIPS OF CELLULAR RETINOIC ACID BINDING PROTEINS ................................................................................................... 87 Abstract ................................................................................................................... 88 Experimental procedures ...................................................................................... . 89 Results ................................................................................................................... 90 Discussion ............................................................................................................... 91 References .............................................................................................................. 94 vii CHAPTER 3 SOLUTION STRUCTURE OF TYPE-II HUMAN CELLULAR RETINOIC ACID BINDING PROTEIN ....................................................................... 95 Introduction ............................................................................................................. 95 Experimental procedures ........................................................................................ 97 Results .................................................................................................................. 103 Discussion ............................................................................................................. 1 26 Conclusion ................................................................................... 1 36 References ............................................................................................................. 137 CHAPTER 4 SOLUTION STRUCTURE OF A MUTANT PROTEIN (R1 11M) OF HUMAN CRABPII ...................................................................................................................... 141 Introduction ......................................................................................................... . 141 Experimental procedures ..................................................................................... 143 Results ................................................................................................................. 148 Discussion ............................................................................................................ 1 70 Conclusion .................................................................................. 1 85 References ............................................................................................................ 186 CHAPTER 5 NMR STUDIES OF HUMAN CRABPII IN COMPLEX WITH ALL-TRANS—RETINOIC ACID ............................................................... 189 Introduction .......................................................................................................... l 89 Experimental procedures ...................................................................................... 191 Results and Discussion ........................................................................................ 193 Conclusion .................................................................................. 2 1 9 References ............................................................................................................ 220 viii LIST OF TABLES CHAPTER 2 Table 2.1. 1H Chemical shifts of the ring protons of the aromatic residues of the wild-type CRABPII, R11 1M mutant, and CHAPTER 3 Table 3.1. lH, 13C and 15N chemical shifis of apo-CRABPII at pH 7.3 Table 3.2. Restraint and structural statistics of apo-CRABPII ........................ 124 CHAPTER 4 Table 4.1. 'H and 15N chemical shifts of Apo-Rl 1 1M at pH 7.3 and 27 °C ............................................................................................ 155 Table 4.2. Restraint and structural statistics ofapo-Rl 1 1M 164 CHAPTER 5 Table 5.1. 1H and 15N chemical shifts of holo-CRABPII at pH 7.3 and 25 °C ......................................................................................... 194 Table 5.2. Summary of proton chemical shifts of CRABPII-bound RA at pH 7.3 and 25 °C ............................................................................ 207 Table 5.3. Relative volumes of intramolecular NOEs of CRBAPII-bound RA ......................................................................... 208 ix CHAPTER 1 Figure 1.1. Figure 1.2. Figure 1.3. Figure 1.4. CHAPTER 2 Figure 2.1. Figure 2.2. Figure 2.3. Figure 2.4. Figure 2.5. Figure 2.6. CHAPTER 3 Figure 3.1. LIST OF FIGURES Chemical structures and the numbering conventions of all-trans-retinol (vitamin A), all-trans—retinal and all-trans-retinoic acid (RA) ................................................... 3 (A) The ligand binding pocket of human CRABPII. (B) expanded view of the interactions between the carboxyl group of RA and the side chains of Argl 11, Argl32 and Tyrl34 ofthe protein 12 The backbone fold and the secondary structural elements of Huma CRABPII in complex with all-trans-RA ............................ 16 The polypeptide backbone and side chains showing the one—bond coupling constants through which the coherences are transferred in double and triple resonance NMR experiments 44 F luorometric titration of the wild type CRABPI (A) and CRABPII (B) ....................................................................................... 90 Competitive binding assays for measuring the relative dissociation constants of CRABPs ....................................................... 91 TOCSY spectrum of the aromatic protons of R132M at 32° showing two sets of resonances for Trp-87 ......................................... 92 Parts of the 500-MHz NOESY spectrum with 150 ms mixing time ofthe wild type CRABPII at 32°C 92 Parts of the SOO-MHz NOESY spectrum with 150 ms mixing time ole 1 1M at 32°C .. 83 Parts of the 500-MHz NOESY spectrum with 150 ms mixing time of R132M at 32°C .......................................................... 93 Strips of the HNCACB (a) and CBCA(CO)NH (b) spectra showing sequential connectivies for residues Gly68—Gln74 . Figure 3.2. Gradient- and sensitivity-enhanced 2D lH—‘SN HSQC spectrum ofuniformly ISN-labeled apo-CRABPII . . . Figure 3.3. The fingerprint region of the homonuclear 2D TOCSY spectrum with40 ms mixing t1me Figure 3.4. Summary of the sequential and medium-range NOEs involving HN and H“ atoms, slow-exchange backbone amide protons and the deduced secondary structures of apo-CRABPII Figure 3.5. The distribution of NOEs along the amino sequence of Figure 3.6. (A) Stereoview of the C“ traces of the superimposed 25 refined structures of apo-CRABPII. (B) Stereoview of the C“ trace of the restrained minimized mean structure of apo-CRABPII (thin line) superimposed with the C“ trace of the crystal structure of holo-CRABPII (thick line). (C) Stereoview of the C“ trace of the restrained minimized mean structure of apo-CRABPII (thin line) superimposed with the C“ trace of the crystal structure of apo-CRABPI (molecule A) (thickline) Figure 3.7. The relative peak intensity of the 2D lH—‘sN HSQC spectrum of apo-CRABPII color-coded along the C“ trace of the solution structure. The strong peaks are in white, the weak peaks in pink and the missing peaks in cyan. RA is positioned on the basis of the superposition of the solution structure of apo-CRABPII and the crystal structure ofholo-CRABPII CHAPTER 4 Figure 4.1. Gradient- and sensitivity-enhanced 2D 'H—‘SN HSQC spectrum ofuniforrnly 15N-labeled of apo-R1 1 1M Figure 4.2. 2D lH-‘SN HMQC spectrum of apo-R1 1 1M selectively labeled with l5N-leucine Figure 4.3. Strip plots extracted from the 3D 15N-edited NOESY- HMQC spectrum of apo-R1 11M showing the sequential and medium-range NOE connectivies of or-helix A (A) xi 105 107 111 118 121 123 131 150 ......152 and B-strand J (B) .............................................................................. 159 Figure 4.4. Summary of the sequential and medium-range NOEs involving HN and H“ atoms, slow-exchange backbone amide protons and the deduced secondary structures ofapo-R111M162 Figure 4.5. The distribution of NOEs along the amino sequence of CRABPII ...................................................................................... 166 Figure 4.6. (A) Stereoview of the C“ traces of the superimposed 28 refined solution structures of apo-R1 1 1M. (B) Stereoview of the superimposed C“ traces of the restrained minimized mean structures of apo-R1 1 1M (thin line) and wild type apo-CRABPII (thick line). (C) Stereoview of the C“ trace of the restrained minimized mean structure of apo- R1 1 1M (thin line) superimposed with the C“ trace of the crystal structure ofholo-CRABPII (thick line) 168 Figure 4.7. C“ atom deviations between apo-R11 1M and wild-type holo-CRABPII (A) and between apo-R1 1 1M and wild-type apo-CRABPII (B) 173 Figure 4.8. The fingerprint regions of the homonuclear 2D TOCSY spectra of apo-CRABPII (A) and apo-R1 11M (B) recorded with 40 ms mixing time . ...............176 Figure 4.9. Relative peak intensities of the 2D lH—‘SN HSQC spectra of apo-R1 1 IM (A) and wild type apo-CRABPII (B) as a function of the residue number. (C) was obtained by subtracting (B) from (A) . ............180 CHAPTER 5 Figure 5.1. Summary of the sequential and medium-range NOEs involving HN and H“ atoms and the deduced secondary structures of holo—CRABPII .............................................................. 198 Figure 5.2. Strip plots extracted from the 3D lsN-edited NOESY-HSQC spectrum (150 ms mixing time)of holo-CRABPII showing the NOE connectivies for residues Met27—Lys38 200 Figure 5.3. Gradient- and sensitivity-enhanced 2D lH—‘SN HSQC spectrum of uniformly 'SN-labeled of holo-CRABPII ......................... 203 xii Figure 5.4. Relative peak intensities of the 2D lH—‘SN HSQC spectra of apo-CRABPII (A) and holo-CRABPII (B) as a function of the residue number. (C) was obtained by subtracting (B) from (A) ...................................................................................... 205 Figure 5.5. Part of the 2D T OCSY spectrum of holo-CRABPII with 33 ms mixing time at 25°C ..................................................................... 210 Figure 5.6. Part of the NOESY spectrum of holo-CRABPII with 100 ms mixing time at 25°C ........................................................................... 212 Figure 5.7. Relative volumes (V "”6) of some of the NOEs observed for bound RA versus the internuclear distance (r) measured from the crystal structure of holo-CRABPII .............................................. 214 xiii BPP theory: CD: CMPG: COSY: CRABPI: CRABPII: CRBPI: CRBPII: CSA: DEPT: DGSA: DMSO: DQF-COSY: E.COSY: FABP: FID: HMQC: HSQC: iLBP: INEPT: IPTG: LIST OF ABBREVIATIONS Bloembergen-Purcell-Pound theory circular dichroism Carr-Purcell-Meiboom-Gill correlated spectroscopy type-I cellular retinoic acid binding protein type II-cellular retinoic acid binding protein type-II cellular retinol binding protein type-II cellular retinol binding protein chemical-shielding anisotropy distortionless enhancement by polarization transfer distance geometry-simulated annealing dimethyl sulfoxide double-quantum-filtered correlated spectroscopy exclusive correlated spectroscopy fatty acid binding protein free-induction decay heteronuclear multiple-quantum coherence heteronuclear single-quantum coherence intracellular lipid binding protein insensitive nuclei enhanced by polarization transfer isopropylthiogalactoside xiv NMR: NOE: NOESY: P.E.COSY: RA: RAR: RMSD: ROESY: TOCSY: TPPI: 2D: 3D: 4D: nuclear magnetic resonance nuclear Overhauser effect nuclear Overhauser enhancement and exchange spectroscopy primitive exclusive correlated spectroscopy retinoic acid retinoic acid receptor root-mean-square deviation rotating frame Overhauser effect spectroscopy retinoid X receptor total correlation spectroscopy time proportional phase incrementation two-dimensional three-dimensional four-dimensional XV CHAPTER 1 Literature Review Part 1. Cellular Retinoic Acid Binding Proteins Section 1. Biological Functions of CRABPII Vitamin A (retinol) and its metabolic derivatives (naturally occurring retinoids) such as retinal and retinoic acid (RA) (Figure 1.1) are essential for vertebrates, with profound effects on a wide variety of biological processes such as vision, cell differentiation, embryonic development and homeostasis (Spom et al., 1994; Chambon et al., 1996). Both the naturally occurring retinoids and their synthetic analogs have been used to treat several kinds of human diseases including skin disorders (Peck & DiGiovanna, 1994), and certain cancers, such as cervical cancer and acute promyelocytic leukemia (Hong & Itri, 1994). A unified picture of the apparent pleiotropic effects of retinoids is gradually emerging, and some key components of the retinoid signal transduction pathway have been discovered over the past five decades. It has been shown that retinal is a key component of visual signal transduction (Wald, 1968), while RAs, Figure 1.1 Chemical structures and the numbering conventions of all-trans-retinol (vitamin A), all-trans-retinal and all-trans-retinoic acid (RA). All-trans-retinoic acid (RA) especially all-trans-RA and 9-cis-RA, are the active forms of vitamin A in most of the other biological processes. RA administration can prevent or reverse the majority of the defects originating from post-natal vitamin A deficiency. Two distinct classes of RA- binding proteins have been found to be involved in retinoid signal transduction: nuclear retinoid receptors and cellular RA binding proteins (CRABPs). Nuclear retinoid receptors. There exist in cells two retinoid receptor families: the RA receptor (RAR) family with three subtypes RAR-0t, ~13 and -y, and the retinoid X receptor (RXR) family also with three subtypes RXR-0t, -B and -y (Mangelsdorf et al., 1994; Mangelsdorf & Evans, 1995; Chambon, 1996). The nuclear retinoid receptors play critical roles in retinoid signal transduction. They have been shown to mediate many of the biological effects observed with administration of vitamin A. For example, all abnormalities exhibited by fetuses of vitamin A deficient dams are recapitulated in RAR or RXR single mutants and RAR-RXR double mutants of mouse (Chambon, 1996). Amino acid sequence comparison suggests that retinoid receptors belong to a protein superfarnily of ligand-activated transcriptional factors (Evans, 1988; Tsai & O’Malley, 1994) that includes also the receptors for steroid and thyroid hormones and vitamin D3. The RAR proteins are activated by both all-trans-RA (K, = ~1-5 nM) and 9- cis-RA with similar effectiveness, whereas the RXR proteins bind exclusively 9-cis-RA (K, = ~10 nM) (Heyman et al., 1992; Levin et al., 1992). In vitro studies have demonstrated that the functional units of the retinoid receptors are either the RAR-RXR heterodimers consisting of one RXR subunit and one RAR subunit, or the homodimers of two RXR subunits. These two forms of dimer appear to have different functions. The heterodimers bind the DNA response elements with direct repeats of the hexamer (purine- G(G/T)TCA) separated by one, two, or five base pairs, while the RXR homodimers prefer the DNA sequence with direct repeats of the hexamer separated by only one base pair. For the same response element the heterodimers have higher affinity than the homodimers. Interestingly, RXR proteins also form heterodimers with other nuclear receptors such as thyroid hormone and vitamin D3 receptors, and are required for the biological functions of these receptors (see, for example, Yu et al., 1991). The diversity of the biological effects of retinoids can be partly explained by the existence of two families of receptors and the interactions between the two families of receptors themselves and with other receptors. It appears that the availability of RA in cells is regulated by other molecules such as the two isoforrns of CRABP. Biological significance of CRABPs. There are two types of cellular retinoic acid binding protein (CRABPI and CRABPII) (Sani & Hill, 1974; Ong & Chytil, 1975; Bailey & Siu, 1988; Kitamoto et al., 1988). They have been found in all vertebrates relying on vitamin A (Ong et al., 1994), and are highly conserved proteins. For example, rat, mouse and cow CRABPI (Nilsson et al., 1988; Stoner & Gudas, 1989; Shubeita et al., 1987) all have identical 136 residues, and human CRABPI has only one residue difference (Astrom et al., 1991). Of 137 residues of mouse CRABPII (Giguere et al., 1990) and human CRABPII (Astrom etal., 1991; Eller et al., 1992) 130 are identical. Although CRABPs have been suggested to play important roles in mediating the biological functions of RA, their physiological functions have not been well defined. They may be involved in directing the spatial organization of cells during development, especially at the time of the limb generation (Maden, 1991; Hofinannn & Eichele, 1994; Morriss-Kay & Sokolova, 1996). Experimental studies have shown that both CRABPI and CRABPII are widely expressed at all stages of mouse embryo development. In particular, the mRNAs of CRABPs form an anteroposterior gradient opposite to the gradient of all-trans-RA in the budding of a chick limb (Maden et al., 1988). It has also been proposed that CRABPs may protect cells by preventing RA from being incorporated into cell membranes. The solubility of RA in aqueous solution is about 200 nM (Szuts & Harosi, 1991), which is higher than its endogenous concentration (~50 nM). However, RA molecules prefer the membrane to the cytoplasm with ~6 x105 fold higher affinity. Indeed, in tissues that express CRABPs, RA is predominately in the protein-bound forrrrs. Both CRABPI and CRABPII have very high affinity for all-trans-RA with an apparent dissociation constant (K) less than 2 nM (Norris et al., 1994; Wang et al., 1997). CRABPs, CRABPI in particular, may participate in the catabolism of all-trans-RA (Fiorella & Napoli, 1991, 1994; Boylan & Gudas, 1992). The rate of the breakdown of RA into inactive metabolites catalyzed by cytochrome P450 increases more than twenty- fold when it binds CRABPI. CRABPs may also be directly involved in the transport of RA to the nucleus (Takase et al., 1986; Blomhoff et al., 1992). Taken together, these in vitro studies lead to the proposal that CRABPs may act as ‘buffer’ to control the actual level of ‘free’ intracellular RA available for binding to the nuclear retinoid receptors. However, the physiological functions of CRABPs remain controversial. Genetic experiments have demonstrated that the mutant mice lacking CRABPI and/or CRABPII are essentially indistinguishable from the wild-type (de Bruijn et al., 1994; Gorry et al., 1994; Fawcett et al., 1995; Lampron et al., 1995) with the exception of a minor limb malformation. These results suggest that their actual function may be to control physiological levels of intracellular RA when the supply of vitamin A from diet is insufficient. The biological functions of CRABPI and II seem to be different. The homology within either CRABPI or CRABPII sequences across different species is much higher than the homology between CRABPI and CRABPII from the same species. For example, human CRABPI and CRABPII have only 74% identity. Moreover, CRABPI and CRABPII exhibit different expression patterns during development or in adults. In mouse embryo they are expressed in distinct, non-overlapping patterns. In adults CRABPI is expressed in a variety of tissues, while CRABPII is mainly restricted to the epidermis (Astrom et al., 1991; Elder et al., 1992). The CRABPII gene, not CRABPI, has a retinoid response element in its promoter region so that CRABPII can be induced markedly in cultured cells and human skin by 9-cis-RA and, to a lesser extent, by all-trans-RA (Giguere etal., 1990; Astrom et al., 1991; Durand et al., 1992; Melhus et al., 1994). It has been shown that production of RA is correlated with the expression of CRABPII but not CRABPI (Bucco et al., 1997). Overexpression of CRABPII enhances cellular response to RA in breast cancer cells (Jing et a., 1997) but overexpression of CRABPI decreases the biological potency of RA in F9 teratocarcinoma cells (Boylan & Gudas, 1991). However, no difference in binding specificity has been reported between CRABPI and CRABPII, and their affinities for retinoids are very similar. Section 2. Biochemical Characterizations of CRABPs Extensive biochemical studies have been performed on CRABPs. These studies have been expedited by cloning and overexpression of the genes encoding CRABPs in E. coli (Giguere et al., 1990; Astrom et al., 1991). The in vitro biochemical functions of CRABPs have been established by ligand binding assays, and the residues critical for such functions have been investigated by site-directed mutagenesis. Binding assay. The ligand-binding properties of CRABPs have been studied by several methods, which could be classified into two categories: non-equilibrium methods and fluorescence titration methods. The non-equilibrium methods remove the unbound ligand either by the use of dextran-coated charcoal (Daly & Redfem, 1988; Fogh et al., 1993) or by chromatography (Siegenthaler et al., 1992). A potential problem with such methods is that the bound ligand can come off the protein during the separation process. The de measured by these methods depend on the off-rate of the ligand, and may be an overestimation of its true value. The fluorescence methods (Cogan et al., 1976) measure the concentration of either the bound ligand or the bound protein at equilibrium. However, reported fluorescence measurements have been plagued by poor experimental designs and data analysis. The problems have been recognized recently (Norris et al., 1994; Wang et al., 1997). Furthermore, because of the low solubility of RA in aqueous solution, the fluorescence methods are unsuitable for measuring the dissociation constants of site-directed mutants with greatly reduced affinity for RA. The reported Kd of CRABPI for all-trans-RA ranges from 0.4 nM to 10 nM (Ong & Chytil, 1978; Fiorella & Napoli, 1991; Norris, et al., 1994; Fogh et al., 1993), and the K, of CRABPII from 1.8 nM to 65 nM (Bailey & Siu, 1988; Fiorella & Napoli, 1993; Fogh et al., 1993; Norris et al., 1994; Wang et al., 1997). These discrepancies may be caused by the different methods used, the assay conditions and the sources of the proteins. The reported relative Kd of CRABPI vs. CRABPII ranges from three-fold (Fiorella, et al., 1993; Wang et al., 1997), as measured by competitive binding assay, to more than ten-fold based on their individual absolute values. The value measured by competitive binding assay appears to be more reliable. The stoichiometry of either CRABPI or II for all-trans-RA has been found to be one-to- one. Binding specificity. No significant differences in binding specificity for retinoids have been reported between CRABPI and CRABPII. Both proteins show no apparent affinities for all-trans-retinol, all-trans-retinal and esters of RA, and have relatively low affinities for the cis-isomers of RA. For example, their affinities for 9-cis-RA range from ~50 nM (Fiorella etal., 1993) to ~200 nM (Norris etal., 1994), and their affinities for 13- cis-RA are at least 3-4 fold lower than that for 9-cis-RA (Fiorella et al., 1993) or too low to be measured (Redfem & Wilson, 1993; Fogh et al., 1993; Norris et al., 1994). However, CRABPs have high affinity for retinoids with a modified B-ionone ring such as 3,4-didehydroretinoic acid (Torma et al., 1991), 4-hydroxyretinoic acid, 4-oxo-retinoic acid, 18-hydroxyretinoic acid, 16-hydroxy-4-oxo-retinoic acid (Fiorella et al., 1993) and acitretin (Norris et al., 1994) (an aromatic retinoid used in the treatment of psoriasis). Both proteins also have high affinity for retinoids with modifications at C-7 and C-8 positions of the isoprene chain (Kleywegt et al., 1994). Although 3,4-didehydroretinoic acid exists in cells and binds CRABPs as tightly as all-trans-RA does, the latter has been shown to be the endogenous ligand for CRABPI and possibly CRABPII (Ross et al., 1980; Sarri et al., 1982). Taken together, these experimental data indicate that the 10 carboxylic group of RA is critical for binding to CRABPs. Modifications at the B-ionone ring and, C-7 and C-8 can be tolerated. Site-directed mutagenesis studies. The interactions between CRABPs and retinoids have been investigated by site-directed mutagenesis. The single mutants R1 1 IO and R131Q, double mutant RllleR13lQ and triple mutant RlllQ/R131Q/Yl33F of bovine/murine CRABPI all have much lower affinity for all-trans-RA than the wild-type (Zhang et al., 1992). The two arginines and one tyrosine are conserved residues in CRABPs and interact with the carboxyl group of RA in the crystal structures of hole CRABPs (Figure 1.2A, B) (Kleywegt et al., 1994). Although no quantitative K, has been reported for these mutants, a CD signal induced by the binding of RA has been observed in the R131Q mutant as in the wild type. No such signal has been detected in the mutant R1 1 IO even with a five-fold excess of all-trans-RA. It appears that in CRABPI Argl 11 is more critical for binding of RA than Argl31. By contrast, it has been reported that the R1 11A mutant of mouse CRABPII has a slightly lower affinity for all-trans-RA than the wild type, while either R132A or R132Q mutant has much lower affinity (Chen et al., 1995), suggesting that only Arg132 is important for binding of RA by CRABPII. However, no quantitative K, has been reported for either R132A or R132Q mutant. Interestingly, there is no change in the retinoid specificity of CRABPI or CRABPII because of the mutations. In particular, the mutants show no increase in affinity for all-trans-retinol and all-trans-retinal compared to their wild-types. In other words, neither CRABPI nor CRABPII could be converted into a retinol-binding protein by the single, double or triple mutations. Similarly, type-II cellular retinol binding protein (CRBPII), a close relative of CRABPs, could not be converted into a RA-binding protein 11 Figure 1.2 (A) The ligand binding pocket of human CRABPII showing the residues (white) interacting with all-trans-RA (green). The two hydrogen atoms (yellow) of a water molecule, the nitrogen atoms (blue) of the side chains, and the oxygen atoms (red) of the side chains and the RA molecule and the water are colored. (B) expanded View of the interactions between the carboxyl group of RA and the side chains of Argl 1 1, Arg132 and Tyr134 of the protein. Dotted line indicates the possible hydrogen bond interaction. 13 by site-directed mutagenesis. The two single mutants Q109R and Q129R (Gln109 and Gln129 of CRBPII are structurally equivalent to Arglll and Argl32 of CRABPII, respectively) display reduced affinity for all-trans-retinol but no apparent affinity for all- trans-RA (Cheng et al., 1991). Surprisingly, the mutant protein Q108R of rat CRBPI (Gln108 in rat CRBPI is structurally equivalent to Arglll of human CRABPII) has similar affinity for all-trans-retinol, all-trans-RA and 13-cis-RA. The mutation reduces the affinity for all-trans-retinol by only three-fold compared to the wild-type (Stump et al., 1991). Section 3. Structural and Dynamical Characterizations of CRABPs The atomic details of the interactions between CRABPs and retinoids, and their internal motions have been analyzed by physical methods, primarily the x-ray diffraction of single crystals and solution NMR spectroscopy. The structural and dynamical information provides a framework for understanding the nature of the interactions and the ligand binding process, and for identifying the factors contributing to the binding affinity and specificity. The structural features of CRABPs. The crystal structures of apo-CRABPI (Thompson et al., 1995), holo-CRABPI and holo-CRABPII (both in complex with all- trans-RA) (Kleywegt et al., 1994) have been determined at 2.7 A, 2.9 A and 1.8 A resolutions, respectively. These studies have demonstrated that the apo- and holo- l4 CRABPs have very similar secondary and tertiary structures. Moreover, they all have structural features typical of a protein family called intracellular lipid binding proteins (iLBPs) (Banaszak et al., 1994). These proteins all bind hydrophobic ligands such as fatty acid, lipids and retinoids. In addition to CRABPs, the crystal structures of more than twenty-five proteins of the iLBP family have been determined at various resolutions with or without bound ligands. The structures reported include rat CRBPI (Cowan et al., 1993) and CRBPII (Winter et al., 1993), several fatty acid binding proteins (FABPs) including those from chicken liver (Scapin et al., 1990), bovine heart (Muller-Fahmow et al., 1991), human muscle (Zanotti et al., 1992; Young et al., 1994) and rat intestine (Sacchettini et al., 1989a,b; Sacchettini et al., 1992; Eads et al., 1993), bovine myelin, (Jones et al., 1988; Cowan et al., 1993) and adipocyte lipid binding protein (ALBP) (Xu et al., 1992; 1993). Although the sequence identities among the proteins are rather low (~20%), they all have similar tertiary structures. All the proteins adopt a compact, single domain structure consisting of a helix-tum-helix motif and two antiparallel B-sheets (Figure 1.3). Both sheets are composed of five B-strands, the first BA, BB, BC, BD and BE, and the second BF, BG, BH, BI and BJ. They are twisted into a nearly orthogonal flattened barrel with a simple up-down topology. The regular hydrogen bond ladders between adjacent B- strands are broken between BD and BE. One B-bulge occurs in the middle of the first strand, resulting in a change of its direction so that it forms parts of the two sheets. BF is also shared by the two sheets. A binding cavity is formed between the two B-sheets by the side chains from eight of the ten B-strands. The surface of the cavity could be divided into two regions: a hydrophobic region and a more centrally located polar region (Figure 1.2A). The former is formed by the side chains from the helix-tum-helix, the BC-BD loop 15 Figure 1.3. The backbone fold and the secondary structural elements of human CRABPII in complex with all-trans-RA (green) showing the structural motif of iLBP family: ten antiparallel B—strands (yellow) and two a. helices (violet) fortning a flattened barrel. The loops or turns are colored in cyan. 17 and the BE-BF loop. The latter is formed by the side chains of the residues deep inside the binding pocket such as Argl l l, Arg132 and Tyrl34 (CRABPII numbering). Generally, only a small opening to the external solvent is observed in the crystal structures, which is supposed to be the ligand entrance. The rear portion of the pocket, opposite to the opening, is completely blocked and inaccessible to solvent. The ligand is either totally buried inside the protein as observed in the majority of the crystal structures or its hydrophobic tail extends to the surface through the ligand entrance as in the crystal structures of CRABPs and ALBP (Xu et al., 1992). The polar group of the ligand is bound in the innermost part of the pocket. Interestingly, in most structures, the volume of the binding pocket is much larger than what is required by the ligand. The extra volume may provide room for the fluctuation of ligand in the bound state. The structural basis for affinity and specificity of CRABPs. The high-resolution structures of the holo-CRABPs provide a basis for the interpretation of their binding affinity and specificity in terms of the relative positions of atoms. The binding specificity and affinity of iLBPs can be partially explained by variation in the positions and the nature of amino acid residues, and by the tightly bound solvent molecules in the binding pocket. Particularly, the interaction between the polar end of the ligand and the protein, (Banazak et al., 1994), seems to be most critical. In the crystal structure of holo- CRABPII, one of the carboxylate oxygens of RA forms hydrogen bonds with the hydroxyl of Tyrl34 and the guanidinium group of Arg132. The other interacts with the guanidinium group of Arglll mediated by a water molecule (Figure 1.2B). Similar interactions between the carboxyl group and the two arginine and one tyrosine resiudes have been observed in other iLBPs. The importance of such interactions for binding of 18 the ligands has been verified by site-directed mutagenesis studies. As described earlier, mutation of these residues to other amino acids reduces greatly the binding affinities of iLBPs. In the structures of holo-CRABPs extensive hydrophobic interactions have been observed between the isoprene chain of RA and CRABPs (Figure 1.2A). Furthermore, a network of water molecules in the binding pocket is often detected in the high-resolution crystal structures. For example, fourteen water molecules have been identified in the crystal structure of holo-CRABPII (Kleywegt et al., 1994). Eight water molecules have been identified in the crystal structure of holo-CRBPI (Cowan et al., 1993) and holo- CRBPII (Winter et al., 1993). They have been suggested to be an integral part of the binding pockets (LaLonde et al., 1994) and to be related to the binding affinity and specificity of these proteins. Some clues on the binding specificity of CRABPs can be drawn from the comparison of the amino acid sequences and the structures of iLBPs. For example, the two arginines and one tyrosine of CRABPs and FABPs interacting with the carboxyl groups of their ligands are replaced, respectively, by two glutamines and one phenylalanine in CRBPs. And, the arginine-carboxyl interactions in CRABPs and FABPs are substituted by the interactions between the two glutamines and the hydroxyl group of retinol. Indeed, replacement of Gln108 of rat CRBPI with arginine has changed the specificity of the protein so that it binds all-trans-RA and 13-cis-RA in addition to all- trans-retinol (Stump et al., 1991). Mechanisms of ligand entry or exit. Previous structural studies, mainly by means of x-ray diffraction, have shown that most of the proteins of the iLBP family undergo 19 only small changes upon ligand binding, and there exists no entrance large enough for the entry or exit of the ligand. In the holo crystal structures of CRBPI, CRBPII, intestinal FABP and ALBP, the ligands are completely buried. The ligand entrances of the unliganded forms are almost identical to those of the holo-forms. The crystal structures of CRABPs show that RA is not completely buried (Kleywegt et al., 1994). The edge of the B-ionone ring of bound RA is exposed to solvent. It appears, however, that significant conformational changes are required for release of the ligand. The ligand entrance of apo- CRABPI is slightly more open than that of holo-CRABPI, and appears to be large enough to admit RA (Thompson et al., 1995). Surprisingly, apo-CRABPI is dirneric in the crystalline state, held together by an intermolecular B-sheet. The slight opening of the ligand entrance has been suggested to be caused by formation of the intermolecular B- sheet. Biochemical studies, by contrast, have suggested that ligand binding induces ccnforrnational and dynamical changes in iLBPs. It has been shown that a microsomal retinol esterifying enzyme, lecithin-retinol acyltransferase, can discriminate between the apo- and the holo-CRBPs (Herr & Ong, 1992). It has also been demonstrated that ligand binding reduces the protease susceptibility of the ligand entrances of CRBPI, CRBPII, CRABPI and heart FABP (Jamison et al., 1994). The results suggest that the apo proteins have more open binding pockets or their entrances are more flexible in solution. These biochemical results are consistent with the physiological functions of these proteins, which require that there exist an entrance large enough for ligand admittance, at least temporarily. Furthermore, if holo-CRABPs and holo—CRBPs are the substrates for metabolic enzymes (Herr & Ong, 1992; Fiorella & Napoli, 1991, 1994; Boylan & Gudas, 20 1992), there should exist structural features that can be used by the enzyme to distinguish the apo-forms from the hole-forms. Indeed, lecithin-retinol acyltransferase can distinguish the apo and holo forms of CRBPs (Herr & Ong, 1992). NMR studies of the iLBP family. NMR spectroscopy has the advantage over x- ray diffraction in providing dynamical information. Such information is critical for understanding the ligand binding process. 1H and 15N resonance assignments have been reported for both apo-CRABPI (pH 7.5) and holo-CRABPI (pH 3.8) (Rizo et al., 1994). Secondary structures derived from NOE and chemical shift indices indicate that both the apo- and the holo-CRABPI share the backbone fold of the iLBP family: ten B—strands and a helix-tum-helix. However, significant differences in internal motions between the apo- and holo-forms have been observed in the ligand entrance region. The ligand entrance of apo-CRABPI is rather flexible in solution, and the flexibility is reduced greatly after the binding of RA. Remarkable motional differences in the ligand entrance region have been reported between rat apo and holo intestinal FABP at pH 7.2 (Hodsdon et al., 1995; Hodsdon et al., 1996; Hodsdon & Cistola, 1997a, b). Furthermore, it has been proposed that the decrease in the mobility of the ligand entrance in the holo form is brought about by a series of long-range cooperative interactions that cap and stabilize the C-terminal half of the second helix. In contrast to what was observed in the crystal structure, the ligand entrance of bovine heart FABP is more mobile than the other parts of the molecule, even in the holo form (Lucke et al., 1992; Lassen et al., 1995). Proton resonance heterogeneity has been observed for Ser22-Thr36 in the helix-tum-helix, Thr53-Thr60 in the BC—BD loop and His] l9-Ala122, indicating that these residues have two or more slowly exchanging conformations in solution. 2] NMR studies of the bound RA. The conformation of CRABP-bound RA has been studied by NMR (Norris et al., 1995). The NMR results indicate that the bound RA in CRABPII adopts a conformation with the 6-3 dihedral angle being -60° skewed from a cis conformation, in contrast to the -33° observed in the crystal structure of holo-CRABPII (Kleywegt et al., 1994). However, RA in holo-CRABPI appears to be quite flexible about the 6-8 bond. The 8-3 and lO-s dihedral angles are 180° i 30° in both holo-CRABPI and holo-CRABPII in solution as well as in the crystalline state. 22 References Astrom, A., Tavakkol, A., Petersson., U., Cromie, M., Elder, J. T., & Voorhees, J. J. (1991).]. Biol. Chem. 266, 17662-17666. Bailey, J. S., & Siu, C. H. (1988) J. Biol. Chem. 263, 9326-9332. Banaszak, L., Winter, N., Xu, 2., Bernlohr, D. A., Cowan, S. W., & Jones, T. A. (1994) Adv. Protein Chem. 45, 89-151. Blomhoff, R., Green, M. H., Berg, T., & Norum, K. R. (1992) Science 250, 399-404. Boylan, J. F ., & Gudas, L. J. (1991) J. Cell. Biol. 112, 965-978. Boylan, J. F., & Gudas, L. J. (1992) J. Biol. Chem. 267, 21486-21491. Bucco, R. A., Zheng, W., Davis, J. T., Sierra-Rivera, E., Osteen, K. G., Chaudhary, A. K., & Ong, D. E. (1997) Biochemistry 36, 4009-4014. Chambon, P. (1996) FASEB. J 10, 940-954. Chambon, P. (1994) Sem. Cell Biol. 5, 115-125. Chambon, P., Olson, J. A., & Ross, A. C. (coordinators) (1996) The retinoid revolution. FASEB. J 10, 938-1048. Chen, L. X., Zhang, Z. -P., Scafonas, A., Cavalli, R. C., Gabriel, J. L., Soprano, K. J ., & Soprano, D. R. (1995) J. Biol. Chem. 270, 4518-4525. Cheng, L., Qian, S-J., Rotthschild, C., d’Avignon, A., Lefkowith, J. B., Gordon, J. 1., & Li, E. (1991) J. Biol. Chem. 266, 24404-24412. Cogan, U., Kopelman, M., Mokady, 8., Shinitzky, M. (1976) Eur. J. Biochem. 65, 71-78. 23 Cowan, S. W., Newcomer, M. E., & Jones, A. (1993) J. Mol. Biol. 230, 1225-1240. Daly, A. K., & Redfem, C. P. F. (1988) Biochim. Biophys. Acta 985, 118-126. de Bruijn, D. R., Oerlemans, F ., Hendriks, W., Baats, E., Ploemacher, R., Wieringa, B., & Geurts van Kessel, A. Differentiation 58, 141-148. Durand, B., Saunders, M., Leroy, P., Leid, M., & Chambon, P. (1992) Cell 71, 73-85. Eads, J., Sacchettini, J. C., Kromminga, A., Gordan, J. I. (1993) J. Biol. Chem. 268, 26375-26385. Eller, M. S., Oleksiak, M. F ., McQuaid, T. J., McAfee, S. G., & Gilchrest, B. A. (1992) Exp. Cell Res. 199, 328-336. Evans, R. M. (1988) Science 240, 889-895. Fiorella, P. D., Giguere, V., & Napoli, J. L. (1993).]. Biol. Chem. 268, 21545-21552. Fiorella, P. D., & Napoli, J. L. (1991).]. Biol. Chem. 266, 16572-16579. Fiorella, P. D., & Napoli, J. L. (1994) J. Biol. Chem. 269, 10538-10544. Fogh, K., Voorhees, J. J ., & Astrom, A. (1993) Arch. Biochem. Biophys. 300, 751-755. Giguere, V., Lyn, s., Yip, P., Siu, C-H., & Amin, s. (1990) Proc. Natl. Acad. Sci. 87, 6233-6237. Giguere, V., Ong, E. S., Segui, P., & Evans, R. M. (1987) Nature 330, 624-629. Herr, F. M., & Ong, D. E. (1992) Biochemistry 31, 6748-6755. Heyrnan, R. A., Mangelsdorf, D. J ., Dyck, J. A., Stein, R. B., Eichele, G., Evans, R. M. & Thaller, C. (1992) Cell 68, 397-406. Hodson, M. E., & Cistola, D. P. (1997a) Biochemistry 36, 1450-1460. Hodson, M. E., & Cistola, D. P. (1997b) Biochemistry 36, 2278-2290. Hodson, M. E., Ponder, J. W., & Cistola, D. P. (1996) J. Mol. Biol. 264, 585-602. 24 Hodson, M. E., Toner, J. J., & Cistola, D. P. (1995) J. Biomol. NMR 6, 198-210. Hofmannn, C., & Eichele, G. (1994) In The Retinoids: Biology, Chemistry, and Medicine (Sporn, M. B., Roberts, A. B., & Goodman, D. S., eds), pp. 387-441, 2“d ed., Raven Press, New York. Hong, W. K., & Itri, L. M. (1994) In The Retinoids: Biology, Chemistry, and Medicine (Spom, M. B., Roberts, A. B., & Goodman, D. S., eds.), pp. 597-630, 2“d ed., Raven Press, New York. Jamison, R. S., Newcomer, M. E., & Ong, D. E. (1994) Biochemistry 33, 2873-2879. Jing, Y., Waxrnan, S., & Mira-y-Lopez, R. (1997) Cancer Res. 57, 1668-1672. Jones, T. A., Bergfos, T., Sedzik, J., & Unge, T. (1989) EMBO J. 7, 1597-1604. Kitomoto, T., Momoi, T., & Momoi, M. (1988) Biochem. Biophys. Res. Commun. 157, 1302-1308. Kleywegt, G. J., Bergfors, T., Senn, H., 1e Motte, P., Gsell, B., Shuao, K., & Jones, T. A. (1994) Structure 2, 1241-1258. LaLonde, J. M., Bernlohr, D. A., & Banaszak, L. J. (1994) Biochemistry 33, 4885-4895. Lassen, D., Lucke, C., Kveder M., Mesgarzadeh, A., Schrrridt, J. M., Specht, B., Lezius, A., Spener, F., & Ruterjans, H. (1995) Eur. J. Biochem. 230, 266-280. Levin, A. A., Sturzenbecker, L. J ., Kazrner, S., Bosakowski, T., Huselton, C., Allenby, G., Speck, J ., Kratzeisen, C., Rosenberger, M., Lovey, A., & Grippo, J. F. (1992) Nature 355, 359-361. Lucke, C., Lassen, D., Kreienkamp, H.-J., Spener, F., & Ruterjans, H. (1992) Eur. J. Biochem. 210, 901-910. Mangelsdorf, D. J., & Evans, R. M. (1995) Cell 83, 841-850. 25 Mangelsdorf, D. J ., Ong, E. S., Dyck, J. A., & Evans, R. M. (1990) Nature 345, 224-229. Mangelsdorf, D. J ., Umesono, K., & Evans, R. M. (1994) The retinoid receptors. In The Retinoids: Biology, Chemistry, and Medicine (Spom, M. B., Roberts, A. B., & Goodman, D. S., eds.), pp. 319-349, 2"d ed., Raven Press, New York. Maden, M. (1991) Dev. Biol. 2, 161-170. Maden, M., Ong, D. E., Surnmerbell, D., & Chytil, F. (1988) Nature 335, 733-735. Melhus, H., Gobl, A., & Ljunghall, S. (1994) Biochem. Biophys. Res. Comm. 200, 1125- 1129. Morriss-Kay, G. M., & Sokolova, N. (1996) FASEB. J 10, 961-968. Muller-Fahmow, A., Egner, U., Jones, A., Ruedel, H., Spener, F., & Saenger, W. (1991) Eur. J. Biochem. 199, 271-276. Nilsson, M. H., Spurr, N. K., Saksena, P., Busch, C., Norlinder, H., Peterson, P. A., Rask, L., & Sundelin, J. (1988) Eur. J. Biochem. I 73, 45-51. Norris, A. W., Cheng, L., Giguere V., Rosenberger, M., & Li, E. (1994) Biochim. Biophys. Acta 1209, 10-18. Norris, A. W., Rong, D., d’Avignon, D. A., Rosenberger, M., Tasaki, K., & Li. B. (1995) Biochemistry 34, 15564-15573. Ong, D. E., & Chytil, F. (1975) J. Biol. Chem. 250, 6113-6117. Ong, D. E., & Chytil, F. (1978) J. Biol. Chem. 253, 4551-4554. Ong, D. E., Newcomber, M. E., & Chytil, F. (1994) In The Retinoids: Biology, Chemistry, and Medicine (Spom, M. B., Roberts, A. B., & Goodman, D. S., eds), pp. 283-317, 2““ ed., Raven Press, New York. Peck, G. L., & DiGiovanna, J. J. In The Retinoids: Biology, Chemistry, and Medicine. 26 (Spom, M. B., Roberts, A. B., & Goodman, D. S., eds.), pp. 631-658, 2"d ed., Raven Press, New York. Petkovich, M., Brand, N. J ., Krust, A., & Chambon, P. (1987) Nature 330, 444-450. Rizo, J., Liu, Z. P., & Gierasch, L. M. (1994) J. Biomol. NMR 4, 741-760. Ross, A. C., Adachi, N., & Goodman, D. S. (1980) J. Lipid Res. 21, 100-109. Sacchettini, J. C., Gordan, J. 1., & Banaszak, L. J. (1988).]. Biol. Chem. 263, 5815-5819. Sacchettini, J. C., Gordan, J. I., & Banaszak, L. J. (1989a) Proc. Natl. Acad. Sci. U. S. A. 86, 7736-7740. Sacchettini, J. C., Gordan, J. I., & Banaszak, L. J. (1989b) J. Mol. Biol. 208, 327-339. Sacchettini, J. C., Scapin, G., Gopaul, D., & Gordan, J. I. (1992) J. Biol. Chem. 267, 23534-23545. Sani, B. P., & Hill, D. L. (1974) Biochem. Biophys. Res. Commun. 61, 1276-1282. Sarri, J. C. ( 1982) Methods Enzymol. 81, 819-826. Scapin, G., Spadon, P., Marnmi, M., Zanotti, G., & Monaco, H. L. (1990) Mol. Cell. Biochem. 98, 95-99. Shubeita, H. E., Sambrook, J. F., & McCormick, A. M. (1987) Proc. Natl. Acad. Sci. U. S. A. 84, 5645-5649. Siegenthaler, G., Tomatis, I., Chatellard-Gruaz, D., Jaconi, S., Erikson, U., & Saurat, J. H. (1992) Biochem. J. 287, 383-389. Sporn, M. B., Roberts, A. B., & Goodman, D. S. (1994) The Retinoids: Biology, Chemistry, and Medicine, 2“d ed., Raven Press, New York. Stoner, C. M., & Gudas, L. J. (1989) Cancer Res. 49, 1497-1504. Stump, D. G., Lloyd, R. S., & Chytil, F. (1991) J. Biol. Chem. 266, 4622-4630. 27 Szuts, E. Z., & Harosi, F. I. (1991) Arch. Biochem. Biophys. 287, 297-304. Takase, S., Ong, D. E., & Chytil, F. (1986) Arch. Biochem. Biophys. 24 7, 328-334. Thompson, J. R., Bratt, J. M., & Banaszak, L. J. (1995).]. Mol. Biol. 252, 433-446. Torma, H., Stenstrom, E., Andersson, E., & Vahlquist, A. (1991) Skin Pharmacol. 4, 246-253. Tsai, M- J ., & O’Malley, B. W. (1994) Ann. Rev. Biochem. 63, 451-486. Wald, G. (1968) Science 162, 230-239. Wang, L., Li, Y., & Yan, H. (1997) J. Biol. Chem. 272, 1541-1547. Winter, N., Bratt, J., & Banaszak, L. J. (1993) J. Mol. Biol. 230, 1247-1259. Xu, 2., Bernlohr, D. A., & Banaszak, L. J. (1992) Biochemistry 31, 3484-3492. Xu, Z., Bernlohr, D. A., & Banaszak, L. J. (1993).]. Biol. Chem. 268, 7874-7884. Young, A. C. M., Scapin, G., Kromminga, A., Patel, A. B., Veerkamp, J. H., & Sacchettini, J. C. (1994) Structure 2, 525-534. Yu, V. C., Delsert, 0, Andersen, B., Holloway, J. M., Devary, O. V., Naar, A. M., Kim, S. Y., Boutin, J. M., Glass, C. K., & Rosenfeld, M. G. (1991) Cell 67, 1251-1266. Zanotti, G., Scapin, G., Spadon, P., Veerkamp, J. H., & Sacchettini, J. C. (1992) J. Biol. Chem. 267, 18541-18550. Zhang, J., Liu, Z., Jones, T. A., Gierasch, L. M., & Sambrook, J. F. (1992) Proteins: Structure, Function, and Genetics 13, 87-99. Part 2. Physical Basis of Protein NMR Section 1. Theory of Fourier Transform Pulse NMR Nuclear magnetic resonance (NMR) spectroscopy, the detection of the Zeeman A levels of nuclear spins in bulk condensed phase by resonant method through electromagnetic means, was pioneered by two research groups headed by Felix Bloch (Bloch et al., 1946a) and Edward M. Purcell (Purcell et al., 1946), respectively. It has since become an important experimental tool for physicists, chemists and more recently for structural biologists as well as radiologists (medical diagnosis). Its wide applications are made possible by continuous improvement in experimental techniques and advances in theory. The former is stimulated by technological breakthroughs in other fields such as electronics, computer science (e. g. fast Fourier transformation algorithm and the sophisticated controls of experimental parameters), and the manufacturing of stable high field superconducting magnets. The latter takes advantage of the weakness of the interactions involving nuclear spins. The weakness makes it relatively easy to manipulate the interaction Hamiltonian through experimental methods such as radio frequency pulses, and justifies the treatment of the evolution of the states of a spin system by perturbation methods. The simplicity of the derived mathematical expressions facilitates 28 29 study biological macromolecules. In this review I will summarize the underlying physical principles of NMR, and will show how they can be applied to practical problems especially those in structural biology. Basic equations. Nuclear spin, a quantum mechanical phenomenon without classical analogs, can only be treated by quantum mechanics. However, for an isolated system consisting of non-interacting spins, the equation for the expectation value of the total magnetization is the same as that for the classical magnetization under the influences of a static field HO and a rotating external radio-frequency field Hl (r. f. field) (Rabi etal., 1954). d/dt=y me (l) where Heff is the total magnetic field in a frame rotating with the frequency of the r. f. field, y is the gyromagnetic ratio. Therefore, it is legitimate to treat such a system classically. In reality the spin system is not isolated but interacts with the molecular environment (bath) including all degrees of freedom other than the spins. The interaction causes a decay of the observed magnetization. With the assumptions that the decay is exponential, and the interactions with the external fields and with the bath are additive, Bloch derived a phenomenological equation for a non-interacting spin system (Bloch, 1946b). dM/dt = y M x H — MVT, — k (M,— M,)/ T, (2) 30 where H is the sum of a static field H0 = kl!O and a r. f. field H], k is the unit vector on the z-axis, M0 = XoHo is the equilibrium magnetization with x0 being the static susceptibility, M is the total magnetization, M is the transverse magnetization. T, and T 2 are, respectively, the longitudinal and transverse relaxation times representing the interactions between the spin system and the bath. The nuclear spins in a real spin system interact with each other. The descriptions of such a spin system must resort to quantum mechanism. For convenience, the Hamiltonian H of the conserved whole system is divided into three parts: the Hamiltonian H,(s, t) of the spin system including the interactions with the external static and r. f. fields, the Hamiltonian F (f) of the bath and the interaction Hamiltonian GO? 3, t) between them. The state of a spin system can be represented by a reduced density operator o(t) (von-Neumann, 1927, 1955; Fano, 1957), which could be defined as an incoherent superposition of pure states. The dynamical evolution of o(t) under the influence of the spin Hamiltonian H,(t) follows the Liouville- von Neumann equation. 50/5t = -i [H.(t), 60)] (3) o(t) is simply related to the density operator p(t) of the whole system: o(t) = tr,{p(t)}, here tr, means the partial trace over the variables of the bath. The expectation value of any spin operator Q(t) can be calculated according to = tr,{Q(t)o}, here tr, means the partial trace over the variables of the spin system. The separation of the whole system into two subsystems (the spin system and bath) is very helpful in the 31 development of NMR theory because of the weakness of the interaction between the spin system and the bath. For example, in the operator product formalism discussed later, an exponential term multiplied into a spin operator can be used to represent the effect of relaxation on the spin system caused by the bath. In relaxation theory, irreversible processes are assumed to occur exclusively in the spin system though only reversible processes could happen in the conserved whole system. Pulse NMR. For most pulse NMR experiments, the exact analytic solutions of the Liouville-von Neumann equation (3) are very difficult to find because of the complexities of the interactions among the spin system itself, and between the spin system and its environment. To make the problem tractable and provide useful qualitative insights into NMR experiments, various approximations have been proposed to simplify the Harniltonians and/or the reduced density operator depending on the nature of the practical system under investigation. One of the simplified versions of the density operator formalism called operator product formalism (Banwell & Primas, 1962; Sorensen et al., 1983; Van de Ven & Hibers, 1983) has been frequently employed to describe the experiments performed on a weakly coupled spin system in the context of high resolution NMR. Following a suggestion by Fano (F ano, 1957), the main idea behind this formalism is to expand the o(t) in a tensor base constructed from the products of spin operators (E, 6,, 0y, 0', of the Pauli matrices) (Sorensen et al., 1983; Ernst et al., 1987). o(t) = 2, b,(t) B, and B, = 2‘4-’>1_[(1,,)“s* (4) k=l 32 where B, is a base operator in the Liouville space consisting of all spin operators, N = the total number of I = '/2 nuclei in the spin system, k = index of nucleus, v = x, y, or z, q = number of single-spin operators in the product, as, = 1 for q nuclei and as, = 0 for the N-q remaining nuclei. The spin Hamiltonian H, of a weakly coupled spin system in the context of high resolution NMR in liquid phase could be approximated by 0,1,2, MHZ] ,, and B1,“, (here B is the flip angle of per unit time about v axis), respectively. The first two are time-independent, representing the interaction with the static field (the chemical shifls) and the interaction between nuclear spins (spin-spin coupling), respectively. The other is a time-dependent term, representing the interaction with external r. f. fields (pulses), which could be made time-independent for a finite time segment by selecting a suitable rotating frame (Rabi et al., 1954). The solution of the equation (3) for a time- independent propagator U is very simple (3 (t) = exp(—iUt) 6(0) exp(+iUt) (5) Most of the NMR experiments could be calculated by means of this equation and commutation relations. With the simplified Harniltonians mentioned above the we could derive all the results from operator product formalism. However, the physical picture was more clear in the latter, whose advantage originates from the simplification of the spin Hamiltonian H, such that each of the three terms is itself a base operator. Therefore, the evolution of any product operator under the effect of any of the three terms could be 33 calculated according to the commutation rules among them. If there exists the following commutation relation among operators A, B and C (and their cyclic permutations) [A, B] = iC (6) the operator A under the propagator U = H, = C will evolve as exp(—i0C) A exp(+i0C) = A cost) + B sin0 (7) Therefore, the space spanned by the three independent operators A, B and C is a subspace of the Liouville space. The majority of pulse NMR experiments applied to the structural biological problems, where the H, could be approximated by the three terms above, could be easily treated in this manner. However, this formalism is not adequate for a complete mathematical description of NMR relaxation in liquid. Fourier Transform NMR. The method of recording NMR signals free of pulse was widely used in the T , and T 2 measurements by the spin-echo method (Hahn, 1950; Carr & Purcell, 1954; Meiboom & Gill, 1958). A non-echo method has also been developed to record NMR signals free of pulse by means of Fourier transform. It has been shown that after applying a strong external r. f. field H, for a very short time 1: (r. f. pulse) the transverse magnetization M (t) = M, + i My evolves in the absence of a pulse according to (Lowe & Norberg, 1957) M (t) = M,, sin (1),“: exp(iroot) I f(a),, + a) ) exp(—i(ot) do) (8) L 4 Git) 34 where M0 is the equilibrium magnetization, (0,, is the central Larmor frequency, a), is the frequency of the applied r. f. field, f(a),, + (o) is the shape function related to the imaginary part of the r. f. susceptibility x". The integral is from —oo to +00. M (t) processes with a time-dependent amplitude G(t) called correlation function, which is the Fourier transform of the shape function f((o(, + 03). Therefore, the recording of the decay of a NMR signal (called free induction decay) represented by G(t) gives the same information about the shape function as the observation of the resonance with a vanishing small r. f. field (the continuous wave NMR). The superiority of the Fourier transform NMR over the continuous wave NMR becomes obvious when it was applied to complex systems (Ernst & Anderson, 1966). One of the advantages is the sensitivity improvement, which becomes more apparent for complicated high resolution spectra in liquid. Furthermore, Fourier transformation consists in the basis for extending the dimensionality of NMR experiments. Multidimensional NMR. Two-dimensional (2D) NMR experiments, first proposed by Jeener (Jeener, 1971), are natural extensions of double resonance experiments (Bloch, 1954, 1958; Royden, 1954; Bloom & Schoolery, 1955). Both techniques are designed to observe the properties of a system, which are related to its nonlinear responses to external perturbations such as the connectivity of transitions in various energy levels. In contrast to the normal double resonance techniques, the parameter t, (time in the evolution period) in 2D techniques is changed systematically as a variable, and a Fourier transform is applied to both time variables t, and t2 (Aue et al., 1976) 35 S ((01, (02) = I dtl exp(—iw,t,) l dtz [exp(-io)2t2)] 5+0], ’2) (9) S+(tl' t2) = trs {PO-(’1' t2)}=zrs Zruexp{(“i(”nM)—A73“!))t2} er tu exp{(_iwm(C)—)\'tu(6))tll (10) 0 (to ’2) = €XP{-(i"7/§d’ + F,(d))t_,} WeXPI-(it‘y/J” + 11“”)0} '4?00 (11) where S((o,, 0),.) represents the cross-peaks of the 2D spectrum, s+(t,, t2) is the observed complex magnetization signal (FID) and F = Z l,+i1,, is the observable spin operator. c’ 4“”, F ‘d’ and 6%,”, 1“,“) are the Hamiltonian and relaxation superoperators in the 3 detection and evolution periods, respectively. Z,, W = F, .% (.9f00)m, here .Jfoo is the rs in density operator after the preparation period, .44 is the preparation superoperator and <50 is the density operator at t = 0. .A’is a rotation superoperator representing the r. f. pulses, which couples the coherences in the evolution period with those in the detection period. In other words, .%induces coherence transfers which are representative of the molecular system under investigation, and is the characteristic feature distinguishing the various forms of 2D experiments. In three- and four-dimensional experiments, there may be two or three evolution periods, and two or three rotational superoperator .Ws transferring the coherences between the different evolution periods and between the last evolution period and the final detection period. Selection of pathway. For the success of a 2D experiment, certain desired coherence transfer pathways must be selected while others are suppressed by either phase-cycling (Wokaun & Ernst, 1977) or pulsed field gradients (Maudsley et a1, 1978; Bax et al., 1980). Both methods rely on the different transformational behaviors of the 36 coherences with different orders under phase-shifting or gradients. In NMR experiments coherences are transferred between the different orders by a propagator U((p) (5), which changes with its phase (p in the following manner. U((+i(-ip,, (29) where angular bracket represents ensemble average. From the exact Liouville-von Neumann equation (3), several authors have derived relaxation theories in terms of the correlation function or spectral density of the molecular environment (bath) appropriate for a spin system with interactions such as scalar, dipolar and quadrupolar couplings (Wangsness & Bloch, 1953; Fano, 1953, Bloch, 1956, 1957; Redfield, 1958, 1965; Abragam, 1961; Hubbard, 1961; Argyres & Kelley, 1964). These theories share features similar to the general Onsager’s thermodynamic theory on irreversible processes (Zwanzig, 1961), and are constructed by means of various simplifying assumptions on the bath and the magnitude of the coupling between the bath and the spin system. Their final results are a set of equations relating the NMR- measurable relaxation parameters such as relaxation time and NOE enhancement to the microscopic properties characteristic of the bath. The key steps of the derivations are recapitulated as follows. Quantum relaxation theory in operator form. An isolated system with total conserved Hamiltonian H is divided into two subsystems, a bath with time-independent Hamiltonian hF(/) and a spin system with time-dependent Hamiltonian hH,(s, t) including 51 the interactions with external r. f. fields (Hubbard, 1961). These two subsystems interact with each other and the interaction Hamiltonian 1560? s, t) is a random function of time. H = 2H,(s, t) +2; F(f)+ hG(/f s, t) (30) The separation is necessary since the two subsystems possess quite different properties and play distinct roles in NMR relaxation theory. Several simplifying assumptions are introduced for the bath at the different stages of the derivation. The bath is assumed to possess quasi-continuous energy levels, and to be in thermal equilibrium at all times and independent of the spin system. p(fi s ,t) = o(s, t) pT(/) = G(S, t) exp(—/iF / k7) / tr,.{exp(—/iF / kT )} (31) where p(fl s ,t), pT(f) and o(s, t) are density operators for the whole system, the bath in thermal equilibrium and the spin system, respectively. T is the absolute temperature, k is the Boltzmann constant and tr, means taking the trace over an ensemble of the bath subsystems. These assumptions characterize the basic feature of a relaxation process: irreversibility. With these assumptions, a syrnmetrized quantum-mechanical correlation function C,,(r ) can be defined as Cu“ ) = .tv = '/2 [trfiprplt + I) F'(t)} + trle‘lt + 09710)] (32) 52 where F’s are the variables of the bath whose dissipative behavior is rather loosely characterized by a temperature-dependent correlation time t, with the property that |C,,,(t ) I z 0 if I >> rc >>/r‘/ kT. It is further assumed that the bath correlation time t, is much shorter than the relaxation time of the spin system, that is, |R|‘I or [N]-1 >> r,, where l R I"1 or | N l’I is a function of both the spin operators and the spectral densities of the bath, the latter is defined as J“ (to) = l C,, (t) exp(iwt) dt {see equation (75) and (76) of Hubbard, 1961}. The magnitude of lRl‘l or lM" represents a measure of the relaxation times of the spin system. Two assumptions are also made for the interaction Hamiltonian hG(s, f, t). It is assumed to be so weak that the relaxation of a spin system caused by it can be treated by time-dependent perturbation up to the second order. ao/at = - tr, {i [60). pm] +1.;dr [G(t). [Ga—r). mom} (33) Furthermore, G(t) is expanded as a sum of products of a bath operator F‘(t) and a spin operator V(t) with the latter itself being expanded as a Fourier series or integral G(t) = z , 15*(z) Va) and V0) = 2 , V,‘ exp (rm/‘1) (34) where 0)," is the difference of the energy levels in frequency of the spin system. With these assumptions the change in o(s, t) for a time interval At satisfying the condition lRl’l or [N |‘1 >> At >> I, could be expressed as 53 o(t+At) = 6(1) + if“ {R (G(t), t') + i [G(t), N (t’)]} dt’ (35) Finally, with the assumption that G(t) in (34) can be replaced by G(t’) (this is equal to the introduction of the irreversibility into the problem according to Fano), the integrodifferential equation (35) becomes a linear differential equation. 66(0/ 6: = i [G(t), N (t)]+ R (G(t), t) (36) This approximation implies that the response of the spin system to an external perturbation is simultaneous without any memory effect (Zwanzig, 1961). Hubbard argued that this assumption was the same as the assumption |R|", WI“l >> At >> t,. In order to take advantage of the symmetrical properties of the bilinear dipolar- dipolar, quadrupolar and chemical shift anisotropy (CSA) interactions G(t) could be expanded as a scalar contraction of second-rank spherical irreducible tensor operators 7", of the spin system and F“,(t) of the molecular degrees of the bath (see, e. g. Brink & Satchler, 1993; Goldman, 1984) G(t) = 2M (—1)" I“, F“_q(t) (37) where u specifies different interactions and the summation q runs from —1 to +1, here 1 represents the rank of the tensor, in this case, I = 2. The corresponding correlation function involving only F",(t) can be defined similarly. This expansion makes it much 54 easier to treat the cross-correlation between different relaxation mechanisms but with the same rank (Werbelow & Grant, 1977; Vold & Vold, 1978). The cross correlation between different ranks vanishes irrespective of the value of q. Semiclassical theory. In semiclassical approximation the bath is further simplified to act as a randomly fluctuating magnetic field with infinite temperature. The classical bath correlation function is identified with the symmetrized quantum-mechanical correlation function (31) without regard to the commutation relation and the density operator of the bath pT(/). Cum = tr.{F*‘.. (38) The semiclassical formalism widely used in the interpretation of protein NMR relaxation data is due to Abragam (Abragam, 1961). In an interaction representation with o’(t) = tr, {p'(t)} = tr, {exp(i H,+F) p exp(—iH,—F)} equation (32) and (34) can be simplified as 0’0) = exp(iH.+F) G(t) exp(-iH.-F) = Z). 1E*(t)A"(t) = Z]. F0) Vrkexpfiwrkt) (39) 6676i = —1/2 2,, ,J,(m,*) [V,"‘,[V,", o‘— 0T1] (40) where spectral density J,,(oo,") = l exp(ico,"t) dt is defined as before, and k"(t) is a spin operator. oTrepresents the reduced density operator in a thermal equilibrium state. There is a difference between the H,(s, t) in Hubbard’s notation and the H,(s, t) in Abragam’s notation, namely, the former includes the average over an ensemble of baths in thermal equilibrium while the latter does not, so no 0’ appears in (36). The cross- 55 correlation between different dipolar-dipolar interactions, and between the dipolar and chemical shift anisotropy (CSA) is neglected in (40). They are supposed to be suppressed experimentally (Kay et al., 1992b). Usually T, and T , relaxation and heteronuclear NOE enhancement o(H,—)N,) such as that between 'H and ”N are dominated by the dipolar-dipolar interaction between ”N and the attached proton, as well as the CSA interaction with an external magnetic field. These three relaxation parameters can be related to spectral densities J((o)’s at four fi'equencies (Abragam, 1961) UT] = 72H Yzth /4r6NH {10)(wH-(DN) 1' 3fl)(mN) + 6J2)((DH+CDN) } 1" 1/3A2 OJZNJCCUDN) ...... (41) 1 / T 2 = 72 H 11214712 /8r6~H {4J°’(0)+J‘°’((DH-co~) + 3.1"(69~) + 6J‘ ”((011) +6J‘Z’(wn+w~)} + l/3A2w2N {2/3 rem) + 1/2 J“(o)N)} (42) G(Hz-Wz) = 72 H 1(2th Mr’iw—r {612’(0)H+wN) —J(O)(O)H_wN)l (43) where J‘“)(co)'s are the spectral densities for the different components of the dipolar- dipolar interaction, and J “((0) is the spectral density for CSA interaction, respectively. The CSA is assumed to have an axially symmetrical axis 2 with A = (5,, — l/2 (0,,+oyy). rm, is the distance between the two nuclei. It is often further assumed that all the spectral densities J (w)’s are the same .110)((r)) .__ 10(0)) = J”)(CO) = J ‘c (0)) = J C“ ((0) / P2(COS9 ) (44) 56 where the J “’(m) is the spectral density for the cross-correlation between the dipolar- dipolar and the chemical shift anisotropy interactions if it is not suppressed, P,(cosG) is the second Legendre polynomial, and 0 is the angle between the internuclear vector and the symmetrical axis of CSA. This assumption is valid for isotropic rotational diffusions, and usually a good approximation for most practical systems in liquid. According to equations (41)-(43) it is obvious that only certain frequencies of the bath are directly responsible for NMR relaxation. Those are 0, a)”, (1),, (DH-(0N and (0,,+05N causing zero, single and double quantum transitions of the spin system. Similar equations for the relaxation of antiphase and two spin order coherences have also been derived (see, e. g. Peng & Wagner, 1992), and the equations keeping the cross-correlation term between the dipolar-dipolar and CSA interactions have been published (Goldman, 1984). Relaxation theory in matrix form. The formalisms of the relaxation theories above are expressed in terms of operators. The main result of the semiclassical formalism equation (39) could also be cast in a matrix form. The result is a set of linear equations for the evolution of the matrix elements owls in a representation of the eigenstates or, or', B, B' etc. of the unperturbed spin Hamiltonian H,(s, t) (Redfield, 1965). doa,./ dt = i (CU—or )6m, + 2139' R,,.BB. [CW — 5,39,01,33,] (45) where BBB. is the Kronecker delta. R ,3. is a time-independent element of the relaxation aa'B matrix, and is related to the spectral density Jaw, of the bath. 57 R.......=J......(9'—a' > + J.,...(a —9> «3.4.2 , J.,... (y —a) — 6.. 2 . J... .. (or — y) (46) The spectral density Jam, is the Fourier transform of the time correlation function Cam, (t). JaBa'B' ((1 ‘13) = 1 dt em 43): CaBa'B' (I) and CaBa'B’ (T) = (GaB (t) Grow—1)) (47) The interaction Hamiltonian G(t) responsible for the relaxation of the spin system is a random function of time, and Gm. (t) could be further expanded as before Gaa' (t) = Z1: 11*“) Vaa' (48) where V‘aa. ’s are the spin operators and H(t) ’s are the random functions representing local fluctuating fields. A spectral density involving only bath operators could be defined as 1.,, (a — 13) = 1 dr ,1. ' B" Wt) Ftp-r» (49) The semiclassical theory in matrix form presented above can be easily generalized into a quantum-mechanical theory. One result of such extensions is that the temperature appears in the definition of the quantum spectral density j, p (or — B) = k,,, (or -— B) exp[/r‘ (B — or) / 2.4T], here / is the Boltzman constant. Therefore, we have 58 Jrr(a‘B) =lrk(B"a) exp[/i(B—0t) //Tl (50) This equation represents mathematically the fact that the bath induced transition where the bath gains the energy Ii (B—or) is more probable than the opposite one by a factor exp[/i (B—a) / / T]. Therefore, in quantum-mechanical relaxation theory Raawi Rama, and the disturbed spin system will approach its equilibrium state as it should be. This is in contrast to the semiclassical theory where R0,,” = Rama, and the approach to an equilibrium state is introduced in (40) and (45) as an ad hoc assumption. BPP theory. The semiclassical theory becomes identical with BPP theory for an isolated system consisting of non-interacting nuclear spins (Bloembergen et al., 1948) by an additional assumption that there exists no phase correlation between the eigenstates of the spin Hamiltonian. Consequently, all the off-diagonal elements of the reduced density operator G(t) can be neglected. The underlying mechanism is clear in this theory. Random fluctuations of the bath cause not only single quantum transitions between the nuclear spin energy levels with energy transfer between the bath and the spin system, but also zero and double quantum transitions with energy conservation within the spin system alone. A master equation about the probability of a transition per unit time Wag (= Rm”) could be derived dpa/dtzzB WaB [(pB_pBO)—(pa—pa0)] (51) 59 where p,l and p, are, respectively, the populations of or and B energy states and p00, p,0 their corresponding equilibrium values. The macroscopic magnetization such as M,, which commutes with the spin Hamiltonian of a system of nuclear spins I could be calculated according to M. (t) = 2'4 m19...(t) (52) where pm(t) is the population of an energy level with magnetic number m. Cross-Relaxation. The BPP theory was extended to a two spin system coupled by dipolar-dipolar interaction with four eigenstates ll) = lord), [2) = laB), I3) = |B0t) and I4) = |BB). The following equations for longitudinal components 1,. 3,, and similar ones for transverse components could be derived (Solomon, 1955). dIz / dt = —-p (1,—10) — o (S,— 0) and dS, / dt = -p’ (S,— 0) — o (I,— 0) (53) p = U,,+2U,+U, and p’ = U0+2U,' +U, (54) 0' = U2— U,, (55) and U0 =W323W23, U, = W,,=W2,, U, = WI, and U,’ = W.,: W,, (56) where W.,, i, j = 1, 2, 3, 4, is the transition probability per unit time between the eigenstates li) and 1]) of the longitudinal component of the combined system of spin I and S. p and p’ are auto relaxation rates of spin I and S, respectively. 0 is the cross-relaxation 60 rate between the two spins I and S. One manifestation of the cross-relaxation is nuclear Overhauser effect (NOE) predicted by Overhauser (Overhauser, 1953) and extended to the case of nuclear dipolar-dipolar interaction by Bloch (Bloch, 1954). The distance restraint used in NMR structure determination is extracted from the measurement of the cross-relaxation rate 6 between protons by a NOESY type experiment. For a system consisting only of two protons 'HI and 'H2 their cross-relaxation rate 0’ depends on the internuclear distance r, and the spectral densities J‘q’(ar) at two frequencies. 6 (lHlHIHZ) = Waltz/4r" {6 12120)”) -J‘°’(0)} (57) In NMR structural determination it is assumed that 6 oc l/r“ for any pair of protons in a protein. This essentially is equivalent to the assumption of an isotropic global rotation without internal motion. However, in reality the situation is much more complicated. The spectral densities are related to both the global and internal motion. Moreover, global motion is anisotropic for non-spherical molecules. For any system with three protons sharing a common partner there exists cross-correlation (Hubbard, 1970). Other mechanisms such as random-field may also contribute to the relaxation (Werbelow & Grant, 1977). The compromise of extracting only distance range rather than one accurate number from the NOESY experiment may be the best choice. However, the structures derived in such a manner may be biased, especially for proteins with high internal flexibility. This offers an excellent example that a clear understanding of the underlying 61 physical principles is a prerequisite for the assessments of the results obtained through NMR techniques. Bloch phenomenological equation. The simplest classical relaxation theory (1) was proposed by Felix Bloch in 1946 (Bloch, 1946b) based on phenomenological arguments. It assumes that relaxation is solely determined by two distinct relaxation times T, and T 2. Spectral density mapping. The theories outlined above correspond to the treatment of the Brownian motion by the spectrum method. Their final results are similar: a set of equations such as (41)—(43) relating the measurable NMR relaxation parameters to the spectral densities of the bath at several frequencies. The equations (41)—(43) and two similar ones for two spin relaxation are the starting point of the so-called spectral density mapping method. The original version (Peng & Wagner, 1992) tried to determine the spectral density at the five frequencies 0, 0),, 0),”, to”, or,” (here X represents heteronucleus) through the measurements of both the one- and two-spin relaxation. Modified versions dealing with the spectral density at only three frequencies have been developed to compensate for the experimental errors at high frequencies by assuming that J,” (00,”) z «Am (0)”) 2: un (0),”) (Farrow et al., 1995; Peng & Wagner, 1995; Lefevre et al., 1996), here ny (0)) is the spectral density of the X-H vector. It is true that spectral densities at certain frequencies are the only parameters that can be directly determined by NMR relaxation experiments at least nowadays. However, the information context about internal motion obtained by the mapping method is very limited. Furthermore, without resort to a certain model the result can not be correlated or compared with the microscopic properties of the bath obtained by means of theoretical or other experimental 62 methods. In fact, employing a model to describe relaxation is related to the diffusion equation method (Wang & Uhlenbeck, 1945) for analyzing the Brownian motion. The two methods are not completely equivalent under all circumstances but they give identical results for the processes much longer than the correlation times characteristic of the Brownian motion of the bath (Chandrasekhar, 1943). In contrast to the mapping method the macroscopic diffusion constant obtained from the diffusion equation method can be related to other microscopic quantities of the bath. Diffusion equation method. This method assumes that NMR relaxation in liquid is caused by the discrete random walk processes through short distances or over small angular orientations. Passing over to the continuous case the probability distribution satisfies a particular partial differential equation of diffusional type with certain boundary conditions (Chandrasekhar, 1943). For a random Markofl process there exists a general solution POT t), and a Green function P, (f,, t; f, t+r), here P(fl t) df is a priori probability (at equilibrium) of finding f in the range of (t: f + df) at a time t, and P,(f,, t; f}, t+t) df, is the conditional probability of given f, at time t one finds f, in the range of (f,, f, + df2) a time I later. These two probabilities are simply related P(f,, t+r) = l P (f, t)P_.(f,, t; f, m) df, (58) Furthermore, lim,_,,, P,(f,, t; f, t+t) is equal to P(f,' t) for the thermodynamic fluctuations at equilibrium but for an irreversibility process it becomes P03, 00) of the newly established equilibrium state. For NMR relaxation we have P03, 00) = P( f, t) according to 63 the assumptions about the bath (31). The equivalence of the spectrum and the diffusion equation methods is established by the following basic relationship. Cult. t+I) = rv = 11 PM, 0PM. tif.» t+1)17.11?)Fz.'(f2)dfr dfz (59) Motional models. As is well known both the overall and internal rotations contribute to NMR relaxation in liquid. The overall rotation can be either isotropic or anisotropic depending on the shape of the molecule under investigation. The internal rotation can be either free or restricted, about one bond or multiple bonds, and the bond can be either fixed or wobbling. A model is distinguished by the particular diffusion equation, and by its boundary condition suggested by physical instinct and the nature of the system. A general model appropriate for an anisotropic overall rotation, and a restricted internal rotation about multiple wobbling bonds will be quite complicated. Some simplifications are usually necessary for extracting the dynamical information from NMR relaxation data. Sphere without internal motions. The simplest model considers a protein in solution as an isotropically rotating sphere without internal motions (Abragarrr, 1961). Its correlation function C(t) is C(T) = expf-ITl/T.) (60) where the parameter r, is the correlation time characteristic of the bath. The irreversibility is encoded in the absolute value of It]. The conditional probability P(Q,,, Q, r), where 0,, 64 and Q are the Euler angles specifying the orientation of the molecule in a laboratory frame, is the solution of a diffusion equation for the free Brownian rotational motion with the initial condition (MS), 0) = 5(0 — £20) dry/6t = D, A,\V (61) where D, is the rotational diffusional coefficient and A, the spherical Laplacian operator. This diffusion equation is the special case of the more general Smoluchowski equation with the potential V But/6t = div[D gradw + vf" grad V] (62) where D and f are the diffusion constant and the frictional coefficient of the rotational motion, respectively (Chandrasekhar, 1943). The solution P620, (1, r ) can be found by first expanding w in spherical harmonics Y,’, and then substituting the result into (61). P610. 9. I ) = 21.... YmP(Qo) Y...’ (Q) “pl-110+1 )D.] (63) With a priori probability P(Q,,) = l/(8Tt2) the corresponding correlation function can be obtained from P(Q,,, Q, t ) by (59). Arbitrary shape without internal motion. The distribution function P(Q,,, Q, t) for the overall motion of an arbitrary shaped molecule is a solution of the following 65 diffusion equation for the free Brownian rotational motion with the initial condition w“), 0) = 8(Q — 00) (F avro, 1960), fill/at = Zkak L k2‘l’ (64) where k = x, y, z, and Lf, Lyz, L," are the components of angular momentum L. The molecule-fixed frame with the axes x, y and z diagonalizes the diffusion tensor D, and D,,, D, and D,z are the three principal diffusional coefficients. The solution could be obtained similarly as before except that w is now expanded in terms of functions D’m,,(Q), which are the elements of Wigner rotation matrix D’(Q) (Hubbard, 1970). They are the eigenfunctions of the total angular momentum of a rigid body (Brink & Satchler, 1993). These eigenfunctions form a complete set in the space of Euler angles. P610. 0. l‘) = 2°10 (21 +1) /(873) tr{lexm-Q '0] D"(Qo) ”(0)1 (65) The (21+!) x (21+1) real, symmetric matrix Q1 has matrix elements QC... = <1ml D+ L2 + (D...— -D+)L.2 + D- (U - 142) |ln> (66) where |lm> is the eigenket of L2 and L,2 with eigenvalues 1 (1+1) and m, respectively, and D, a 1/2 (Dni- D,,,). If one introduces the quantities 66 b... slD.+m2(D..-D.) = b.-. (67) then in the special case with axially symmetry (D_ = 0) the matrix Q1 becomes diagonal, [Cxp(—Q l(t)]rnn : 5 mn exp(—blmt) (68) and P(Q(lr Q! t) = Elm," (21 +1) /(81'[2) [exp(—blmt)] Dlnm(00) Dlnm(Q) (69) where summation 1 runs from 0 to +00 and m, n from —1 to +1. For relaxation dominated by the dipolar-dipolar and CSA interactions 1 is equal to 2. Isotropic rotation of a sphere corresponds to D”: D”: D,, and (69) becomes (63). With a priori probability P(Q,,) = 1/(8112) the corresponding correlation functions can be obtained from (59). Model with internal motion. With internal motion the solutions to the diffusion equation become much more complicated. As the first step of simplification it is usually assumed that the internal motion is independent of the overall rotation. This is valid for the isotropic overall rotations of spherical molecules, and normally a good approximation for arbitrary shaped molecules. The correlation function in the laboratory frame can be related to that in a molecule-fixed frame diagonalizing the diffusion tensor by Wigner rotation matrices D2 '5, since the elements of the correlation functions of most practical "Ill NMR systems are second-rank spherical tensors (Brink & Satchler, 1993). CM, (0 = 2 (17.... [0(0)] D7 [0(01) (02.0. [9(0), ¢(0)] Dim [G(t), ¢(t)]) (70) 67 The summation n, n’ runs from —2 to +2. Where Q is referred to the laboratory frame and 0, d) to the molecule-fixed frame, respectively. The overall rotation is characterized by the first average while the internal motion by the second average over a bath ensemble. Model-free approach. One of the simplest ways to include internal motion is to assume that the correlation function C,(t) for the internal motion can be divided into two parts (Lipari & Szabo, 1982a, b): the average value S2 of the components of C,(t) representing the restrictions of the internal motion, and a weighted normal correlation function Q(t) with the properties that lim,_,,, Q(t) = 0 and Q(O) = 1 characterizing the randomness of the motion. In other words, the average value of Q(t) is zero. C,(t) = S2 + (1— S2)Q(t) (71) S2 is called order parameter and equal to lim,_,,o C,(t). It does not depend on time and is not a random function. With the assumption that the overall rotation is isotropic, and is independent of the internal motion the total correlation firnction C(t) becomes C(t) = C0(1) C[(t) = [”5 CXPH /IM)] C10) (72) where C,(t) is the correlation function for the overall motion. The corresponding spectral density J(o)) is J((r)) = 2S2tM /5 [1+(th)2] + 2/5 (1— S2) 1,,” dt cosmt Q(t) exp(—t /rM) (73) 68 If we assume that Q(t) = exp(— t /r,) we have J(o3) = 2/5{S"t,,, / [1+(corM)2] +(1— SZ) 1: / [1 +(wr)2]} (74) where t " = 1M" + t,", and t, is defined according to r, (1— S") = 10“” (C,(t) — S2) dt. If the internal motion is much faster than the overall motion, and lies in the extreme narrowing range, that is, the correlation time for internal motion r, satisfies the condition 1:, << 1,, and t, << (1), then (74) becomes J((r)) = 2/5 {Sth / [1 +(corM)2]+( 1 —S")r,} (75) For internal motion much slower than overall motion we instead have J((r)) = ZIM /5 [1+(wrM)2]. Order parameter. The physical meaning of the order parameter S’ is quite clear according to the definition of the correlation function (59). With the modified spherical harmonics C2,, (Brink & Satchler, 1993) substituting for the general random function in (59) we have 52 =1im..... an, t; 0... 1+1) =1im E Av T—‘NXJ ”I = 2m Av Av: zm iAvi2 (76) 69 where the summation runs from —2 to 2, and C2,,1(Q,, 00) is the spherical harmonics at the final equilibrium state reached by relaxation. We have Av = ,,= Av based on the assumptions about the bath introduced before (31). In terms of the probability distribution function the order parameter S2 is 5" =1im..., 2,, ll P(o,, t)P,(o,, 1; 9,, 1+1) C,,,(o,, t) C,,,‘(o,, 1+1) do, do, = 2,1113“), t)P,(Q,, :; o, oo) C,,,,(Q,, z) C,,,‘(rr,, 00) do, do, = 2,, {lP(o,, t) C,,,(o,. t) (10,} {l C2,,‘(o,, 0102(1), t) 210,} = 2.. l{lP(Q. t) €2.19. 9 amt = 2,. I112 (77) where the relations P,(Q,, t; (2,, oo) = P,(Q,, co) = P,(Q,, t) = P,(Q,, t) = P(Q, t) of the equilibrium distribution have been used. Therefore, S2 is nothing more than the average value over a bath ensemble of the spherical harmonics of the Euler angles specifying the orientation of a vector relative to a molecule-fixed frame. It is also the first nontrivial term in the expansion of P(Q, t) in a series of Legendre polynomials (Lipari & Szabo, 1980). However, how well it represents P(Q, t) depends on how fast this series converges. Therefore, a model is needed to establish the relationship between S2 and the distribution of the orientation. The information context of S2 corresponds to the description of a random process by only the first set of the probability distribution P(Q, t) (Wang & Uhlenbeck, 1945). The conformation with high energy or rare probability is not being sampled properly by S2. The claim that S" is insensitive to motions slower than nanosecond is misleading. The insensitivity is not due to the time scale but due to the energies of these conformations. Though P(Q, t) can not be easily obtained for a complex 70 system such as a protein even a random Markofl process must be specified by a joint probability or conditional probability P,(Q,, t; (2,, 1+1) together with the equilibrium distribution P(Q, t). Extension of model-free approach. For some proteins it was found that two exponential terms instead of one were needed to fit their relaxation data by the Lipari & Szabo approach (Clore et al., 1990). Moreover, an exchanging term R, has been added to the transverse relaxation rate constant 1/ T 2 to account for the contribution to it from the possible slow motions (Kay et al., 1989; Clore et al., 1990) other than the dipolar-dipolar and CSA interaction. The term R“ is usually explained as the contribution fi'om conformational exchanges. However, the physical meanings of these added terms are not well defined. They are related to the diffusion tensor in some complicated ways. Furthermore, for proteins with arbitrary shapes the total correlation function can not be separated rigorously into an overall and an internal part. Therefore, one should always bear in mind the range of validity of the original formalism or its extension when using it to interpret NMR relaxation data. Recently the Lipari & Szabo approach has been extended to the cross-correlation by several authors (Kay & Torchia, 1991; Zhu et al., 1995; Daragan & Mayo, 1995; Daragan & Mayo; 1997). The cross-correlation is responsible for the difference of relaxation behaviors in the two outer lines from the two inner ones, and manifests itself through the nonexponential behavior of the relaxation. It has been treated in detail in two reviews (Werbelow & Grant, 1977; Vold & Vold, 1978). Restricted rotational diffusion model. Several more sophisticated models are in vogue. The first class may be referred to as restricted rotational diffusion model (Wittebort & Szabo, 1978, London & Avitabile, 1978). The conditional probability P(y', t 71 ; y, 0) of the internal rotation is the solution of one dimensional diffusion equation dw/o‘t = D EN! /827 with the initial condition w“), 0) = 5(0 — £20) and the boundary condition (hwy/61y | in. = 0, where y is azimuthal angle. The boundary condition is the mathematical statement that the internal rotational diffusion is restricted between iy,,. P(r'. t: v. 0)=1/(210){1 + 2..=zcos[nn(r-ra)/2101 coslnn(r’-ro)/2nl exr)(-t/I..)} (78) where t, = 4702 /(Dn"rt"). With a priori probability P620) = 1/(2y0) the corresponding correlation function can be obtained from (59). With this model we can include excluded volume effects accounting for the restricted space an internuclear vector could have to diffuse. The Woessner model (Woessner, 1962) of free internal rotation is included as a special case of this general model with 17,, = i180°. The model above is good for the internal rotation about one bond. It can be easily generalized into a model for internal rotations about N bonds (Wallach, 1967). The transformation from a frame F attached to a nucleus whose relaxation is of interest to the laboratory frame L can be achieved by a series of Wigner’s rotation matrix D2's with Euler angles (20,, (2,2. QM, N specifying the orientations of the successive coordinate systems. D2q()(QLF) = Za,b,---b"D2qa(QLD)D2 (QB/)D2b,b2(012)‘”Dzbn,,bn(QN,N—I) D2b"0(QNF) (79) ab, The internal rotations are usually assumed to be independent of each other in order to simplify the problem. 72 Discrete-jump model. The discrete-jump model with M configurations (Wittebort & Szabo, 1978) tries to give a more realistic description of concerted internal rotations with mutual dependence. It is also an extension of the Woessner jump model with only three configurations (Woessner, 1962). The underlying mathematics is the same: the conditional probability P(k, t; l, 0) of the side chain of a macromolecule has configuration k at time t if it has configuration I at time 0 is the solution of the mater equation (81) with the initial conditions (80): P(k, t; 1, 0) = 5,, 1 = 0 and lim,_,,,, P(k, t; 1, 0) = P,,(k) for all 1. (80) dry/6t: ZMJ.=,R,j\V (81) where R, is the rate constant for the transition from configuration j to i with M total available configurations. P,q(k) is a priori probability at equilibrium satisfying the equation )3 ”H R]. P,q(i) = 0. The solution can be obtained by matrix method, and is expressed in terms of the eigenvectors X ,'= (X ,0, mX 1M) and eigenvalues 1.,/s of the matrix Q with elements Q”: (R, R1,)”. 1’09 t: 1, 0) = X10 (1110)"l 2,. X1" X1" eXP(-1~,.t) (82) where X ,0 = [P,q(k)]"'2. Wobbling in a cone model. This model assumes that an internuclear vector is wobbling uniformly in a cone (Kinoshita et al., 1977; Lipari & Szabo, 1980). The 73 conditional probability P(QO; Q, t) is the solution of a diffusion equation for the free Brownian rotational motion with the initial condition w(Q, 0) = 5(0 — Q”), and the boundary condition dul(0)/60 l80 = 0, here 0 is the polar angle. (NI/61 =1), AN (83) where D, is the diffusion coefficient of the wobbling motion. The boundary condition is the mathematical statement that the internal rotational diffusion is restricted within a polar angle 0,, but does not depend on the azimuthal angle. The solution could be obtained as an expansion of associated Legendre functions. However, the resulting expressions for P(Qo; Q, t), and consequently for the correlation function C,(t) are not closed analytic firnctions. CAI) = 21' Ai exp(—Dwt/Gi) (84) where A , and o, are functions of 00. Instead an approximate expression for C,(t) is derived which is exact at time t = 0 and t = 00 (order parameter S2), and has the property that the area under the approximate correlation function is the same as that under the exact correlation function. C,(t) = .S‘2 + (l— SZ) exp (-D,t/ ) (85) 74 where S2 = A,o and = Z,,.,, A,o, / (l- 82). The a priori probability P(00) = [21t(1—c0300)]". The wobbling motion has been combined with a free rotation to give a more realistic description of the internal motion (Richarz et al., 1980). Intermediate and slow motions. The NMR relaxation measurement discussed above can provide dynamical information about the fast motion on the time scale ranging from 10"2 to 10'8 second. The motions in the intermediate time scale from microsecond to millisecond are much more difficult to be studied by NMR. The measurement of the exchange term R“, sensitive to motions on the microsecond to millisecond time scale, will be discussed later. The slow motion in the time scale ranging approximately from milliseconds to seconds gives cross-peaks between separate resonances which could be observed in homonuclear NOESY, ROESY and in heteronuclear longitudinal magnetization or two-spin order exchange experiments. Most of NMR techniques for measuring intermediate or slow motions have been discussed in a recent review (Palmer et al., 1996). Pulse sequences for measuring NMR relaxation parameters. The 2D pulse sequences for NMR relaxation experiments have been reviewed in a recent paper (Peng & Wagner, 1994), and the versions with pulsed field gradients are also available (Farrow et al., 1994). Frequently the inversion-recovery (Carr & Purcell, 1954; Vold et al., 1968) and the Carr-Purcell-Meiboom-Gill (CPMG) (Carr & Purcell, 1954; Meiboom & Gill, 1958) techniques are used to measure T, and T2, respectively. 114.0) = M0 [(1-2 eX13(-t/ T 1) l (86) M,, (211:) = M, (0) exp(—2nt/T,) exp(—2y2DG2nt3/3) (87) 75 where G is the gradient, D diffusion coefficient and n the number of echo. The heteronuclear NOE enhancement is measured by steady-state NOE technique. The experimental techniques for measuring two-spin order or antiphase coherence have also been developed (Peng & Wagner, 1992). These pulse sequences normally include pulses to suppress cross-correlation for minimizing its contribution to the relaxation (Boyd et al., 1990; Kay et al., 1992; Palmer et 1a., 1992). The interference from the chemical exchange or slow motions on the microsecond time scale can be measured by T ,p measurement (Szypersky et al., 1993) or spectral density mapping method. The exchange term Ra manifests itself by an anomalous decrease in T 2 relative to that predicted for dipolar, CSA or quadrupolar relaxation in a free-precession NMR spectrum (the line width), or in a CPMG (Allerhand & Thiele, 1968), or T ,0 experiment. In the T ,p experiment R, can be identified by directly measuring 1,, according to the equation (Davis et al., 1990, where a number of general cases are also discussed) R... = (5(0)2 PIPE T.,. / (1+1...2(D.2) (88) by varying the effective spinlock field a), (Akke & Palmer, 1996). Where 801) is the difference in chemical shift, and p A, p, are the time fractions spent in the conformation A and B, respectively. T he future of NMR relaxation study. A more realistic model needs to be derived from a general diffusion equation with external forces (Chandrasekhar, 1943) with the 76 considerations of more relevant factors such as solvent effect and cross-correlation effect. A unified relaxation theory is also very helpful for distinguishing and quantitating the contributions to the total relaxation of the various internal motions on different time scales. However, even with a quite realistic model the concerted movements of large fragments, which may be the most important internal motions for protein functions, could not be obtained because the dynamical information obtained from NMR experiments is local. Certain calculation procedures need to be developed to identify the collective motion. Most importantly, experimental techniques need to be improved or developed to measure accurately the relaxation parameters, or to measure new relaxation parameters. Finally, dynamical information should be combined with biochemical data to establish the roles of internal motions in the biochemical functions of proteins. 77 References Abragam, A. (1961) Principles of Nuclear Magnetism, Clarendon Press, Oxford. Chapter VIII, XXI. Akke, M., & Palmer A. G. (1996) J. Am. Chem. Soc. 118, 911-912. Allerhand, A., & Thiele, E. (1968) J. Chem. Phys. 45, 902-916. Anil-Kumar, Ernst, R. R., & Wfithrich, K. (1980a) Biochem. Biophys. Res. Commun. 95, 1-6. Anil-Kumar, Wagner, G., Ernst, R. R., & Wiithrich, K. (1980b) Biochem. Biophys. Res. Commun. 96, 1156-1163. Argyres, P. N., & Kelley, P. L. (1964) Phys. Rev. 134A, 98-111. Aue, W. P., Bartholdi, E., & Ernst, R. R. (1976) J. Chem. Phys. 64, 2229-2246. Banwell, C. N., & Primas, H. (1962) Molec. Phys. 6, 225-256. Basus, V. J. (1989) Methods Emzymol. 177, 132-149. Bax, A., Clore, G. M., & Gronenbom, A. M. (1991).]. Magn. Reson. 88, 425-431. Bax,A., & Davis, D. G. (1985) J. Magn. Reson. 65, 355-360. Bax, A., DeJong, P. G., Mehlkopf, A. F. , & Smidt, J. (1980) Chem. Phys. Lett. 69, 567- 570. Bax, A., Driscoll, P. C., Clore, G. M., Gronenbom, A. M., Ikura, M., Kay, L. E. (1990) J. Magn. Reson. 87, 620-627. Bax, A., Griffey, R. H., & Hawkins, B. L. (1983) J. Magn. Reson. 55, 301-315. 78 Bax, A., & Pochapsky, S. S. (1992) J. Magn. Reson. 99, 638-643. Bloch, F. (1946b) Phys. Rev. 70, 460-474. Bloch, F. (1954) Phys. Rev. 93, 944-944. Bloch, F. (1956) Phys. Rev. 102, 104-135. Bloch, F. (1957) Phys. Rev. 105, 1206-1222. Bloch, F. (1958) Phys. Rev. 111, 841-853. Bloch, F ., Hansen, W. W., & Packard, M. (1946a) Phys. Rev. 69, 127-127. Bloembergen, N., Purcell, E. M., & Pound, R. V. (1948) Phys. Rev. 73, 679-712. Bloom, A. L., & Schoolery, J., N. (1955) Phys. Rev. 97, 1261-1265. Bodenhausen, G., & Ruben, D. J. (1980) Chem. Phys. Lett. 69, 185-188. Bodenhausen, G., Vold, R. L., & Vold, R. R. (1980) J. Magn. Reson. 37, 93-106. Boyd, J ., Hommel, U., Campbell, I. D. (1990) Chem. Phys. Lett. 175, 477-482. Braunschweiler, L., & Ernst, R. R. (1983) J. Magn. Reson. 53, 521-528. Brink, D. M., & Satchler, G. R. (1993) Angular Momentum, Oxford University Press, New York. Burum, D. R., & Ernst R. R. (1980) J. Magn. Reson. 39, 163-168. Carr, H.Y., & Purcell, E. M. (1954) Phys. Rev. 94, 630-638. Cavanagh, J., Palmer, A. G., Wright, P. E., & Rance, M. (1991) J. Magn. Reson. 91, 429-436. Chandrasekhar, S. (1943) Rev. Mod. Phys. 15, 1-89. Clore, G. M., Bax, A., Driscoll, P. C., Wingfield, P. T., & Gronenbom, A. M. (1990) Biochemistry 29, 8172-8184. Clore, G. M., Driscoll, P. C., Wingfield, P. T., Gronenbom, A. M. (1990) Biochemistry 79 29, 7387-7401. Clore, G. M., & Gronenbom, A. M. (1991a) Science 252, 1390-1399. Clore, G. M., & Gronenbom, A. M. (1991b) Annu. Rev. Biophys. Chem. 21, 29-63. Clore, G. M., Szabo, A., Bax, a., Kay, L. E., Driscoll, P. C., & Gronenbom, A. M. (1990).]. Am. Chem. Soc. 112, 4989-4991. Daragan, V. A., & Mayo, K. H. (1995) J. Magn. Reson. B107, 274-278. Daragan, V. A., & Mayo, K. H. (1997) Prog. NMR Spectrosc. 31, 63-105. Davis, D. G., Perlman, M. E., London, R. E. (1990).]. Magn. Reson. B104, 266-275. Davis, A. L., Boelens, R., & Kaptein, R. (1992).]. Biomol NMR 2, 395-400. Dayie, K. T., Wager, G., & Lefevre, J. F. (1996) Annu. Rev. Phys. Chem. 47, 243-282. Delaglio, F., Grzesiek, S., Vuister, G. W., Zhu, G., Pfeifer, J ., & Bax, A. (1995).]. Biomol.NMR 6, 277-293. Doddrell, D. M., Pegg, D. T., & Bendall, M. R. (1982).]. Magn. Reson. 48, 323-327. Ernst, R. R., & Anderson, W. A. (1966) Rev. Sci. Instrum. 37, 93-102. Ernst, R. R., Bodenhausen, G., & Wokaun, A. (1987) Principles of Nuclear Magnetic Resonance in One and Two Dimensions, Clarendon Press, Oxford. Fano, U. (1953) Phys. Rev. 96, 869-873. Fano, U. (1957) Rev. Mod. Phys. 29, 74-93. Farrow, N. A., Muhandiram, R., Singer, A. U., Pascal, S. M., Kay, C. M., Gish, G., Shoelson, S. E., Pawson, T., Kay, J. D. F., & Kay, L. E. (1994) Biochemistry 33, 5984-6003. Farrow, N. A., Zhang, 0., Forrnan-Kay, J. D., & Kay, L. E. (1997) Biochemistry 36, 2390-2402. 80 Farrow, N. A., Zhang, 0., Szabo, A., Torchia, D. & Kay, L. E. (1995) J. Biomol. NMR 6, 153-162. Farvo, L. D. (1960) Phys. Rev. 119, 53-62. Fesik, S. W., & Zuiderweg, F. R. P. (1988) J. Magn. Reson. 78, 588-593. Garret, D. S., Seok, Y., Liao d., Peterkofsky, A., Gronenbom, A. M., & Clore, G. M. (1997) Biochemistry 36, 2517-2530. Goldman, M. (1984) J. Magn. Reson. 60, 437-452. Griesinger, C., Otting, G., Wfithrich, K., & Ernst, R. P. (1988) J. Am. Chem. Soc. 110, 7870-7872. Griesinger, C., Sorensen, O. W., & Ernst, R. P. (1985) J. Am. Chem. Soc. 107, 6394- 6396. Gronenbom, A. M., & Clore, G. M. (1995) CRC Crit. Rev. Biochem. Mol. Biol. 30, 351-385. Grzesiek, S., Anglister, J., & Bax, A. (1993) J. Magn. Reson. BI 01, 114-119. Grzesiek, S., & Bax, A. (1992a) J. Magn. Reson. 96, 432-440. Grzesiek, S., & Bax, A. (1992b) J. Am. Chem. Soc. 114, 6291-6293. Grzesiek, S., & Bax, A. (1993a) J. Am. Chem. Soc. 115, 12593-12594. Grzesiek, S., & Bax, A. (1993b) J. Biomol. NMR 3, 185-204. Hahn, E. L. (1950) Phys. Rev. 80, 580-594. Havel, T. F., & Wiithrich, K. (1984) Bull. Math. Biol. 46, 673-698. Hubbard, P. S. (1961) Rev. Mod. Phys. 33, 249-164. Hubbard, P. S. (1970) J. Chem. Phys. 52, 563-568. Huntress, W. T. (1970) Adv. Magn. Reson. 4, 1-37. 81 Ikura, M., Clore, G. M., Gronenbom, A. M., Zhu, G., Klee, C. B., & Bax, A. (1992) Science 256, 632-638. Ikura, M., Kay, L. E., & Bax, A. (1990a) Biochemistry 29, 4659-4667. Ikura, M., Kay, L. E., Tschudin, R., & Bax, A. (1990b) J. Magn. Reson. 86, 204-209. Jeener, J. (1971) Ampere International Summer School, Basko Polje, Yugoslovia Jeener, J., Meier, B. H., & Ernst, R. R. (1978).]. Chem. Phys. 71, 4546-4553. John, B. K., Plant, D., Heald, S. H., & Hurd, R. D. (1991) .1. Magn. Reson. 94, 664-669. Karplus, M. (1959) J. Phys. Chem. 30, 11-15. Karplus, M., & McCammon, J. A. (1983) Ann. Rev. Biochem. 53, 253-300. Karplus, M., & Petsko, G. A. (1990) Nature 347, 631-639. Kay, L. E. (1995) Prog. Biophys. Molec. Biol. 63, 277-299. Kay, L. E., & Bax, A. (1990) J. Magn. Reson. 86, 110-126. Kay, L. E., Clore, G. M., Bax, A., & Gronenbom, A. M. (1990c) Science 249, 411-414. Kay, L. E., Ikura, M., & Bax, A. (1990b) .1. Am. Chem. Soc. 112, 888-889. Kay, L. E., Ikura, M., Tschudin, R., & Bax, A. (1990a) J. Magn. Reson. 89, 495-514. Kay, L. E., Keiffer, P., & Saarinen, T. (1992a) J. Am. Chem. Soc. 114, 10663-10665. Kay, L. E., Nicholson, L. D., Delaglio, F., Bax, A., & Torchia, D. (1992b) J. Magn. Reson. 97, 359-375. Kay, L. E., & Torchia, D. (1991).]. Magn. Reson. 95, 536-547. Kay, L. E., Torchia, D., & Bax, A. (1989) Biochemistry 28, 8972-8979. Kay, L. E., Xu, G. Y., Singer, A. U., Muhandiram, D. R., & Forman-Kay, J. D. (1993) J. Magn. Reson. BIO], 333-337. Kinoshita, K. Jr., Kawato, S., & Ikegami, A. (1977) Biophys. J. 20, 289-305. 82 Kuszewski, L., Qin, J., Clore, G. M., & Gronenbom, A. M. (1995).]. Magn. Reson. 3106, 92-96. Landau, L. D., & Lifshitz, E. M. (1969) Statistical Physics, pp. 6, Addison-Wesley, Reading, Massachusetts. Lefevre, J. F., Dayie, K. T., Peng, J. W., & Wagner, G. (1996) Biochemistry 35, 2674- 2686. Levitt, M. H., Freeman, R., & Frenkiel, T. (1983) Adv. Magn. Reson. 11, 47-1 10. Lipari, G., & Szabo, A. (1980) Biophys. J. 30, 489-506. Lipari, G., & Szabo, A. (1982a) J. Am. Chem. Soc. 104, 4546-4559. Lipari, G., & Szabo, A. (1982b) J. Am. Chem. Soc. 104, 4559-4570. Logan, T. M., Olejniczak, E. T., Xu, R. X., & Fesik, S. W. (1993) J. Biomol. NMR 3, 225-231. London, R. E., & Avitabile, J. (1978) J. Am. Chem. Soc. 99, 7765-7776. Lowe, I. J., & Norberg, R. E. (1957) Phys. Rev. 107, 46-61. Macura, S., & Ernst, R. R. (1980) Mol. Phys. 41, 95-117. Marion, D., Driscoll, P. C., Kay, L. E., Wingfeld, P. T., Bax, A., Gronenbom, A. M., & Clore, G. M. (1989d) Biochemistry 28, 6150-6156. Marion, D., Ikura, M., & Bax, A. (1989c) J. Magn. Reson. 84, 425-430. Marion, D., Ikura, M., Tschudin, R., & Bax, A. (1989a) J. Magn. Reson. 85, 393-399. Marion, D., Kay, L. E., Sparks, S. W., Torchia, D. A., Bax, A. (1989b) J. Am. Chem. Soc. 111,1515-1517. Marion, D., & Wuthrich, K. (1983) Biochem. Biophys. Res. Commun. 113, 967-974. Maudsley, A. A., Wokaun, A., & Ernst, R. R. (197 8) Chem. Phys. Lett. 55, 9-14. 83 McCoy, M. A., & Miieller, L. (1992a) J. Am. Chem. Soc. 114, 2108-2112. McCoy, M. A., & Mueller, L. (1992b) J. Magn. Reson. B98, 674-679. Meiboom, S., & Gill, G. (1958) Rev. Sci. Instrum. 29, 668-691. Messerle, B. A., Wider, G., Otting, G., Weber, G, & Wiithrich, K. (1989) J. Magn. Reson. 85, 608-613. Morris, G. A., & Freeman, R. (1979) J. Am. Chem. Soc. 101, 760-762. Miieller, L. (1979) J. Am. Chem. Soc. 101, 4481-4484. Miieller, L. (1987) J. Magn. Reson. 72, 191-196. Muhandiram, D. R., & Kay, L. E. (1994) J. Magn. Reson. B 103, 203-216. Muhandiram, D. R., Xu, G., & Kay, L. E. (1993).]. Biomol. NMR 3, 463-470. Nilges, M., Clore, G. M., & Gronenbom, A. M. (1988) FEBS Lett. 229, 129-136. Overhauser, A. W. (1953) Phys. Rev. 89, 689-700; ibid, 92, 411-415. Palmer, A. G., Cavanagh, J., Wright, P. E., & Rance, M. (1991) J. Magn. Reson. 93, 151-170. Palmer, A. G., Skelton ,N. J ., Chazin, W. J ., Wright, P. E., & Rance, M. (1992) Mol. Phys. 75, 699-711. Palmer, A. G., Williams, J ., & McDermott, A. (1996) J. Phys. Chem. 100, 13293-13310. Peng, J. W., & Wagner, G. (1992) J. Magn. Reson. 98, 308-332. Peng, J. W., & Wagner, G. (1994) Methods in Enzymology 239, 563-596. Peng, J. W., & Wagner, G. (1995) Biochemistry 34, 16733-16752. Piantini, U., Sorensen, O. W., & Ernst, R. P. (1982).]. Am. Chem. Soc. 104, 6800-6801. Piotto, M., Saudek, V., & Sklenar, V. (1992) J. Biomol. NMR 2, 661-665. Prigogine, I. (1962) Non-Equilibrium Statistical Mechanics, John Wiley & Sons, New 84 York, New York. Purcell, E. M., Torrey, H. C., & Pound, R. V. (1946) Phys. Rev. 69, 37-38. Rabi, I. I., Schwinger, J., & Ramsey, N. F. (1954) Rev. Mod. Phys. 26, 167-171. Redfield, A. G. (1958) IBM J. Res. Dev. 1, 19-31. Redfield, A. G. (1965) Adv. Magn. Reson. 1, 1-32. Redfield, A. G., & Gupta, R. K. (1971) Adv. Magn. Reson. 5, 81-115. Resing, H. A., & Torrey, H. C. (1963) Phys. Rev. 131, 1102-1104. Rice, S. O. (1944) in Selected Papers on Noise and Stochastic Processes edited by Nelson Max pp. 133-294, Dover, New York, New York. Richarz, R., Nagayama, K., & Wuthrich, K. (1980) Biochemistry 19, 5189-5196. Royden, V. (1954) Phys. Rev. 96, 543-544. Shaka, A. J., Barker, P. B., & Freeman, R. (1985) J. Magn. Reson. 64, 547-552. Shaka, A. J ., Keeler, J., Frenkiel, T., & Freeman, R. (1983) J. Magn. Reson. 52, 335-338. Shaka, A. J., Lee, C. J., & Pine, A. (1988).]. Magn. Reson. 77, 274-293. Shukert, S. B., Hajduk, P. J., Meadows, R. P., & Fesik, S. W. (1996) Science 274, 1531- 1534. Sklenar, V., & Bax, A. (1987) J. Magn. Reson. 74, 469-479. Snyder, G. H., Rowan, R., Karplus, S., & Sykes, B. D. (1975) Biochemistry 14, 3765- 3777. States, D. J ., Haberkon, R. A., & Ruben, D. J. (1982) J. Magn. Reson. 48, 286-292. Stejskal, E. O., & Tanner, J. E. (1965) J. Chem. Phys. 42, 288-292. Solomon, 1. (1955) Phys. Rev. 99, 559-565. Sorensen, O. W., Eich, G. W., Levitt, M. H., Bodenhausen, G., & Ernst, R., R. (1983) 85 Prog. NMR Spectrosc. 16, 163-192. Szypersky, T., Luginbuhl, P., Otting, G., Guntert, P., & Wiithrich, K. (1993) J. Biomol. NMR 3,151-164. Torrey, H. C. (1953) Phys. Rev. 92, 962-969. Torrey, H. C. (1954) Phys. Rev. 96, 690-690. Van de Ven, F. J. M., & Hibers, C. W. (1983).]. Magn. Reson. 54, 512-520. Vold, R. L., & Vold, R. R. (1978) Prog. NMR Spectrosc. 12, 79-133. Vold, R. L., Waugh, J. S., Klein, M. P., Phelps, D. E. (1968) J. Chem. Phys. 48, 3831- 3832. Von-Neumann, J. (1927) J. Gottinger Nachr. 245 and 273. Von-Neurnann, J. (1955) Mathematical Foundations of Quantum Mechanics, Princeton University Press, Princeton. New Jersey. Viiister, G. W., & Bax, A. (1993) J. Am. Chem. Soc. 115, 7772-7777. Wagner, 6., & Wiithrich, K. (1982) J. Mol. Biol. 155, 347-366. Wallach, D. (1967) J. Chem. Phys. 4 7, 5258-5268. Wang, Y-X., Freedberg, I. D., Yamazaki, T., Wingfield, P. T., Stahl, S. J ., Kaufman, J. D., Kiso, Y., & Torchia, D. A. (1996) Biochemistry 35, 9945-9950. Wang, M., & Uhlenbeck, G. E. (1945) Rev. Mod. Phys. 17, 323-342. Wangsness, R. K., & Bloch, F. (1953) Phys. Rev. 89, 728-739. Werbelow, L. G., & Grant, D. M. (1977) Adv. Magn. Reson. 9, 189-299. Wider, G., Hosur, R. V., & Wiithrich, K. (1983) J. Magn. Reson. 52, 130-135. Wider, G., Riek, R., & Wiithrich, K. (1996) J. Am. Chem. Soc. 118, 1 1629-11634. Wider, G., & Wiithrich, K. (1993) J. Magn. Reson. A102, 239-241. 86 Williamson, M. P., Havel, T. F., & Wiithrich, K. (1985) J. Mol. Biol. 182, 295-315. Wittebort, R. J., & Szabo, A. (1978) J. Chem. Phys. 69, 1722-1737. Wittekind, M., & Mfieller, L. (1993) J. Magn. Reson. 101, 201-205. Woessner, D. E. (1962).]. Chem. Phys. 36, 1-4. Wokaun, A., & Ernst, R. R. (1977) Chem. Phys. Lett. 52, 407-412. Wiithrich, K. (1986) NMR of Proteins and Nuclear Acids, Wiley, New York. Wiithrich, K., & Wagner, G. (1975) FEBS Lett. 50, 265-268. Zhang, 0., Kay, L. E., Olivier, J. P., & Forman-Kay, J. D. (1994) J. Biomol. NMR 4, 845-858. Zhu, G., & Bax, A. (1990) J. Magn. Reson. 90, 405-410. Zhu, L., Kemple, M. D., Landy, S. B., & Buckley, P. (1995) J. Magn. Reson. 3109, 19- 30. Zuiderweg, E. R. P., & Fesik, S. W. (1989) Biochemistry 28, 2387-2391. Zuiderweg, E. R. P., McIntosh, L. P., Dahlquist, F. W., & Fesik, S. W. (1990) J. Magn. Reson. 86, 210-216. Zuiderweg, E. P. R., Petros, A. M., Fesik, S. W., Olejniczak, E. T. (1991).]. Am. Chem. Soc. 113, 370-372. Zwanzig, R. (1961) Phys. Rev. 124, 983-992. CHAPTER 2 Structure-Function Relationships of Cellular Retinoic Acid Binding Proteins (Reprinted with the permission of the American Society of Biochemistry and Molecular Biology) 87 88 1115 Juuaxru. or 8101.04.21ch Crtsunmrv O 1997 by The American Society for Biochemistry and Molecular Biology. Inc. Vol. 272. No. 3. Issue ofJanuary 17, pp. 1541-1547. 1997 Printed in USA. 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” (Received for publication, August 1, 1996, and in revised form, September 27, 1996) Lincong Wang, Yue Li, and Honggao Yanil: From the Department of Biochemistry, Michigan State University, East Lansing, Michigan 48824 It has been suggested that electrostatic interactions are critical for binding of retinoic acid by cellular reti- noic acid-binding proteins (CRABP-I and CRABP-II). However, the roles of two conserved arginine residues (Art-111 and Arg-l3l in CRABP-I; Arg-lll and Arg-132 in CRABP-II) that interact with the carboxyl group of ret- inoic acid have not been evaluated. A novel competitive binding assay has been developed for measuring the relative dissociation constants of the site-directed mu- tants of CRABPs. Arg-lll and Arg-l32 of CRABP-II were replaced with methionine by site-directed mutagenesis. The relative dissociation constants of R1 11M and R132M (K4 (auras/Ks (CRABP-fl) ”“1 Kd (army/Kdtcmpanl were determined to be 40-45 and 6-8, respectively. The ring protons of the aromatic residues of the wild-type CRABP-II and the two mutants were sequentially as- signed by two-dimensional homonuclear NMR in con- junction with three-dimensional heteronuclear NMR. Detailed analysis of the nuclear Overhauser effect spec- troscopy spectra of the proteins indicated that the con- formations of the two mutants are highly similar to that of the wild-type CRABP-II. These results taken together showed that Arg-ll l and Arg-l32 are important for bind- ing retinoic acid but contribute to the binding energy only by ~2.2 and 1.2 kcal/mol, respectively. In addition, the relative dissociation constant of CRABP-II and CRABP'I (Kd (CRABP-TD’Kd (cmp_[)) was determined to be 2-3, in close agreement with that calculated using the apparent K, values determined under the same condi- tions by fluorometric titrations. Retinoic acid (RA).1 a hormonally active metabolite of vita- min A, has profound effects on cell growth, differentiation, and morphogenesis. Two types of proteins have been found to bind RA: nuclear retinoic acid receptors (RARs and RXRs) and cel- lular retinoic acid-binding proteins (CRABPs). RARs and RXRs are RA-activated transcriptional factors that regulate expres- sion of target genes (1). Although the physiological roles of ‘ This work was supported by funds from the REF Center of Protein Structure and Design and the Cancer Center at Michigan State Uni- versity. The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked ‘advertisement' in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. 3 To whom correspondence should be addressed. Tel.: 617-353-8786; Fax: 517-353-9334; E-mail: yan®nmr1.bch.msu.edu. ‘The abbreviations used are: RA, all-trans-retinoic acid; CRABP. cellular retinoic acid-binding protein; DQF-COSY, double quantum- filtered correlation spectroscopx NOE. nuclear Overhauser effect; NOESY, nuclear Overhaqu effect spectroscopy; RAR, retinoic acid receptor; RXR, retinoid X receptor; TOCSY, total correlation spectros- coPY; WT. wild-type CRABP-II; MT, CRABP-Il mutant. This paper is available on line at hitp://www-ibc.stanford.edulibcl CRABPs are not clear at. present, they are thought to be in- volved in cellular transport and metabolism of RA (2). Two isoforms (CRABP-I and CRABP-II) have been characterized. Both CRABP-I and CRABP-II bind specifically to all-trans- retinoic acid, but they differ in several respects: (1) CRABP-I has higher affinity for RA than CRABP-II (3—6); (ii) CRABP-I is expressed in many adult. mouse and human tissues, but the expression of CRABP-II is limited to skin (7, 8); and (iii) RA stimulates expression of CRABP-II but not that of CRABP-I (8). It appears that the two isoforms may have distinct functions. The idea is supported by the fact that. the sequence identity of human and mouse CRABP-I (99.3%) or human and mouse CRABP-II (93.5%) is much higher than the sequence identity (73.7%) between the two isoforms from the same source. CRABPs are members of a family of intracellqu lipid-bind- ing proteins that bind small hydrophobic molecules such as retinoids and fatty acids (9). The family of proteins includes CRABPs, cellular retinol-binding proteins, fatty acid-binding proteins, P2 myelin protein, a mammary gland protein, and gastropin. The structures of 11 different intracellular lipid- binding proteins, including CRABP-I and CRABP-II, have been determined by x-ray crystallography (9-11). Although the se- quence identity among intracellular lipid-binding proteins is rather low (~20%), these crystal structures are remarkably similar with respect to backbone folding. They are composed of two nearly orthogonally packed five-stranded B-sheets and two short a-helices. The ligand binding pocket in each protein is deep inside the interior of the B-barrel formed by the two B-sheets. The helix-turn-helix motif is located at the ligand entrance. These proteins apparently lack a true hydrophobic core that is important for folding and stability of many other proteins (12). Electrostatic interactions are thought. to play major roles in binding of RA by CRABPs because the two proteins bind only retinoic acid but not. retinol and retinal. No arginine residues (Arg-lll and Arg-131 in CRABP-l; Arg-lll and Arg-132 in CRABP-II) have been identified by crystallography to interact with the carboxyl group of the bound RA (10). However, previ- ous site-directed mutagenesis studies suggest that both Arg- 111 and Arg-131 are critical for CRABP-I to bind RA (13) but that only Arg-132 is important for CRABP-II to bind BA (14). Because of the limitations of the current methods for meas- uring RA binding, the dissociation constants of the critical mutants have not been measured. Furthermore, the conforma- tions of the CRABP-I mutants have been characterized only by CD, and the conformations of the CRABP-II mutants have not been characterized. Thus, the contributions of the electrostatic interactions between the arginine residues and RA to binding are still uncertain. We are interested in studying the quantitative structure- function relationships of CRABPs with respect to binding of 1541 89 1542 RA. In this paper, we report a novel competitive binding assay developed for measuring the dissociation constants of the site- directed mutants of CRABPs. We have used the novel method to evaluate the contributions of both Arg-lll and Arg-132 of CRABP-II to binding of RA in conjunction with sitedirected mutagenesis and NMR. The results show that like CRABP-I, both Arg-lll and Arg-132 in CRABP-II are important for bind- ing of RA, contrary to the results of Chen et a1. (14). However, Arg-lll and Arg-132 contribute to the overall binding energy only by ~2.2 and 1.2 kcal/mol, respectively. EXPERIMENTAL PROCEDURES Materials—Nonredioactive RA was purchased from Sigma. [11,12- ’H]RA was purchased from DuPont NEN. A DNA sequencing kit was obtained from US. Biochemical Corp. Enzymes for recombinant DNA experiments were purchased from Life Technologies. Inc., or New Eng- land Biolabs. Other chemicals were analytical or reagent grade from commercial sources. Cloning and Expression—The cDN As encoding human CRABP-I and CRABP-II were kindly provided by Dr. Anders Astrom (8). 'Ihe genes were subcloned into the expression vector pET-17b (Novagen) by po- lymerase chain reaction. The primers for polymerase chain reaction were: 5’-GGAA'!'1‘CCATATGCCCAAC’I'I‘CGCCGGC-3' (forward, CRABP-I), 5‘-CGGGATCCTCA'I'I‘CCCGGACATAAA’I'I‘C.3' (reverse, CRABP-I), 5‘-CGGGATCCATATGCCAAAC'ITCTCTGGCAACTG-3' (forward, CRABP-II), and 5'-GGAA'I'I‘C’I‘CAC'I‘CTCGGACGTAGAC-3' (reverse, CRABP-II). To ensure that there were no undesired mutations in the amplified genes, they were sequenced by double-stranded DNA sequencing from both orientations. The expression constructs were then transformed into the bacterial strain BL21(DE3)pLysS (15). The trans- formed bacterial cells were grown at 37 'C in a LB agar plate containing both ampicillin (100 ng/ml) and chloramphenicol (25 pg/ml). Expression of CRABPs was verified by growing the colonies in 5 ml of LB medium followed by induction with isopropyl-l-thio-R-ego' ‘ r, " and SDS-polyacrylamide gel electrophoresis of the harvested cells. Site-directed Mutagenesis—The oligonucleotides for making human CRABP-II mutants were 5'-TCGTGGACCATGGAACTGACCA-3' (RlllM) and 5'-GTGTGCACCATGGTCTACGTC-3' (R132M). The mu- tants were generated by the method of Kunkel (16) and screened by DNA sequencing. In order to ensure that there were no unintended mutations in the mutants, the entire sequences of the mutated genes were determined. Protein Purification—All proteins were purified by the same proce- dure as described below. 100 ml of LB medium containing 100 ug/ml of ampicillin and 25 pg/ml of chloramphenicol was inoculated by a small piece of a frozen seed culture and incubated at 37 'C with vigorous shaking (200 rpm) overnight. It was then used to inoculate 4 liters of LB medium containing both antibiotics and grown at 37 'C. When Am of the culture reached 0.6-0.8, isopropyl-l-thio-fi-D-galactopyranoside was added to a final concentration of 0.4 mu. The culture was allowed to grow further at 30 'C for 3-5 h. The bacterial cells were then har- vested by centrifugation and suspended in 100 ml of buffer A (10 mil Tris-HCl, 1 ma dithiothreitol, pH 8.2). The suspension was aonicated on ice and centrifuged (27,000 x g) at 4 'C for 30 min. The supernatant was applied to a DEAE-cellulose column equilibrated with the same buffer. The column was washed with buffer A until Am of the eluent was less than 0.05. Elution of the column was achieved by a linear N aCl gradient (0—200 mil in buffer A) and monitored by A,” and 15% SDS-polyacrylamide gel electrophoresis. The fractions containing CRABP were pooled and concentrated by an Amicon ultrafiltration cell using a YM 10 membrane. The concentrated protein solution was cen- trifuged (27,000 x g) for 20 min. The supernatant was loaded onto a Sephadex G-50 column equilibrated with phosphate-buffered saline (4 mt! NaH,PO,, 16 mil Na,HPO., 150 mM NaCl. pH 7.3) and eluted with the same buffer. The fractions containing CRAB? were located byA,.o and 15% SD8polyacrylamide gel electrophoresis. The fractions of greater than 99% purity were pooled and concentrated. The protein solution was dialyzed against doubledistilled water and lyophilized. Fluorometric Titration—Fluorescence binding assays were carried out by a procedure modified from that of Cogan et al. (17) using a Hitachi 4500 fluorometer. Briefly, CRABPs were dissolved in phos- phate-buffered saline. The concentrations of the protein stock solutions were measured by Am. The absorption coefficients used were 19,868 is“ cm" for CRABP-I and 19,480 M" cm‘l for CRABP-II determined by the method of Gill and Hippel (18). RA stock solutiOns were prepared in absolute ethanol. The concentrations of the RA stock solutions were Cellular Retinoic Acid-binding Proteins determined by A,“ using the absorption coefficient of 45,000 It" cm". The final ethanol content for each titration was kept less than 2%. The samples were excited at wavelength of 283 nm with a slit width of 2.5 nm. The emission wavelengths were 330 nm for CRABP-I and 338 nm for CRABP-ll with a slit width of 20 nm. The excitation shutter was closed between measurements. The inner filter effects were not cor- rected because they were negligible at the protein and RA concentra- tions we used in the experiments. The data were analysed by nonlinear least square fit to Equation 1. F 1 (F. l)x(P,+R,+K,,—JP,+R,+K,,)’-4P,R, r,‘ i F,‘ 2?. ) (Eq. 1) where F is the observed fluorescence, I", is the fluorescence of the bound CRABP. F,is the fluorescence of the free CRABP. P, is the total protein concentration, and R, is the total RA concentration. Competitive Binding Assay—The assay was carried out in a Spectral Por equilibrium dialyzer at room temperature. The two compartments of each dialysis cell were separated by a semipermeable membrane with a molecular mass cut-off of 6—8 kDa. One compartment was filled with the wild-type CRABP-ll (1 ml), and the other was filled with a CRABP-Il mutant or the wild-type CRABP-I. The buffer used was phosphate-buffered saline plus 5 mil dithiothreitol. An equal amount of [’HIRA (100 nu) was added to each compartment. The protein concen- trations in both compartments were much greater (at least 20—fold) than the RA concentration. 100 pl of samples was taken from the two com- partments after various times of incubation at room temperature, mixed with 5 ml of scintillation fluids, and counted by a liquid scintil- lation counter. The equilibria in the two compartments that contained the wild-type CRABP-Il and mutant proteins can be described by Equa- tions 2 and 3 (similar equations can be written for CRABP-I and CRABP-II). [WTIRA] Km " W “5“. 2’ IM'I‘ RA Klilfl‘l . [MT F RAll (Eq. 3) where WT. MT, WT-RA, and MT-RA represent the wild-type CRABP-II, a CRABP-II mutant, and their RA complexes, respectively. Therefore. ’31“ .. L______M I I w I'm Km—n [WTIMT ' RA] Since the concentrations of the proteins were much greater than their respective dissociation constants and the concentration of RA, [WT] ~ lw'rlm... [MT] ~ (mm... [WT'R-Al » [RA]. and [MT'RM >9 [M- Then the relative dissociation constant can be calculated by Equation 6, Km .. immen- (Eq 5) ann [WTlianCirr ' where Cw... and Cm. are the measured radioactivities of the two com- partments containing the wild-type CRABP-II and the mutant, respec- tively. It turned out that the system could not reach equilibrium in 2 days, presumably because of few free RAs in solution to diffuse across the membrane. Since RA is not stable even in the dark, the assay was redesigned to match the equilibrium conditions by varying the ratio of the protein concentrations of the wild-type and the mutant ([MTIWJ lWT],,,.,). Thus, the concentration of the mutant was varied while keeping the concentration of the wild type at ~2 use. Initially, the concentration of the mutant was increased in an exponential manner (e.g. 2. 20, 200 nM). Then it was varied in a small range. Since an equal amount of RA was added to the two compartments of the dialysis cell, the two compartments should have the same RA concentration and radioactivity at the beginning of the assay. If [MT]«,uuillii‘i'l‘lh,“| e Kama/K.,, «am: there would have a net transfer of RA across the semi- permeable membrane separating the two compartments. Thus, the radioactivity counts of the two compartments (CW,- and Cm) would differ after incubation for a certain period. When [M'I‘lm/Ml‘lw < Karim/Knurn- then er ‘ Cm > 0- When maul/masts! > Kenn-l KltW'hv then CWT ‘ Ctr-r < 0- When mun/mural ' Kaiser/Karena. than Cm " Cm - 0. NMR Spectroscopy—All NMR measurements were performed at 32 'C on a Varian VXR-SOO spectrometer operating at a proton fre- quency of 500 MHz. The proteins were dissolved in 20 mil sodium phosphate. pH 7.5 (direct pH meter reading). 100 mu NaCl, 5 mu dithiothreitol in D20. The protein concentration was ~2 mu. The spec- (Eq. 4) 90 Cellular Retinoic Acid-binding Proteins tral width was 7200 Hz. For wild-type CRABP-II, one phase-sensitive DQF-COSY spectrum (19), two MLEV-l'l clean TOCSY spectra (20-22) with mixing times of 20 and 40 ms and one NOESY spectrum (23. 24) with a mixing time of 150 ms were acquired. For RlllM. one DQF- COSY spectrum, three TOCSY spectra with mixing times of 20, 40. and 75 ms and one NOESY spectrum with a mixing time of 150 ms were acquired. For R132M. one TOCSY spectrum with a mixing time of 40 ms and one NOESY spectrum with a mixing time of 150 ms were acquired. All of the spectra were acquired in the hypercomplex mode with stand- ard phase cycling schemes. The data were usually acquired with 2048 complex points in the t2 dimension and 256 complex points in the t1 dimension. 96 transients were collected for each FID. Data processing was performed on a Sun Sparc 10 station using VNMR software from Varian. The time domain data were zero-filled once and multiplied by shined sine bell or Gaussian functions before Fourier transformation in both dimensions. Base lines were corrected in t2 dimensions using a 5-order polynomial. Chemical shifts were referenced to internal sodium 3-(tn'methyl silyl)-propionate-2.2.3.3-d.. RESULTS Fluorometric Titration—Probably because of the simplicity of the method. fluorometric titration has been frequently used for measuring the affinities of CRABPs for RA. However, there are large discrepancies in the K., values obtained by the method. These could be due to variations in the assay method- ology and/or conditions. Some problems with the method have been recently discussed (2. 4. 6). In order to compare the ligand binding properties of CRABPs under the same conditions, we carried out fluorometric titration experiments. The results are shown in Fig. 1. The apparent K, values obtained by nonlinear fitting were 0.63 nu for CRABP-I and 1.9 nu for CRABP-II. While our work was in progress, Norris et al. (6) reported the apparent Kd values of mouse CRABPs measured by an im- proved fluorescence titration method. The major changes were the lowering of protein and RA concentrations and the use of nonlinear least square fitting for data analysis. These modifi~ cations were in line with our approach. Our fluorometric titra- tion data were analyzed by nonlinear least square fitting. The protein concentrations we used in the assays (~50 nM) were the lowest among those used in previous studies except those of Norris et al. (6). We did not include the additive gelatin used by Norris et al. (6) in the assays. The apparent Kd values deter- mined by us are in close agreement with those (0.4 all for CRABP-l and 2 nu for CRABP-II) reported by Norris et al. (6). Competitive Binding Assay—In order to measure the affini- ties of the mutants of CRABPs for RA. we developed a novel competitive binding assay. We first used the method to meas- ure the relative affinity of CRABP-I and CRABP-II for RA. The result is shown in Fig. 2A. When the ratio of the concentrations of the two proteins ([CRABP-IH/[CRABP—ll) was 52, there was a not transfer of RA from the compartment containing CRABP-ll to the compartment containing CRABP-I. When the ratio of the concentrations of the two proteins was 23, there was a net transfer of RA from the compartment containing CRABP-I to the compartment containing CRABP-ll. As de- scribed under “Experimental Procedures,‘ the relative K., of the two proteins lies between the points with opposite net trans- fers. Thus the Kd of CRABP-II relative to that of CRABP—I (K.,(CRABP-IIVKJCRABP-D) is 2—3. It is in close agreement with the results of the fluorometric titration studies. We then used the competitive binding assay to measure the K., values of RlllM and R132M relative to that of the wild-type CRABP-II. The results are shown in Fig. 2 (B and C). In the case of R111M (Fig. 28), there was a net transfer of BA from the compartment containing RlllM to the compartment containing the wild- type CRABP-II when [R111Ml/[WT] s 40. When [R111MIIIWTl z 45. there was a net transfer of BA from the compartment containing the wild-type protein to the compartment contain- ing R111M. Therefore. the relative affinity of R111M for RA 1543 1.0- o.._ A CRABP-l a: ‘ 8 he m 03- m o. 5 2 0.6- E . 1.3 0.44 -. a are- a: ~xe-.._-..______. 0.2- 6 2'0 ' 4'0 ' e’o ' 33 T 100 Retinoic Acid (nM) 1.0-I 0 \. B CRABP-Il 5 03- " 0 m 9 ‘ 'b. O g os~ o ‘ in“ .. 0.4-4 “N... é ‘ k“.“~— .--—._.D 02 1 V I " I V I ' I ' I O 20 40 60 80 100 Retinolc Acid (nM) Fla. 1. Fluorometric titration of wild-type CRABP-l (A) and CRABP-II (B). The concentrations of CRABP-l and CRABP-II were 40 and 60 ml. respectively. (Kd‘RI‘lu/de is 40-45. As shown in Fig. 20, the relative affinity of R132M for RA (Kama/Kay") is 6—8. Sequential Resonance Assignment of the Ring Protons of the Aromatic Residues of the Wild-type CRABP-II and Site-directed Mutants—Since perturbations in the ligand binding property of a mutant may be due to conformational changes caused by the mutation (25), we characterized the conformations of the mu- tants by NMR. A prerequisite for NMR structural analysis is sequential resonance assignment. Currently, we are in the process of making total sequential resonance assignments of the wild-type CRABP-Il and R111M by multidimensional multinuclear NMR. We have sequentially assigned the ring protons of aromatic residues of the wild-type and mutants based on the homonuclear two-dimensional NMR (DQF-COSY. TOCSY, and N OESY) and heteronuclear three-dimensional NMR.3 The chemical shifts of the ring protons are listed in Table l. The sequential resonance assignment is labeled in Fig. 3 for R132M. In comparison with the chemical shifts of the wild-type ring protons, there are quite a few chemical shift changes in the ring protons of the mutants. Some chemical shift changes can be easily rationalized, while the causes for other chemical shift changes are not obvious. For example, the guanidino group of Arg-132 stacks on the aromatic ring of ' L. Wan. Y. Li, and H. Yan. unpublished results. 91 Cellular Retinoic Acid-binding Proteins 200+ . 4004 40m [CRABP-lMCRABP-l] [mural/[camera [R132M1/[CRABP-ll] Fla. 2. Competitive binding assays for measuring the relative dissociation constants of CRABP; The relative radioactivity inA is the radioactivity count of the compartment containing CRABP-l minus that of the compartment containing CRABP-II. The relative radioactivity in B is the radioactivity count of the compartment containing CRABP-ll minus that of the compartment containing R111M. 'flie relative radioactivity in C is the radioactivity count of the compartment containing CRABP-Il minus that of the compartment containing R132M. Tan Chemical shifts of the ring protons of the aromatic residues as I of wild-type CRAB? II. RIIIM mutant. and R132M mutant The underlined values are the mutant resonances that differ by >0.02 ppm from the corresponding resonances of the wild-type CRABP-II. Residue mm”. ems-n sum 3132M Pin-3 7.60‘ 7.34‘ 7.152‘ ‘L§_1‘ g 7.46! 7.60- my 7.52‘ Pile-15 7.42- 7.02‘ 7.11‘ 7 39- LB" 110‘ 7.40- 1._1_1‘ 7.11c Pine-50 6.66' 7.26‘ 6.99‘ 6.68' 7.24‘ 6.99‘ 6.64- 7.240 6.99' Phe-65 6.42- 6.70‘ 6.48‘ §-3_8‘ 6.70‘ 6.44‘ 6.42‘ 6.69“ w Phe-‘ll 7.48‘ 7.00‘ 6.70‘ 7 44- 6.99‘ m 7.48‘ 7.00‘ 6.69‘ I‘m-.51 7.12‘ 6.79' 7.124 6._74_r 7.134 6.78‘ Tyr-134 6.98“ 6.65‘ 7.00!' §.8_6‘ 7.034 w- Trp-‘l 7.25! 7.00! 7.41‘ 7.30‘ 7_._2_2_’ 6.99! 7.39‘ 7.29" 7.25! 6.99! 743* 730‘ W7 8.16’ 7.13! 7.39‘ 7.80‘ s 16’ M 7.39‘ 116‘ 8.17’ 19;! 7.39“ 7.80‘ Trp-B'l" 8.07’ 7.05! 7.33“ 7.76‘ Trp-109 7.0V 6.63! 6.61“ 6.87‘ 'L_l’ gm! 6.59" 6.89‘ 7.09’ 6.62! 6.61“ 6.87‘ '" 2.68. 3.514. and 4B of phenylalanine. respectively. "‘ 2.6!! and 3.5}! of tyrosine. respectively. "‘ 4H. 5H. 6H. and 7B of tryptophan. respectively. ’ Tip-87 has two sets of resonances because of conformational heterogeneity. Phe-15 and is hydrogen-bonded to the hydroxyl group of 'lyr- 134 in the crystal structure of the wild-type CRABP-II (10). Removal of the guanidino group is likely to be the cause for the chemical shift changes in Phevl5 and Tyr-134 of R132M. The crystal structure also reveals that the guanidino group of Arg- 111 interacts with Trp-109 through an amide/aromatic-ring hydrogen bond (10). Thus. removal of the guanidino group of Arg-lll may cause the chemical shift changes in M109 of R111M. Some small chemical shift differences may result from changes in pH. temperature. and other conditions. They are not necessarily indicative of conformational changes. Conforma- tional changes are best characterized by analysis of NOESY spectra as described below. Interestingly. 4H and 7H of the Tip-87 ring protons in R132M have two sets of NMR signals (Fig. 3). indicating that there is localized conformational het- erogeneity and that the 'hp ring is in two distinct conforma- tional states. The lifetimes of the two conformational states are at least 50 ms. The ratio of the pepulations in the two states (major and minor forms) is about 4:1. 1H of the Tip-87 ring protons in the wild-type protein and RlllM also show two distinct resonance signals (data not shown). indicating that the two proteins also have localized conformational heterogene- ities. The lifetimes of the two conformational states are at least 40 ms in the wild-type protein and at least 33 ms in R111M. The ratio of the major and minor forms is about 3:1 in the wild-type protein and about 4:1 in RlllM. Conformational dynamics may be important for RA to get into the deep binding pocket of the protein. However. Tip-87 is not at the RA-binding pocket and is far away from the ligand entrance. The functional significance of the conformational heterogeneity is not clear. Conformational Comparison of the Wild-type CRABP-II. RIIIM. and R132M—The aromatic-aromatic NOEs are shown in the lower panels of Fig. 4 (wild-type CRABP-II). Fig. 5 (R111M). and Fig. 6 (R132M). Some aromatic-aliphatic NOEs are shown in the upper panels of the figures. Interresidue NOEs are labeled. because they are indicative of the conforma- tion of a protein. The interresidue NOE peaks of the three proteins are the same except for some minor variations in peak intensity. The observed NOEs are also consistent with the atomic distances measured from the coordinate of the crystal structure of CRABP-II complexed with RA. The identities of the aliphatic protons shown in the upper panels of these figures are not known at present. However. qualitatively. the three pro- teins have very similar aromatic-aliphatic NOE patterns. The results of the N OESY analysis suggest that the conformations of R111M and R132M are highly similar to that of the wild-type protein. Thus. the decreases in affinity for RA of the two muo tants are most likely due to the disruption of the interactions of the arginine residues with RA by mutagenesis. Arg-lll and Arg-132 contribute to the binding energy by ~2.2 and 1.2 kcal/mol. respectively. DISCUSSION Competitive Binding Assay—Two types of methods have been in general use for measuring binding of RA to CRABPs: fluorometry and radiometry. The radiometric method involves 92 Cellular Retinoic Acid-binding Proteins , "I . . .. . ‘I . 4 'l . 0 n (on) FIG. 3. TOCSY spectrum of the aromatic protons of R132M at 32 '0 showing two sets of resonances for e mixing time was 40 ms. All of the cross-peaks have been sequentially assigned. The chemical shifts of the resonances are listed in Table 1. separation of bound from free RA by dextran—coated charcoal. gel filtration, and other means. Substantial loss of bound li- gand during the separation process makes the method unsuit- able for measuring the dissociation constants of site-directed mutants with greatly decreased affinity for RA. The very lim- ited solubility of RA in water (~200 nM) (26) also makes the fluorometric method inapplicable for determining the dissocia- tion constants of these mutants. Studies of the quantitative structure-function relationships of CRABPs have been ham- pered by the lack of methods for measuring the affinities of site-directed mutants for RA (13, 14). We have developed a novel competitive binding assay for measuring the affinities of site-directed mutants for RA relative to that of the wild-type CRABP. The essence of the method is to monitor the competi- tion between a mutant and the wild-type protein for binding of limited RA. The method uses an equilibrium dialyzer. The two compartments of each dialysis cell are filled with the wild-type and mutant proteins. respectively. The absolute concentration of RA is not important as long as the concentration of free RA is much smaller than that of bound RA. There is no need to separate bound from free RA by dextran-coated charcoal, gel filtration. and other means. The transfer of RA fi'om one com- partment to the other is determined by measuring the radio- activities of the samples taken from the two compartments. The direction of the net transfer is dependent on the relative affinity of the proteins and the ratio of the protein concentra- tions of the two compartments. If the solubility of a mutant is the same as that of the wild-type CRABP-II (~2 mM). then the ratio of the concentrations of the wild-type and mutant proteins can be varied by more than 1.000-fold. and a relative K.,, of as large as 1000 (Kd(m/Kd (w-n = 1000) can still be measured by the method. More importantly. determination of the relative dissociation constant of a point mutant is sufficient for estimat- ing the energetic contribution of the amino acid residue to ligand binding (AAG= RTln(Kd (MT/KtHWT) In addition to vitamin A, many other hydrophobic molecules such as vitamin D and steroid hormones play vital roles in a variety of cellular processes. Studies of the stmcture-function relationships of the receptor proteins for these bioactive hydro— phobic molecules are of fundamental interest and biomedical 1545 0.: 0.0 7.0 7.6 7.4 7.2 7.0 6.6 6.6 6.6 6.2 '2 (vs-I) I rout I I I I I I I I l I s.z 6.6 7.s 7.6 7.: 7.: 7.0 6.0 6.6 6.6 6.: n (vs-I) FIG. 4. Parts of the two-MHz NOESY spectrum of the wild-type CRABP-II at 32 ‘C. The mixing time was 150 ms. Only the interresi- due NOE: are labeled. The identities of the NOEs are as follows. A, Phe-65 2. 6H to Phe-71 2, 6H; 8 Phe-65 2. 6H to Phe-71 3. 5H; C. Rho-65 4H to’l‘rp-1096;H D. 4Phat-35 3.6H99to'l‘I-p~106;H E. Phe-BO 2.6Ht0 87 7H;F,Phe-71to-’I‘rp 109 7H; G. Phe-65 4H to Tip-109 7H; H, Phe-65 2. 6H to 'l‘rp-109 7H; 1. Phe-60 3. 5H to Phe-65 4H; J. 'l‘rp-8'I 5H to Phe-3. ~ Trp-;877HM.Phe-504Hto'l‘rp— 877H ;.N Phe-50 2. 6H to 'I‘rpv87 6H; P. Phe-50 2. 6H to [I‘ll-2.611; Q. 'h'p-S‘ISH Phe-3 2. 6H; R. Phe-S 4H to ’I‘rp-87 significance. The method can also be used for measuring the relative dissociation constants of the site-directed mutants of these important proteins and may facilitate these studies. Relative Afl'inity of CRABP-I and CRABP-II for RA— CRABP-I and CRABP-II differ in affinity for RA, expression pattern, and regulation. Although it is generally accepted that CRABP-I has a higher affinity for RA than CRABP-II. it has been difficult to determine precisely the relative affinity of the two proteins. A wide range of apparent Kd values have been reported. The apparent K., values for purified CRABP-I are in the range of 0.4-39 nM (4—6, 13. 27-29). The apparent K., values for purified CRABP-Il are in the range of 2-64 nu (3-6, 14) The best estimation of the relative affinity of the two proteins for RA was made by Fiorella et al. (5) In their study, RA was first incubated with CRABP- II and then mixed with CRABP-I I. The protein RA complexes (CRABP-I- -RA and CRABP-II~RA) were separated by chromatofocusing. The result indicated that CRABP-II has about 3-fold lower affinity for RA than CRABP-I. More recently. Norris et al. (6) compared the affinities of CRABP-I and CRABP-II for RA by fluorometric titration under the same conditions. The results showed that 93 1546 2.1 ° 3"? l ' £23 0 fit as no 5232 a: $9.63? .. ~ 3:? v o 0 so?» 0 ° . 0.: 0.0 7.0 7.6 7.0 7.: 7.0 6.0 s.s 0.0 6.: n (0000) ' uuuuu ' I ' i""‘ I | 0:: I 0.0 1.0'1it '74 '7.0 1.0 6.0 I0.‘ I: (00-) FIG. 5. Parts of the GOO-MHz NOESY spectrum of RlllM at 32 ‘C. The mixing time was 160 m0. Only the interresidue N020 are labeled. The identities of the N030 ars given in the legend to Fig. 4. 0.0 0.3 CRABP-I has 4~fold higher affinity for RA than CRABP-II (although a 10-fold stronger affinity of CRABP-I for RA. com- pared with CRABP-II. was stated in the paper. apparently due to an error in calculation). The apparent dissociation constant or CRABP-II relative to that Of CRABP-I (K4 (cmp.u/ Kamp.n) is 3, based on our fluorometric measurements. The results of our competitive binding assays indicated that the relative apparent dissociation constant of CRABP—II (K4 (CRABPoll/Kd(cmP-D) is “.1118 mats Of our fluorometric measurements and competitive binding studies are in close agreement. They are also in line with the results of Fiorella et al. (5). The relative affinity of CRABP-I and CRABP-II meas- ured by Fiorella et al. (5) may be the upper limit. because some RA bound to CRABP-Il could be preferentially lost during separation of the protein'RA complexes by chromatofocusing. The difference in affinity for RA between CRABP-I and CRABP-II is rather small (1—2-fold). Both Ami-111 and Amt-132 of CRABP-II Are Important for Binding of RA—Arg-lll. Arg-132. and 'l‘yr-134 of CRABP-II have been predicted to be involved in binding of RA on the basis of homologous modeling (13, 30, 31). The crystal structure of CRABP-II in complex with RA reveals that one of the carbox- ylate oxygens of RA is hydrogen-bonded to the guanidino group Arg-132 and the hydroxyl group of 'l‘yr-134. and the other carboxylate oxygen of RA interacts with Arg-lll via a water molecule (10). We have substituted Arg-lll and Arg-l32 with methionine. Our biochemical and structural characterizations show that guanidino groups ot'Arg-lll and Arg-132 contribute to the binding energy by ~2.2 and 1.2 kcal/mol, respectively. The results indicate that both Arg-lll and Arg-132 are indeed Cellular Retinoic Acid-binding Proteins 0.3 0.0 '2 (son) - L——.-___..-.-- ..._. - _.--.————- _- 0.0 0.0 1.0 1.0 1.0 1.3 1.0 0.0 0.0 0.0 0.3 '2 (0’) FIG. 6. Parts of the mun: NOESY spectrum of R132M at 32 'C. The mixing time was 150 m0. Only the interresidue NOEs are labeled. The identities of the NOEs are given in the legend to Fig. 4. involved in binding of RA and that Arg-lll contributes more to the binding energy than Arg-132. Fluorometric titration also suggests that R132M has higher afiinity for RA than RlllM (data not shown). Zhang et al. (13) have replaced the corre- sponding residues in CRABP-I with glutamine (RlllQ and R131Q). Quantitative comparison of our results with those of Zhang et al. (13) is not possible, because the dissociation con- stants of their mutants have not been determined. Qualita- tively, the ligand binding properties of our mutants are similar to those of the corresponding CRABP-I mutants. Zhang et al. (13) observed that both the wild-type CRABP-I and R131Q can induce CD of RA in the wavelength range between 320 and 360 am. but R111Q cannot induce the CD signal. The results of the CD experiments with CRABP-I are consistent with our meas- urements of the energetic contributions of Arg-lll and Arg-132 to the binding of RA. Chen et al. (14) have mutated Arg-lll of CRABP-II to alanine (R111A) and Arg-132 of CRABP—II to alanine and glutamine (R132A and R132Q). In contrast to our R11 1M mutant, the apparent dissociation constant of their R111A mutant is very similar to that of the wild-type CRABP-II (only 50% higher than that of the wild type). The causes of the discrepancy are not clear, but several possibilities can be ruled out. (i) It is unlikely that there are mutations other than substitution of Arg-lll with a methionine in our RlllM mutant. We have sequenced the entire gene after mutagenesis and subcloning into the expression vector. (ii) Steric hindrance is unlikely to be the cause. because the side chain of methionine is smaller than that of arginine. (iii) Although small conforma- tional changes cannot be ruled out. NMR characterizations suggest that RlllM is properly folded, and its conformation is 94 Cellular Retinoic Acid-binding Proteins very similar to that of the wild-type protein. Furthermore, RlllM has been crystallized. The preliminary crystal struc- ture at 2.2-A resolution shows that the structure of RlllM is indeed very similar to that of the wild-type protein.3 The ligand binding property of our R132M mutant is qualitatively similar to those of the R132A and R132Q generated by Chen et al. (14). However, quantitative comparison is again not possible, be- cause the apparent dissociation constants of the two mutants have not been determined. The results of our mutagenesis studies appear to be partly inconsistent with the crystallographic data. The crystal struc- ture of holo~CRABP-II suggests that Arg-132 interacts directly with the carboxyl group of RA. whereas the interaction between Arg-lll and the carboxyl group of RA is mediated by a water molecule (10). Our results agree with the crystallographic ob- servation that both Arg-lll and Arg-132 are involved in bind- ing of RA. However, in contrast to the crystallographic data, our mutagenesis results indicate that Arg-lll is more impor- tant than Arg-132 for RA binding. But this is not necessarily inconsistent with the crystallographic studies. Since Arg-132 and Tyr-134 are hydrogen-bonded to the same carboxylate ox- ygen of RA, the loss of the hydrogen bond between Arg-132 and RA in R132M may be partially compensated by the interaction between Tyr-l34 and RA. On the other hand, the other carbox- ylate oxygen of RA only interacts with Arg-lll (albeit via a water molecule), the loss of the interaction in RlllM may not be compensated. Alternatively, in solution, the position of the carboxyl group of the bound RA may be slightly different from that observed in the crystals. It is noted that the RA binding cavity is much larger than necessary to accommodate the li- gand, and a large portion of the deep binding cavity is not occupied by RA. It appears that the fi-ionone ring of the bound RA is well fixed, but the isoprene tail and carboxyl group have room to move. The guanidino group of Arg-lll could be closer to the carboxyl group of RA than to that of Arg-132 in solution. Solution NMR study of holo-CRABP—II is in progress. CRABPs have very stringent retinoid specificity (2). The proteins bind only RA and reject both retinol and retinal. On the other hand, cellular retinol-binding proteins bind both ret. inol and retinal and reject RA. Amino acid sequence analysis reveals that the two conserved arginine residues in CRABPs are replaced with glutamine in cellular retinal-binding pro- teins. It has long been thought that the electrostatic interac- tions between the guanidino groups of the two conserved argi- nine residues and the carboxyl group of RA play major roles in binding of RA (31). Our mutagenesis studies indicate that Arg-lll and Arg-132 are indeed involved in RA binding, but ’ Ji, K., L. Wang, Y. Li, and H. Yan, unpublished results. 1547 the electrostatic interactions between the guanidino groups of Arg-lll and Arg-132 and the carboxyl group of RA contribute to the overall binding energy only by ~2.2 and 1.2 kcal/mol. Thus, amino acid residues that interact with the hydrophobic moiety of RA may be more important for binding of RA. Acknowledgment—We are indebted to Dr. Anders Astrom for provid- ing the cDNA clones of human CRABP-I and CRABP-II. We thank Drs. David G. McConnell and Yanling Zhang for critical reading of the manuscript and Dr. Xinhua J i for useful discussion. REFERENCES . Mangelsdorf, D. J., Umesono, K, and Evans. R. M. (1994) in The Retinoids: Biology, Chemistry, and Medicine (Sporn, M. 8., Rorberts, A. 3., and Good- man, D. S., eds) pp 319-349, Raven Press, New York 2. Ong, D. E., Newcomer, M. E.. and Chytil, F. (1994) in The Retinoids: Biology, Chemistry, and Medicine (Sporn, M. 8., Rorberts. A. 3.. and Goodman. D. S., eds) pp 283-317, Raven Press. New York . Bailey, J. S., and Siu, C.-H. (1988) J. Biol. Chem. 203. 9326-9332 . Fogh, K. Voorhees, J. J., and Astroin (1993) Arch. Biochem. Biophys. soo, 751-755 5. Fiorella, P. D.. Giguere, V., and Napoli, J. L (1993) J. Biol. Chem. 298. 21545-21552 . Norris, A. W., Cheng, L. Gigubre, V., Rosenberger, M., and Li. 3. (1994) Biochim. Biophys. Acta 1209. 10-18 . Giguere, V., Lyn, S., Yip, P., Siu, C.-H., and Amin, S. (1990) Proc. Natl. Acad. Sci. U. S. A. 87. 6233- 6237 Astrom, A., Tavakkol, A., Pettersson, U., Cromie, M., Elder, J. T., and Voo- rhees, J. J. (1991) J. Biol. Chem. 206, 17662-17666 . Banaszak, L, Winter, N., Xu, 2.. Bernlohr, D. A., Cowan, S. W., and Jones, T. A. (1994)Adv. Protein Chem. 45. 89-151 10. Kleywegt, G. J., Bergfors, T., Senn, H., be Motto. P., Gsell, 8.. Shudo. K., and Jones, T. A. (1994) Structure 2. 1241-1258 11. Thompson, J. R, Bratt, J. M., and Banesuk. J. (1995) J. Mol. Biol. 852. 433—446 12. Scapin, G., Gordon, J. 1., and Sacchettini, J. C. (1992) J. Biol. Chem. 987. 4253-4269 13. Zhang, J., Liu, Z.-P., Jones, T. A., Gierasch, 1.. M., and Sambrook, J. I". (1992) Proteins Slrucl. Funcl. Genet. 13. 87-89 14. Chen, L. X., Zhang, Z.-P., Scafonas, A., Cavalli, R. C., Gabriel, J. L, Soprano, K. J., and Soprano, D. R. (1995) J. Biol. Chem. 270, 4518-6525 15. Studier. P. W., Rosenberg, A. K., Dunn. J. J., and Dubendorfl'. J. W. (1987) Methods Enzymol. 1985. 60-89 16. Kunkel, T. A. (1985) Proc. Natl. Acad. Sci. U. S. A. 92. 488-492 17. Cogan, U., Kopelman, M., Mokady, S., and Shinitzky, M. (1976) Eur. J. Bio- chem. “. 71-78 18. Gill, 8. C., and Rippel, P. T. ( 1989) Anal. Biochem. 193, 319-326 19. Piantini, 0., Sorensen, O. W., and Ernst, R. P. (1982) J. Am. Chem. Soc. 10‘. 6800-6801 20. Braunschweiler, L, and Ernst. R. P. (1983) J. Magn. Reson. 53. 5521-5528 21. Bax, A., and Davis, D. V. (1985) J. Magn. Reson. 86, 355—360 Griesinger, C.. Otting, G., Wuthrich, K., and Ernst, R. P. (1988) J. Am. Chem. Soc. 110, 7870-7872 Jenneer, J., and Ernst. R. P. (1979) J. Chem. Phys. 71. 4548-9653 Macura, S., and Ernst. R. P. (1980) Mol. Phys. ‘1. 95—117 Tsai., M.-D., and Yan. H. (1991) Biochemistry 30. 6806-6818 Szuts, E. Z., and Harosi, F. l. (1991) Arch. Biochem. Biophys. 287. 297-304 Daly, A. K., and Redfem, C. P. P. (1988) Biochem. Biophys. Acta 995. 118-128 Fiorella, P. D., and Napoli, J. L. (1991) J. Biol. Chem. 906. 16572-16579 Ong, D. 8., and Chytil, P. (1978) J. Biol. Chem. 253, 4551-‘554 Cowan, S. W., Newcomer, M. S., and Jones, T. A. (1993) J. Mol. Biol. 8”. 1225-1246 . Jones, T. A., Bergfors, T., Sedzik, J., and Unge, T. (1988) EMBO J. 7. 1597-1604 b“ D‘ Q“ g 8883383? a one CHAPTER 3 Solution Structure of Type-ll Human Cellular Retinoic Acid Binding Protein by NMR Introduction One of the key questions concerning the family of iLBPs including CRABPs is how the ligands get in and out of their respective binding pockets, since they are encapsulated by the B-sheets and helix-turn-helix motif in most cases. The crystal structures of holo-CRABPI and holo-CRABPII (both in complex with all-trans-RA) have been determined (Kleywegt et al., 1994). In each structure, RA is buried deeply in the RA-binding pocket, with its carboityl group interacting with the side-chains of two arginines and one tyrosine at the bottom of the pocket. The B-ionone ring which is fixed snugly at the entrance of the RA-binding pocket with only one edge of the ring accessible to the solvent, is forced to adopt an unusual cis—like conformation. It appears that RA cannot enter or exit the deep binding pocket in the absence of major conformational changes in the protein. More recently, the crystal structure of apo-CRABPI has been solved (Thompson 95 96 More recently, the crystal structure of apo-CRABPI has been solved (Thompson et al., 1995). The ligand entrance of apo-CRABPI is slightly more open than that of holo- CRABPI and thus a bit more accessible to RA. Surprisingly, apo-CRABPI is a dimer in the crystalline state, held together by an intermolecular B-sheet. Dimerization has been suggested to be the mechanism that allows RA to enter or exit the RA-binding pocket and by which CRABPI interacts with RA-metabolizing enzymes such as cytochrome P4503. However, it is unclear whether apo-CRABPI is dimeric in solution. CRABPI has been studied by NMR (Rizo et al., 1994), but no solution NMR structure has been reported. In this chapter, the solution structure of human apo-CRABPII determined by NMR spectroscopy is described. The sequential assignments of the 1H, 13C and ”N resonances of apo-CRABPII were established by multinuclear multidimensional NMR experiments. The solution structure of apo-CRABPII was derived from 2,382 experimental NMR restraints. The solution structure of apo-CRABPII is similar to the crystal structure of holo-CRABPII, but significant conformational differences were observed, especially in the ligand entrance region. In comparison to holo-CRABPII, apo- CRABPII exhibits a concerted movement of the second helix, the BC-BD loop and the BE-BF loop, so that the ligand entrance of apo-CRABPII is greatly enlarged and readily accessible to RA. Furthermore, the ligand-binding pocket of apo-CRABPII is rather disordered. The results suggest that binding of RA induces significant changes in the conformation and dynamics of CRABPII. 97 Experimental Procedures Sample preparation. ['3C] Db-glucose, ISNH4C1, ['SN]-leucine and ['SN]-va1ine were purchased from Cambridge Isotopic Laboratories. Unlabeled amino acids and other compounds were purchased from Sigma (St. Louis, Missouri). Human CRABPII gene was over-expressed and the unlabeled protein was purified as detailed in Chapter 2. The same procedure was used for isotopic labeling of CRABPII except for some minor modifications as described below. M9 medium was used for the uniform isotopic labeling of CRABPII with 15NH4C1 and ['3C] Dé-glucose as the sole nitrogen and carbon sources for 15N- and 13C~-labeling, respectively. For 15N--labeling of leucine and valine residues, a rich medium containing [”NJ-valine or ['SN]-leucine was used (Muchmore et al., 1989). The E. coli strain DL49PS pLysS (kindly provided by Dr. David M. LeMaster) was used for the selective isotopic labeling. D20 samples were prepared by dissolving lyophilized unlabeled CRABPII in PBS buffer (20 mM sodium phosphate, 150 mM sodium chloride in 99.6% D20, pD 7.5, uncorrected). H20 samples were prepared by directly concentrating CRABPII fractions eluted with the PBS buffer (pH 7.3) from a gel-filtration column at the final step of the purification procedure (Chapter 2). About 10% D20 was added to the concentrated protein solutions. The protein concentrations of the NMR samples were ~2 mM. NMR spectroscopy. All NMR spectra were acquired at 25°C. Typical carrier frequencies for various experiments were as follows: lH, 4.70 ppm; l5N, 119.7 ppm; ”CO, 175 ppm; 13C“, 54 ppm and 13CW", 43 ppm. Quadrature detection in the 98 nonacquisition dimensions was achieved by the hypercomplex method (States et al., 1982) for homonuclear experiments and by the States-TPPI method (Marion et al., 1989a) for heteronuclear experiments. Homonuclear 2D spectra were recorded on a Varian IN OVA-600 spectrometer. The acquisition times and numbers of complex points were as follows: DQF-COSY (Piantini et al., 1982) F 1 61.0 ms, 512 and F2 243.8 ms, 2048 (128 transients); NOESY (Jeener et al., 1978; Macura & Ernst, 1980; Kumar et al., 1980) F1 38.1 ms, 320 and F2 243.8 ms, 2048 (64 transients); SSNOESY (Smallcombe, 1993) and a WATERGATE (Piotto et al., 1992) NOESY, Fl 61.0 ms, 512; F2 243.8 ms, 2048 (64 transients and 150 ms mixing time); clean-TOCSY (Griesinger et al., 1988) F1 38.1 ms, 320 and F2 243.8 ms, 2048 (64 transients). Two TOCSY spectra were acquired, one with a mixing time of 30 ms and one of 50 ms. Two NOESY spectra were recorded, one with a mixing time of 100 ms and one of 150 ms. The data were processed with the program VNMR v. 5.2F (V arian Associates). A gaussian and minus LB combination was applied in F2 dimension and a ~75°~shified sine-bell in F1 dimension in processing NOESY and TOCSY spectra. A ~30°-shificd sine-bell was applied in both F1 and F2 dimensions in processing of DQF-COSY data. The data sets were zero-filled to 4096 x 2048 real points for TOCSY and NOESY and 8192 x 4096 real points for DQF-COSY. In general, a five-order polynomial was applied for baseline correction in F2 dimension after Fourier transformation. Heteronuclear double resonance NMR spectra of uniformly ”N labeled protein were recorded with the following acquisition times and complex data points: 2D lH-‘sN HSQC (Bodenhausen & Ruben, 1980; Kay et al., 1992; Zhang etal., 1994), 'H (F2) 121.9 99 ms, 1024, ”N (F1) 53.3 ms, 128 (32 transients); 3D lH-”N NOESY-HSQC (Zhang et al., 1994), 1H (F3) 141.5 ms, 1024, ”N (F2) 34.0 ms, 68, 1H (F1) 33.2 ms, 200 (16 transients, 150 ms mixing time). 2D lH-”N HMQC spectra (Miieller, 1979; Bax et al., 1983) of the selectively labeled proteins ([”N]-valine or [”N]-leucine) were recorded with the following acquisition times and complex data points: lH (F2) 141.5 ms, 1024 and (F1) ”N 32.0 ms, 64 (32 transients). Heteronuclear triple resonance NMR spectra (Kay et al., 1990; Grzesiek & Bax, 1992a; Muhandiram & Kay, 1994; Kay, 1995) of ”N/”C-doubly labeled protein were recorded with the following acquisition times and complex data points: CT-HNCO (Grzesiek & Bax, 1992a), ”N (F1) 16.0 ms, 32, ”CO (F 2) 34.0 ms, 64, lH (F3) 127.80 ms, 1024 (16 transients); CT-HNCA and CT-HN(CO)CA (Grzesiek & Bax, 1992a), ”N (Fl) 16.0 ms, 32, l3CCl (F2) 10.9 ms, 64, 1H (F3) 127.80 ms, 1024 (16 transients); HNCACB (Wittekind & Miieller, 1993; Muhandiram & Kay, 1994), ”0"" (Fl) 7.1 ms, 64, 15N (F2) 15.2 ms, 32, 1H (F3) 128.0 ms, 1024 (32 transients); CBCA(CO)NH (Grzesiek & Bax, 1992b; Muhandiram & Kay, 1994), ”CM3 (Fl) 6.4 ms, 64, ”N (F2) 16.0 ms, 32, lH (F 3) 127.80 ms, 1024 (32 transients); C(CO)NH(G1'zesiek et al., 1993; Muhandiram & Kay, 1994), 13c (Fl) 9.7 ms, 96, ”N (F2) 15.2 ms, 32, ‘H (F3) 128.0 ms, 1024 (32 transients); H(CCO)NH (Grzesiek et al., 1993; Muhandiram & Kay, 1994), IH (Fl) 15.9 ms, 128, l5N (F2) 25.0 ms, 50, 1H (F3) 127.80 ms, 1024 (32 transients); gradient HCCH-TOCSY (Bax etal., 1991; Kay et al., 1993), 1H (F1) 35.8 ms, 128, 13C (F2), 10.5 ms, 75, 1H (F3) 127.80 ms, 1024 (16 transients, 16 ms mixing time). A low power GARP-I sequence (Shaka et al., 1985) was applied for ”N broad-band decoupling during the data acquisition. 'H broadband decoupling in the CT-HNCO, CT- HNCA, CT-HN(CO)CA, C(CO)NH and H(CCO)NH experiments were achieved by a 100 DIPSI-2 sequence (Shaka et al., 1988), and in the CBCA(CO)NH experiment by a WALTZ-l6 sequence (Shaka et al., 1983). The CO decoupling during the CM3 chemical shift and Cam-”N scalar coupling evolution periods in the HNCACB experiment, and Ccl decoupling during the constant time t2 (”N) in the CBCA(CO)NH experiment were accomplished by applying a SEDUCE-l sequence (McCoy & Mueller, 1992a, b). ”C isotropic mixings in the 3D HCCH-TOCSY (Bax et al., 1991; Kay et al., 1993), H(CCO)NH and C(CO)NH experiments were accomplished by a DIPSI-3 sequence (Shaka etal., 1988). The HSQC, HMQC, HNCACB and C(CO)NH spectra were acquired on the INOVA-600 spectrometer and processed with the program VNMR v. 5.3 (Varian Associates). All other spectra were recorded on a Briiker DMX-SOO spectrometer at the National Magnetic Resonance Facility at Madison and processed by Felix 95 (Biosym Technologies inc.). A ~45°-shified sine-bell was applied in both dimensions of the 2D HSQC and HMQC data. The HSQC data was zero-filled to 2048 x 512 real points and the HMQC data to 2048 x 256 real points. A ~75°-shified sine-bell was applied in both F1 and F3 dimensions, and a cosine-bell in the F2 dimension of the 3D NOESY-HSQC and HCCH-TOCSY data. For the other 3D data, a 75°-shified sine-bell was applied in the F3 dimension and a cosine-bell in the other two dimensions. The first points of both F1 and F2 dimensions were calculated by linear predictions. The data size in indirect detection dimension was extended by backward-forward linear predication (Zhu & Bax, 1992). Solvent suppression was improved by convolution of time domain data (Marion et al., 1989b). The data size was doubled in each dimension by zero-filling once. The empty regions of the 3D data (the aliphatic parts in the F3 dimension of the ”N-edited data and the amide part of the HCCH-TOCSY data) were removed to reduce the data size. 101 Derivation of structural restraints. Two types of structural restraints were derived from experimental NMR data: interproton distance restraints and hydrogen bond restraints. Approximate interproton distance restraints were derived from the NOE data. NOE cross peaks between aliphatic protons were picked from the homonuclear 2D NOESY spectrum with a mixing time of 100 ms using the program VNMR v. 5.2. NOE cross peaks involving amide protons were picked from the 3D 'H-”N NOESY-HSQC with a mixing time of 150 ms using the program Felix 95, except for those of Ile9, IlelO, Ile63, Arg132, Va1133 and Tyrl34, which were picked from the 2D homonuclear NOESY data. The integrated peak volumes were converted into approximate interproton distances by normalizing them against the calibrated volumes of NOE cross-peaks between backbone protons within the identified B-sheet regions. The upper limits of the interproton distances were calibrated according to the equation Va = V, (rb/ra)6, here Va, V, were the volumes and rd, rb the internuclear distances. The distance bounds were set to 1.8—2.7 A (1.8—2.9 A for NOE cross peaks involving amide protons), 1.8—3.3 A (1.8—3.5 A for NOE cross peaks involving amide protons) and 1.8—5.0 A corresponding to strong, medium and weak NOEs, respectively. Pseudoatom corrections were made for non- stereospecifically assigned methylene and methyl resonances (Wfithrich et al., 1983). An additional 0.5 A was added to the upper bounds for methyl protons. Protein backbone hydrogen bonds were derived from patterns of cross peaks remaining in the fingerprint regions of the homonuclear 2D TOCSY and NOESY spectra recorded in D20. Hydrogen bond restraints were set to 2.0 i 0.5 A and 3.0 i 0.3 A for HNmO and N---O, respectively. 102 Structure calculation. The structures were calculated by the hybrid distance geometry-simulated annealing (DGSA) protocol (Nilges et al., 1988) in the program X- PLOR v. 3.1 (Briinger, 1992) on an SGI IndigoII workstation. A square-well potential function with a force constant of 50 kca1.mol“.A’2 was applied for the distance restraints. The X-PLOR f function was used to simulate van der Waals interaction with atomic repel radii set to 0.80 times their CHARMM values (Brooks et al., 1983). Hydrogen bond restraints within the regions of regular secondary structures were introduced at a later stage of structural refinement. Seventy structures were generated using this protocol at the beginning. The structures were inspected with the programs QUANTA96 (Molecular Simulations. inc.) and InsightII (Biosym Technologies inc.), and analyzed by PROCHECK—NMR (version 3.4.4) (Laskowski et al., 1993; 1997). An iterative strategy was used for the structural refinement. The first round calculation employed mainly NOEs between backbone protons. Then the newly computed NMR structures were employed to assign more NOE restraints, to correct wrong assignments and to loosen NOE distance bounds if spectral overlapping was deduced. Then another round of structural refinement was carried out with the modified NMR restraints. Afier several rounds of such refinement, an ensemble of 25 structures was selected according to their best-fits to the experimental NMR restraints and the low values of their total energies. 103 Results Sequential resonance assignment. HNCACB and CBCA(CO)NH spectra were first analyzed to obtain sequential connectivities. Representative results are shown in Figure 3.1. About one hundred residues with relatively strong peaks in the 2D lH-”N HSQC (Figure 3.2) spectrum could be easily linked on the basis of these two sets of data. These residues were then sequentially assigned by the combined analyses of the 3D C(CO)NH, H(CCO)NH, HCCH-TOCSY, and homonuclear 2D TOCSY spectra. The majority of the remaining residues had weak or no detectable peaks in the 2D 'H-”N HSQC spectrum, even with minimum water perturbation using the water flip-back technique (Grzesiek & Bax, 1993). Because these residues also had weak or no detectable peaks in the triple resonance experiments, their sequential assignment posed a big challenge. The following residues had weak cross-peaks in the 2D lH-”N HSQC spectrum: LysS—Glu17, Leu28, Lys30, Ala32—A1a35, Ala40, Va141, Glu62, Ile63, Leu119, Leu121 and Thrl31—Va1135. They were assigned by spin system identifications combined with sequential and long range (inter-strand) NOEs. Because of the excellent chemical shift dispersion of the protein, 2D homonuclear TOCSY and NOESY data were especially useful for assigning these residues. In particular, most of the inter-strand NOE cross-peaks between Ha-H“ protons of the adjacent B-strands were well resolved in the 2D NOESY spectra. The sequential assignments of many of these residues were also facilitated by the comparisons with the sequential assignments of the wild type holo- CRABPII in complex with all-trans-RA (Chapter 5) and the apo-form of the site-directed 104 Figure 3.1. Strips of the HNCACB (a) and CBCA(CO)NH (b) spectra showing sequential connectivities for residues Gly68—Gln74. Both inter-residue and intra-residue peaks in the HNCACB strips (a) are labeled. The dotted lines indicate negative peaks. 105 —25.o .3 ' A B l3__ P 4 —3o.o é}?— [xi-1) -—35.0 B —4o.o 9 ..,,_._o $__. ““5” —45.o —so.o D a 0 $ ., a(i-1) 3:3 .. a D . 4 _ a a @__o éd 55.0 5. b o ”G —60.0 so ' © C018 3 b a a a b a a (ppm) G68 E69 E70 F71 E72 E73 Q74 Residue 106 Figure 3.2. Gradient- and sensitivity-enhanced 2D lH—”N HSQC spectrum of unifomly ”N-enriched apo-CRABPII. Sequential assignments are indicated with one-letter amino acid code and residue number. The side-chain amides of Asn and Gln are indicated by #. The unlabeled peaks are most probably from the side-chains of arginines. F1 (ppm)? 110 112 114 116 118 120 122 124 V 126 128 130 132 134 136’ 1.. .A.__L_L_...._._._._..s_... - _ 11.0 . I . .- l . . 10.5 0 107 678°.623 0' 30451: 0451 9 G117 6102‘) Gs G104 o T110 0 N91o_l:l4#N14# N141! 1.125410?“ 0 D77 N251: ° ° S10132 V135 T75 0.0034 N2#I=N91# 8891.7”154; ~091#N115#N64# N64#°337 09711 a T124 011.6 0 037,, 812'? T114 90V33 0W1099 .°K30 ’ 357%” 1 K101 4 L22 6470 °r1o7' E16E103 (85389 11}, 0mg F15 87K?” 0'23 V94 $572 ~64" M27 T1220 K92 L34 131 °°V26 Na R11 K660 R111W87oo127 £69 “79 $01134 a M1230 OF71 R136 @7135 V21 097 F3 D126QV76 E623V133°°Y513L100L18N24 W7 I43 F50 15112:3 128/13826 0 W7N€ {Aw/E13 “T8231 9 265%“ A125 sV41 ”20 19 o V129 M93 L99 9 0 flE137 p 0 348 L113 0 Y134 95'& O a 523%.”: 045 E96 R132 K44 o E42 E46 0130 fives E70 0 K53 61.121 ‘1 """ 1i? ’l' 'l’l'FWT'll 0 9.5 9.0 8.5 8.0 7.5 7.0 6.5 108 mutant R1 1 1M (Chapter 4). Most weak peaks in the 2D lH-”N HSQC spectrum of wild type apo-CRABPII became much stronger upon the binding of RA or with the point mutation, and thus could be assigned unambiguously for holo-CRABPII and apo-R11 1M. Furthermore, neither the binding of RA nor the mutation changed the chemical shifts of the majority of residues. Such comparisons were especially helpful for the assignments of the selectively lSN-labeled residues such as Leu119, Va133, Val41, Val76, Va1133 and Leu121. Such comparisons also expedited greatly the assignments of the forty-six residues with amide protons remaining in the fingerprint regions of the homonuclear 2D spectra in D20 (Figure 3.3). Almost exactly the same residues remained in the 2D homonuclear D20 spectra of apo-CRABPII and apo-R1 11M (Figure 4.8 of Chapter 4). In particular, such comparisons helped the assignments of residues LysS, Ile9, IlelO, Ser12, Ile63, Thr131, Arg132 and Tyrl34. Nine residues, namely Glu13, Asn14, PhelS, Lys30, Ile3l, A1340, Asp77, Gly78 and Cysl30, were assigned based on NOEs and with the help of the comparisons. The HN and ”N resonance assignments of Va133 were based on selective labeling, spin system identification and several NOEs between the ring H53 of Phe15 and HS of Va133 after the first round of structural calculation. The HN and ”N resonances of Ala32, Ala34 and Ala35 were assigned based on the comparisons and a few weak, sequential NOEs. The HN or ”N resonance of Ser37 was assigned tentatively by spin system identification and the comparisons of the three sets of 2D lH-”N HSQC spectra. The peaks corresponding to the two adjacent residues, Ala36 and Lys38, could not be observed in the 2D lH-”N HSQC spectrum of apo-CRABPII. The aliphatic protons of the residues with weak peaks in the 2D lH-”N HSQC spectrum were assigned by the inter-strand NOEs, by spin system identifications obtained from the homonuclear 109 2D spectra and 3D HCCH-TOCSY spectrum, and by the comparisons with the assignments of holo-CRABPII and apo-R1 11M. For examples, the spin systems of Ser12, A1a40, Val4l, Thr54, Glu62 and Val76 could be easily identified. The H“ of Arg132 was assigned by its inter-strand NOE to Ser12 observed in the homonuclear 2D NOESY spectrum. The following eleven residues were not observed in the 2D lH-”N HSQC spectrum: Asn2, Arg29, Ala36, Lys38, Thr56—Thr61 and Asn115. They were first grouped into several spin systems based on the homonuclear 2D DQF-COSY, TOCSY and 3D HCCH-TOCSY experiments. The aliphatic protons of Asn2, Thr56, Thr57, Va158, Thr60, Thr61 and Asn115 were assigned by spin system identifications, the inter- strand NOEs, and by the comparisons with the assignments of holo-CRABPII and apo- Rl 11M. The assignments of the aliphatic protons of Thr56 and Thr57 were further helped by the similar but unique chemical shifts of their H“ and HB resonances in the spectra of apo- and holo-CRABPII and apo-R1 11M: H“ 4.83 ppm and HB 4.61 ppm for Thr56, and H“ 4.00 ppm and H“ 4.23 ppm for Thr57 (in apo-R1 1 1M). Two possible alanine peaks from Ala32 and Ala36 could be identified in the 2D homonuclear and 3D HCCH-TOCSY spectra, but their sequential assignments could not be obtained due to the lack of strong, well-resolved sequential NOEs. None of the 1H, ”C and ”N resonances of Arg29, Lys38, Pro39 and Arg59 were assigned (Table 3.1). The ring protons of aromatic residues, except for PhelS and Tyrl34, could be easily assigned by analysis of the 2D DQFCOSY and TOCSY data, and by the comparisons with the assignments of holo-CRABPII and apo-R11 1M (Chapters 4, 5). Their identities were confirmed later by sequential assignment and the NOEs between H”- 110 Figure 3.3. The fingerprint region of the homonuclear 2D TOCSY spectrum with 40 ms mixing time. The TOCSY experiment was initiated more than 24 h after dissolving a lyophilized protein sample in PBS, pD 7 .5, in D20. lll __.._ _. _-____._ _..___ _.___ 7 0 F1 '; T122(HB) ' (pm); 4.4 V94 K86 l s 635 ° 4-6‘3 s 0 F50 ”2‘0ch K44 "0 .0573 4'8 3 0 . A125 ; I120 s .63 5.0‘j . K5 M123 0.8 V133 T: R132 W7 . 5.2; 7110 , F71 L119 Y51 K08 C130 '43 1.113 5.4 :f V135/E 2 L121 4 , K. 5.6"}:9 ° 1 O 92, Y134 0 111 0 5'8: T122 51°12 W109 "7‘71'l'TT'iNUTT'T‘TTTT‘ITE' " l! TIT!) ll lljlinllquTqfi WITH W11 TITO T1 TT'lTiTTTl'lTTTlUTTTPlTTTlTTTITTTTTTqu 10.2 9.8 9.4 9.0 8.6 8.2 7.8 F2 (PM) 112 Table 3.1. 'H, ”N and ”C assignments of apo-CRABPII in PBS at 25°C. Residue P 1 N2 F3 54 GS N6 W7 K8 19 110 R11 512 E13 N14 F15 E16 E17 L18 L19 K20 V21 L22 023 V24 N25 V26 M27 P39 T49 F50 Y51 152 K53 T54 S55 T56 T57 V58 ISN 121.9 (9.82) 112.5 (8.49) 110.6 (9.22) 118.6 (8.15) 123.3 (9.35) 123.2 (l0.l7) 125.6 (9.05) 118.8 (9.24) 120.8 (7.57) 115.8 (8.33) 124.0 (8.71) 117.8 (9.25) 119.3 (8.72) 117.4 (8.36) 118.5 (8.36) 121.4 (7.59) 114.5 (7.36) 118.9 (8.18) 121.3 (7.83) 117.5 (7.24) 106.5 (7.69) 121.4 (7.62) 127.9 (8.68) 120.3 (8.47) 119.7 (7.95) 118.6 (7.82) 116.5 (7.66) 119.2 (7.16) 122.6 (8.54) 116.7 (8.66) 121.0 (7.31) 121.6 (8.44) 113.9 (7.43) 123.9 (8.71) 125.2 (7.93) 130.0 (9.30) 123.1 (9.30) 128.9 (9.56) 128.0 (8.58) 128.9 (8.53) 1 17.9 (9.10) 126.6 (8.54) 116.7 (8.08) 124.2 (9.08) 122.5 (8.35) 123.3 (8.41) 133.1 (9.30) 124.4 (8.85) 122.7 (9.43) 124.0 (8.71 ) CO 171.9 173.5 171.8 167.9 171.5 172.5 172.4 173.3 173.3 171.0 172.1 175.9 174.9 176.0 178.3 174.9 173.6 172.0 173.6 172.8 173.4 177.7 174.8 174.4 171.5 172.1 173.9 170.8 172.2 173.5 170.6 172.5 170.1 173.1 172.8 171.6 172.0 173.5 C0. 63.0 (4.32) 52.7 (5.03) 60.4 (4.32) 60.7 (4.70) 44.9 (3.96, 3.76) 53.0 (5.52) 56.7 (5.14) 54.1 (5.28) 55.5 (3.92) 61.3 (4.61) 62.8 (5.12) 61.7 (4.80) 55.7 (4.95) 55.3 (4.59) 63.2 (3.30) 59.6 (3.68) 58.5 (3.73) 57.6 (3.62) 60.5 (3.83) 66.0 (3.78) 55.4 (3.91) 45.9 (3.96, 3.58) 63.3 (3.72) 54.2 (4.38) 66.7 (3.46) 58.5 (4.14) (4.08) (4.32) 59.6 (3.87) 64.3 (3.72) 66.6 (3.62) (4.08) (4.21) 52.2 60.4 (4.70) 51.9 (4.96) 61.2 (5.13) 54.6 (5.36) 60.0 (5.34) 56.4 (4.60) 53.9 (4.45) 55.1 (4.37) 48.0 (3.94, 3.58) 54.5 (5.18) 63.0 (4.70) 57.6 (4.56) 56.3 (5.24) 61.0 (4.60) 54.5 (5.08) 62.1 (4.88) 58.0 (5.30) (4.85) 64.9 (4.00) 62.2 (4.16) (:13 33.2 (2.37, 1.91) 39.3 (2.95, 2.89) 32.3 (3.28, 2.70) 64.5 (4.30, 3.99) 40.2 (2.69, 2.52) 32.2 (3.06, 2.66) 36.1 (1.90, 1.85) 33.6 (1.76) 41 .8 (1.83) (3.61, 3.43) 40.1 (2.19, 2.08) 41.4 (3.31, 3.22) 33.7 (2.98, 2.85) 32.6 (1.96) 28.1 (1.84, 1.73) 42.4 (2.17, 1.34) 41.1 (1.62, 1.56) 32.6 (1.93, 1.61) 31.7 (2.15) 41.5 (1.39) 31.7 (1.45) 39.3 (3.09, 2.74) 32.1 (2.03) 32.0 (2.07, 2.01) (1.63, 1.32) (2.20, 1.97) 32.2(189, 1.81) 39.2 (1.80) (2.07) (1.38) (1.38) 19.1 (4.23, 4.02) 20.9 (1.45) 35.3 (1.73) 33.4 (2.06, 1.99) 40.8 (2.27) 35.3 (1.87, 1.76) 30.6 (1.82, 1.71) 31.0 (1.62, 1.26) 42.1 (3.15, 2.68) 69.6 (4.14) 44.6 (1.65) 41.0 (3.01, 2.70) 40.4 (1.84) 34.9 (1.76, 1.58) 70.0 (3.86) 65.0 (3.83, 3.74) (4.61) 66.7 (4.23) (1.96) Others H" 2.10, 2.08; H“ 3.32, 3.25 2, 6H 7.42; 3, 5H 7.16; 4H 7.34 2H 6.78; 4H 7.07; 5H 6.82; 6H 7.23; 7H 7.12 H“ 9.20; N‘ 124.6 H" 1.63, 1.52; H“ 1.37, 1.33; H‘2.85 H" 1.65, 1.56; H“ 0.48; H""' 1.39 H""‘0.87 H’l.93 H“ 7.35, 7.00; N“ 109.8 2, 6H 7.14; 3, 5H 6.84; 4H 6.93 H" 2.16 H" 2.38, 2.19 H" 0.81; H“ 0.74, 0.70 H" 1.12; H“ 0.36, 0.32 H" 1.35, 1.14; H“ 2.91 H" 1.10, 0.95 H“ 0.49 H" 0.83 H“ 7.73, 6.91; N“ 112.3 H" 0.96, 0.91 H" 2.59, 2.46 H" 1.09; H“ 0.82, 0.76 H" 1.74; H“ 2.70 H" 1.68, 1.59 H""‘ 0.88; H“ 0.80 H" 0.94 H" 0.63, 0.46 H" 2.25 H1 1.66; H"‘“ 0.85; H“ 0.51 H" 1.54, 1.41; H“ 1.33, 1.24 H" 2.03; H“ 7.83, 7.11; Nc 106.8 H" 2.02 H" 0.99 2, 6H 6.48; 3, 5H 7.08; 4H 6.81 2, 6H 6.94; 3, 5H 6.61 H’ 1.26; H""' 0.80; H“ 0.31 H" 1.17; H“ 0.44, 0.29; H“ 2.84 H" 0.94 H" 1.07 H" 1.27 H" 0.79, 0.74 113 Residue R59 T60 T61 E62 163 N64 F65 K66 V67 G68 E69 E70 F7 1 E72 E73 Q74 T75 V76 D77 G78 R79 P80 C8 1 K82 583 L84 V85 K86 W87 E88 S89 E90 N91 K92 M93 V94 C95 E96 Q97 K98 L99 L100 K101 6102 E103 G104 P105 K106 T107 8108 W109 T110 R111 5112 L113 T114 N115 D116 Gll7 E118 ISN 123.5 (8.58) 122.9 (8.73) 120.4 (8.10) 117.2 (8.34) 120.8 (9.32) 127.5 (9.01) 114.9 (8.91) 120.2 (7.94) 130.3 (9.21) 121.4 (9.06) 119.6 (8.67) 122.4 (8.44) 117.9 (7.59) 111.8 (9.18) 122.4 (9.18) 112.0 (8.54) 107.2 (7.93) 121.0 (7.86) 118.8 (8.87) 120.0 (8.87) 124.9 (9.06) 120.7 (7.36) 130.1 (9.20) 123.8 (8.47) 120.2 (8.65) 127.7 (9.21) 112.5 (8.59) 116.7 (8.71) 111.2 (8.05) 120.1 (7.84) 127.0 (9.35) 119.3 (8.86) 127.3 (8.85) 127.7 (7.70) 121.0 (7.43) 123.1 (8.94) 126.1 (8.78) 121.9 (7.77) 117.4 (8.05) 108.6 (8.18) 117.2 (8.10) 109.1 (8.14) 123.5 (8.58) 118.7 (8.79) 112.3 (7.80) 116.3 (8.21) 111.2 (9.56) 120.1 (8.79) 123.3 (8.87) 126.4 (8.25) 116.9 (9.35) 115.7 (7.75) 107.6 (7.96) 117.5 (7.43) CO 170.5 169.7 169.1 175.2 174.1 170.7 171.5 172.6 169.5 173.1 171.6 174.3 175.4 171.4 173.6 168.5 173.0 168.7 173.7 171.5 173.3 174.9 173.3 171.2 173.1 171.3 171.0 171.0 170.0 171.4 172.6 173.1 172.9 175.4 174.8 172.7 169.7 173.1 172.8 175.1 170.8 171.2 171.8 171.7 170.9 170.5 175.1 173.1 173.5 171.7 172.0 CG 60.3 (5.20) 60.1 (4.61) (4.98) 59.5 (4.88) 52.3 (5.56) 55.9 (4.65) 53.9 (4.65) 67.1 (3.23) 45.7 (4.38, 3.61) 54.9 (4.77) 57.6 (4.88) 55.8 (5.11) 55.3 (5.05) 55.3 (4.73) 54.2 (5.30) 62.4 (4.55) 65.6 (3.82) 45.9 (4,10, 3.55) 54.0 63.8 (4.66) 55.9 (5.05) 54.5 (4.73) 58.7 (4.86) 54.9 (4.23) 60.6 (4.35) 54.2 (4.48) 57.0 (5.1 1) 58.0 (4.35) 57.0 (4.55) 59.1 (3.66) 52.9 (4.96) 56.7 (5.46) 53.4 (5.18) 60.3 (4.30) 55.6 (4.48) 55.2 (4.31) 54.7 (4.91) 54.7 (4.60) 56.0 (4.00) 57.0 (4.07) 55.4 (4.37) 45.1 (4.02, 3.71) 54.9 (4.60) 45.5 (4.10, 3.90) 62.7 (4.48) 56.6 (4.38) 60.1 (5.36) 57.3 (4.63) 56.1 (5.76) 61.0 (5.24) 56.1 (5.57) 53.9 (5.78) 52.8 (5.28) (4.76) 55.5 (4.44) 54.3 (4.74) 46.1 (4.17, 3.69) (4.98) (:13 71.7 (3.90) 71.7 (3.97) (2.02, 1.91) 42.2 (1.85 ) 42.8 (2.54, 2.50) 39.8 (2.15) 35.4 (1.63, 1.49) 31.9 (1.85) 35.3 (2.16, 1.96) 31.4(1.99,1.91) 41.4 (3.53, 3.20) 31.9 (202,191) 29.7 (202,190) 33.1 (1.73, 1.70) 71.2 (4.16) 32.1(200) 29. 8 33.0 (2.38, 1.86) 31.6 (2.58) 34.3 (1.54, 1.47) 68.0 (2.78, 1.54) 45.2 (2.00, 1.58) 32.6 (1.35) 36.9(1.6l, 1.43) 29.5 (3.22, 2.99) 32.3 (1.80, 1.66) 65.0 (4.1 1, 3.98) 29.5 (1.84, 1.78) 40.4 (3.15, 2.85) 37.2 (1.98, 1.47) 35.1 (1.99) 34.9 (1.77) 27.9 (0.36, -0.28) 31.6(1.81,1.73) 30.8 (1.75) 35.0 (2.10) 42.0 (1.98, 1.63) 42.5 (1.64, 1.47) 34.9 (1.72, 1.61) 32.9 (1.96, 1.78) 32.7 (2.03, 1.33) 32.6(1.81, 1.51) 72.5 (4.46) 66.8 (3.86, 3.78) 33.1 (3.48, 3.29) 72.2 (4.16) 35.6(185. 1.50) 34.9 (1.94) 43.9 (1.25) (4.70) 38.1 (2.79, 2.76) 41.2 (2.79, 2.39) (2.02, 1.98) Others H" 1.06 H" 1.06 1171.16, 101;va 1.09;H“0.80 H“ 7.62, 6.92; N“ 113.1 2, 6H 6.24; 3, 5H 6.52; 4H 6.32 117 1.45, 1.28; 115 2.92 H, 0.64, 0.43 H" 2.19 H" 2.20, 2.13 2, 6H 7.32; 3, 5H 6.82; 4H 6.52 H" 2.30, 2.13 H" 2.14, 2.14 H" 2.17, 1.99; H‘ 7.49, 6.70; N" 110.3 H" 1.03 H" 1.01, 0.81 H" 2.15, 1.99 H" 1.26, 1.22; H“ 1.03 H" 1.17; H“ 0.68, 0.51 H" 0.23, 0.02 H" 1.26, 1.12 2H 7.30; 4H 7.98; 5H 6.95; 6H 7.21; 7H 7.62 H‘ 10.66; N“ 129.6 H" 1.99 H" 2.08 H“ 7.50, 6.83; N“ 112.6 H" 1.22; H“ 1.10; H8 2.68 H" 2.19, 2.08; H“ 0.41 H" 0.73, 0.70 H" 2.14, 1.95 H" 2.43, 2.02; H“ 7.38, 6.71; Nc 114.4 H" 1.96, 1.78; H“ 1.24; H‘ 3.16, 2.84 H" 1.36; H“ 0.86, 0.65 H" 1.35; H“ 0.80, 0.76 H" 1.37, 1.27 H"2.10 H" 1.84, 1.76; H“ 3.60, 3.29 H" 1.35, 1.30; H" 1.13 2H 6.89; 4H 6.90; 5H 6.35; 6H 6.43; 7H 6.59 H“ 11.37; N‘ 137.8 H" 1.09 H" 1.35, 1.23; H“ 3.00 H" 2.03 H" 0.94; H“ 0.51, 0.44 1.25 H 7.67, 6.95; N“ 112.6 114 Residue L119 1120 L121 T122 M123 T124 A125 D126 D127 V128 V129 C130 T131 R132 V133 Y134 V135 R136 E137 ISN 124.6 (8.73) 125.2 (9.47) 135.5 (9.88) 120.1 (9.13) 121.7 (9.29) 115.2 (8.26) 125.3 (8.39) 122.4 (9.28) 121.1 (8.48) 123.4 (8.41) 125.7 (8.38) 129.4 (9.56) 124.6(8.11) 128.3 (9.77) 122.2 (8.50) 127.2 (9.79) 112.1 (9.55) 123.2 (8.55) 125.6 (8.08) CO 172.6 172.8 171.4 171.5 172.1 172.7 172.7 172.8 171.2 172.9 171.2 172.5 172.8 174.5 178.7 CG 53.4 (5.08) 60.8 (4.83) 53.3 (5.56) 60.4 (5.76) 55.1 (5.14) 60.7 (5.46) 51.6 (4.84) 57.3 (4.14) 54.7 (4.50) 63.6 (3.95) 61.4 (4.50) 56.4 (5.16) (5.13) (5.01) 62.2 (4.90) 56.5 (5.59) 59.7 (5.33) 57.3 (3.89) 58.9 (3.86) (:13 43.3 (0.89) 39.4 ( 1.77) 45.3 (2.15) 71.0 (4.03) 38.3 (1.84, 1.81) 71.2 (3.93) 21.3 (1.03) 39.6 (2.72) 41.4 (2.66) 32.9 (2.14) 34.4 (1.83) 35.4 (3.00, 2.66) (3.74) (1.78, 1.67) (1.66) 43.6 (3.23, 3.01) 36.2 (2.36) 30.3(l.29,1.11) 31.6 (1.80, 1.68) Others H" 0.44; H“ 0.18, —0.36 H" 1.35; H"“’0.75; H“ 0.99 H" 1.22; H“ 0.85, 0.74 H" 1.06 H" 2.55, 2.28 H" 0.96 H" 0.92, 0.69 H" 0.77, 0.69 H" 0.98 H" 1.50 H" 0.85, 0.80 2, 6H 6.82; 3, 5H 6.37 H" 1.04, 0.85 H" 0.59, 0.06; H“ 2.70, 2.79 H" 2.05, 1.90 115 H“ and between Ha-H“ for tyrosines and phenylalanines, between HB-H“l and between H“- H“3 for tryptophans. The ring protons of PhelS and Tyrl34, which were overlapped with other peaks in the aromatic regions, were assigned based on the NOEs between HB-H“ and between Ha-H“ and the comparisons. In the corresponding spectra of holo-CRABPII and apo-R1 1 1M they were separated from the other peaks. Stereospecific resonance assignment. In order to better determine the solution structure of CRABPII, stereospecific assignments were made whenever possible. The methylene H‘3 protons were stereospecifically assigned on the basis of the qualitative estimations of 33613 coupling constants from DQF-COSY spectrum in conjunction with the relative volumes of the NOE cross-peaks between H"‘—Hl3 picked from the 2D homonuclear NOESY with 100 ms mixing time and the NOE cross-peaks between HN—H“ from 3D lH-‘SN NOESY-HSQC (Basus, 1984). The methyl protons of valine residues were stereospecifically assigned in a similar manner except that the relative volumes of the NOE cross-peaks of H“—H" and HN—H" were used. Methyl protons of leucine residues and the ring protons of Phe and Tyr residues were stereospecifically assigned after two rounds of structure refinement. Stereospecific assignments were obtained for H'3 methylene protons of 43 residues, for methyl protons of six valine residues and four leucine residues. Secondary structures. Secondary structural elements were obtained from NOE data. The NOEs used to characterize secondary structures were d,m and dag from the homonuclear 2D NOESY and dam d- and dim from the 3D 'H-‘SN NOESY-HSQC and/or homonuclear 2D NOESY data. Regular a-helix was characterized by strong short-range NOEs dNN, do“, (i, i+3) and de (i, i+3) and weak dew, and regular B-sheet by strong short- 116 range NOEs dew and de and strong or medium long-range NOEs do“, and do,” between adjacent strands. The following ten B-strands forming two anti-parallel B-sheets were identified (Figure 3.4): Ser4—SerlZ (BA), Ala40—Glu46 (BB), Thr49—Ser55 (BC), Thr61— Lys66 (BD), G1u69—Thr75 (BE), ProSO—G1u88 (BF), Lys92—Lys98 (BG), K106—L113 (BH), L119—A125 (BI) and V128—R136 (BJ). Long-range NOEs, a",m and daN, were observed between adjacent BA, BB, BC and BD, between adjacent BE, BF, BG, BH, BI and BJ, and also between the N-terminal half of BA and the C-terminal half of BJ. No such NOEs were observed between the adjacent BD and BE. LysS-Ser12 of BA, LysS3 of BC, Ile63 of BD and Thr131—Vall35 of BJ had weak peaks in the 2D lH-‘SN HSQC spectrum. However, the amide protons of Lys8, Ile10, LysS3, Ile63 and Arg132—Va1135 were observed in the homonuclear 2D spectra recorded in D20 (Figure 3.3). In particular, most of their inter-strand NOE distances due, and do,” was observed in the 2D NOESY spectra in D20. Therefore, Lys8—Ser12 and Thrl31—Va1135 were considered to adopt B-sheeted structures. They formed part of the so-called B-bulge structure. A1a40—Val4l of BB and Thr60—Glu62 of BD showed weak or no observable peaks in the 2D lH-‘SN HSQC and 3D lH-‘SN NOESY-HSQC spectra. However, strong NOEs (dad) were observed for these residues. The first two B-strands were linked by a long segment (Glul3-Pr039). Along this segment a short a-helix 01A, Glul7—Gly23, could be recognized. Glu17 was weak in the 2D lH-‘SN HSQC spectrum. It was assigned as the N-terminus of the helix according to the NOE (110,,3 (i, 1' +3) observed in the D20 spectra. 117 Figure 3.4. Summary of the sequential and medium-range NOEs involving backbone HN and H“ atoms, slow-exchange backbone amide protons and the deduced secondary structures of apo-CRABPII. Line thickness for (11,,N and a"NN sequential NOE distances reflects the intensities of the cross peaks. Asterisks indicate the residues whose amide protons remained after 24 h of exchange with D20 in PBS buffer at room temperature and pD 7.5 as determined by 2D homonuclear NMR experiments. 118 10 20 30 | 1 I PNFSGNWKIIRSENFEELLKVLGVNVMLRKIAVAAA Slow NH exchange ' " ‘ " "‘ “ ‘ ‘ “ daN(i,i+1) I — _-_ _ H .— dNN(i,i+1) — _ I - _- .— daN(i,i+2) — —"=="——-— '— dNN(i,i+2) —‘—=—— — 36140.16) —— —‘===—- NN('.°+3) Secondary/“structures Ct, ,j PA ,f I > . OLA 1 4'0 5'0 60 70 l l SKPAVEIKQEGDTFYIKTSTTVRTTEINFKVGEEFEE Slow NH exchange ‘ “ ’ ‘ ‘ ‘ ‘ ‘ “ " ‘ daN(i,i+1) — _ I - - __ dotN(i,i+2) dNN(',1+3) _ Secondar; structures CDZEIBZII» Snag” cur-£12m) 1:231:53: 1° 1° 11° QTVDGRPCKSLVKWESENKMVCEQKLLKGEGPKT Slow NH exchange ‘ "' ‘ " “ “ ‘ d01N(i,i+l) - — _——-—- - dNN (1,1+1) — — _ - - - fl — - daN(i,i+2) —-=-— dNN(i,i+2) — — dNN(i,1+3) BF ' " BO Secondarystructuresuzb “11.11.,fitjp 111111114> ccn 1'10 120 130 SWTRELTNDGELILTMTADDVVCTRVYVRE Slow NH exchange . t I t I O t t t I 3 t t t 3 t t i . daN(i,i+l) _ __ __- dNN (Li-F1) — - i — - — daN(1,1+2) — dNN(i,i+2) “:— G10: (1 1+3) 1 H Secondarystructures 12:21:23) Egg” mxrxlgjnxxb 119 Glul3—Glu16 had very weak peaks in the 2D 'H-‘SN HSQC, homonuclear 2D SSNOESY and WATERGATE NOESY spectra. NOE distance restraints. A total of 2290 structurally useful NOE distance restraints were obtained from the analyses of the homonuclear 2D NOESY and 3D IH-‘SN NOESY-HSQC spectra. Of the NOE restraints, 647 were intra-residue, 560 were sequential, 248 were medium-range and 835 were long-range NOEs (Table 3.2). On average, each residue had ~17 NOE restrains, but the numbers of NOE restraints per residue varied greatly along the amino acid sequence (Figure 3.5). The residues situated in several turns or loops and the entrance region had NOE restraints well below the average. A total of 92 hydrogen bond restraints obtained fiom 46 hydrogen bonds were included in the structural refinement. Quality of the structure. Figure 3. 6A shows the C“ traces of the superimposed 25 computed structures with no NOE violations exceeding 0.3 A. The statistics of the structures are summarized in Table 3.2. Except for the entrance region, the structures were well defined. Excluding residues Ala32-Pro39 and Thr57-Glu62, the RMSD of the 25 coordinates was 0.54 A for the backbone (N, C“, C’, O) and 0.92 A for all heavy atoms. The stereochemical qualities of the structures were checked by the program PROCHECK—NMR (Table 3. 2). The restrained minimized mean structure (Clore & Gronenbom, 1989) had 55.6% residues in the most favorable regions of the Ramachandran plot, and 37.1% in the additional allowed regions. Two residues, Phe3 and Asn115, and eleven others were at the disallowed and generously allowed regions, respectively. The relative low percentage of the residues in the most favorable regions 120 Figure 3.5. The distribution of NOEs along the amino acid sequence of CRABPII. For intra-residue NOEs only those that are structurally useful are included. No. of NOEs No. of NOEs 121 Long Medium Sequential lntraresidue 1 .113 ' ‘1 gt}? 1" ’ .1 I 1 “1:33.35: 11: ' ,1. 1; PNFSGNWKIIRSENFEELLKVLGVNVMLRK1AVAAASKPAVEIKOEGDTFVIKTSTTVRTTEINFKVGE 5 10 15 20 25 30 35 40 45 50 55 so es Residue Number so 40 20 .l.‘§ ' 1] 3311.3‘ig : "113'; 5" l' 1.; ' 1‘; o s...€.~.«.24:~391..<.§.u.~ 1. 1 it EFEEQTVOGRPCKSLvxweseuxuvceoxtLKGEGPKTSWtRELTNDOEL1Lruerovvcravvvns 70 75 80 85 90 95 100 105 110 115 120 125 130 135 Residue Number 122 Figure 3.6. (A) Stereoview of the C“ traces of the superimposed 25 refined structures of apo-CRABPII. (B) Stereoview of the C“ traces of the restrained minimized mean structure of apo-CRABPII (thin line) superimposed with the C“ traces of the crystal structure of holo—CRABPII (thick line). (C) Stereoview of the C“ traces of the restrained minimized mean structure of apo-CRABPII (thin line) superimposed with the C“ traces of the crystal structure of apo-CRABPI (molecule A) (thick line). 123 124 Table 3.2. Restraint and structural statistics of apo-CRABPII Restraint Statistics Numbers of experimental NOE restraints Intra-residue 647 Sequential 560 Medium 248 Long 835 Total 2290 Number of hydrogen bonds 92 Structural Statistics 3 ,b RMSD from experimental distance restraints (A) (2290) 0.036 t 0.001 0.028 Deviation from idealized covalent geometry Bonds (A) 0.004 i 0.001 0.003 Angles (deg) 0.742 t 0.028 0.583 Impropers (deg) 0.604 3: 0.006 0.482 Measures of Structural Quality (By Procheck) % residues in most favorable regions of the Ramachandran plot 55.6 i 2.4 54.8 % residues in additional allowed regions of the Ramachandran 37.1 :1: 1.6 37.1 plot No. of bad contacts 8 :1: 2 8 H-bond energy 0.40 i 0.10 0.50 Overall G-factor -0.30 :1: 0.02 -0.30 Coordinate Precision RMSD for Ca trace (A) 0.84 i 0.21 0.47 :t 0.15 ° RMSD for backbone atoms (A) 0.89 3: 0.25 0.54 i 0.18 ° RMSD for all heavy atoms (A) 1.12 :t 0.28 0.92 :1: 0.20c “ are the final 25 simulated annealing structures. “ r is the restrained minimized mean structure obtained by restrained regularization of the mean structure, which is obtained by averaging the coordinates of the individual SA structures best fitted to each other. ° Residues from Ala32-Pro39 and Thr57-Glu62 were excluded from the RMSD calculations. 125 was due to the looseness or absence of the backbone NOE restraints for more than one- fourth of the total 137 residues. Description of the solution structure of apo-CRABPII. The solution structure of apo-CRABPII mainly consists of two nearly orthogonal anti-parallel B-sheets twisted to form a flattened barrel, a structural feature shared by all the published structures of iLBPs (Banaszak et al., 1994). As in other iLBPs, owing to a large gap, there are no main-chain hydrogen bonds between BD and BE. The gap gradually widens fi'om the N-terminus to the C-terminus of BD (Figure 3. 6A). The C-terminal half of BC and the N-terminal half of BD together with the loop connecting them move away from the main protein body and extend into the solvent. The C-tenninal half of BB together with the loop connecting BE and BF also moves away the main protein body. The entrance to the binding pocket is formed by the hydrophobic side-chains of Leu19, Va124, Leu28—Pro39 and the residues in the BC—BD loop and BE—BF loop. The distance from the C“ atom of Va158 in the BC— BD loop to the C“ atom of Val76 in the BE—BF loop is about 20.7 A. The distance between the C“ atom of Va124 of the loop connecting 01A to 018 and the C“ atom of Val76 is 17.0 A. The ligand binding pocket is wide open and the side chains of the trio (Argl 1 l, Arg132 and Tyrl34) are readily accessible to the external ligand. 126 Discussion Assignment and secondary structures. The amide resonances of eleven residues were not observed, and about twice the number of residues had weak peaks in the 2D 1H- 15N HSQC spectrum. This posed a challenge for the sequential assignment method based on a series of ISN-edited heteronuclear NMR experiments because these experiments depend on magnetization transfer of amide resonances. Their assignments could not be accomplished without the help of the comparisons with the sequential assignments of both the wild type holo-CRABPII and the site-directed mutant apo-R11 1M. However, the intensities of the amide proton signals from Arg29, Ala36, Lys38 and Arg59 were so low in both the 2D lH-‘SN HSQC and the homonuclear 2D SSNOESY and WATERGATE NOESY spectra that none of their 1H and 15N resonances has been assigned. Secondary structure elements were identified according to the NOE patterns. For residues with weak or no observable amide peaks in NOESY spectra only the a",m and daB(i, i +3) NOEs were used. Several residues had strong peaks in the fingerprint regions of the 2D homonuclear spectra recorded in DZO but had only very weak peaks in the 15N-edited heteronuclear spectra. Their secondary structural identities were established by the backbone hydrogen bond, and the NOEs observed in the homonuclear spectra. All the available data, in particular, the homonuclear 2D SSNOESY, WATERGATE NOESY spectra were analyzed carefully to assign the resonances of Met27-Ala36, which assume an or-helical structure (018) according to the crystal structure of holo-CRABPII. Some weak sequential 127 NOEs typical of the regular helical structure were observed for Met27-Ile31 but their intensities and numbers were well below the average (Figure 3.5). No structurally useful sequential NOEs could be identified for residues Ala32-Ala35. However, the chemical shift indices suggested that Lys30-Ala35 are helical. No NOEs could be identified unambiguously for Ser37 except its intra-residue NOEs. These experimental data suggested that the residues Met27-Ala35 adopt an unstable or-helical structure. Residues Ala32-Pro39 are disordered in the NMR structures because no chemical shift indices were included in the structure refinement (Figure 3.6A). Comparisons with the crystal structures of halo-CRABPII and apo-CRABPI. Figure 3.6B shows the overlay of the C“ traces of the restrained minimized mean structure of the apo-CRABPII and the crystal structure of holo-CRABPII. The structure can be superimposed with C“ deviations of 3.52, 2.04, and 1.65 A for all residues, the regular secondary structures, and all residues excluding Va124-Ser37, Glu74-Pro80 and Leu100- Gly104, respectively. The main differences between the two structures are located in the regions around the ligand entrance. The second helix is well defined in the holo crystal structure but appears flexible in solution, especially the C-terminal half. The BC-BD loop moves away fiom the main body of the protein by ~5 A in the apo solution structure. The C-terminal half of BE and the N-terminal half of BF in the apo solution structure do not twist toward the center of the protein to constrict the ligand entrance as they do in the holo crystal structure. Consequently, the long loop between BG and BH also moves away from the main body of the apo protein. The distance between the C“ atom of Val76 in the BE—BF loop and the C“ atom of V3158 in the BC—BD loop is ~8.0 A longer in the apo solution structure than in the holo crystal structure. The distance between the C“ atom of 128 Val76 and the C“ atom of Va124 in the orA-orB loop is ~6.7 A longer in the apo solution structure than in the holo crystal structure. As a result, the ligand entrance in the apo solution structure is greatly enlarged with a much more exposed binding pocket. In particular, residues Va124, Leu28, Ile31, Va158, Arg59, and Val76 move away from each other so that the side-chains of Argl 1 l, Arg132 and Tyrl34 that interact with the carboxyl group of RA are easily accessible to the ligand. However, the relative positions of Argl l 1, Arg132, and Tyrl34 are quite similar to those observed in the holo crystal structure: their backbone atoms can be superimposed rather well. Compared with other parts of the protein, the ligand entrance region of apo- CRABPII is not well defined in the solution structure. Most residues in the region have weak or no observable peaks in all 'SN-edited spectra. The number of the experimental distance restraints per residue for the region is well below the average (Figure 3.5). However, except for Ala32—Pro39, the differences between the apo solution structure and the holo crystal structure (Figure 3.6B) are larger than between any two conformations from the ensemble of 25 refined NMR solution conformations (Figure 3.6A). Thus, the differences between the apo solution structure and the holo crystal structure are real. Most of the differences between the apo solution structure and the holo crystal structure of CRABPII are also observed between the solution structure of apo-CRABPII and the crystal structure of apo-CRABPI (Thompson et al., 1995) (Figure 3.6C). Dynamical properties. As detailed earlier, about a quarter of the residues showed weak or no signals in the 2D 'H-‘sN HSQC spectrum (Figure 3.2). Some of them were situated in the turns or loops between elements of regular secondary structure. Surprisingly, many others assumed regular a-helical or B-sheet structures. Only the 129 residues in the BC-BD loop and the BE-BF loop and Asn115 in the BH-BI loop showed weak or no cross-peaks in the lH-‘SN HSQC spectrum. The residues in the longest loop between BG and BH showed the highest intensities in the lH-‘SN HSQC spectrum. By contrast, residues in the long loop of holo-CRABPII had high temperature factors in the X-ray structure. Interestingly, most of these residues that showed weak or no cross-peaks in the lH-‘SN HSQC spectrum are located in the RA-binding pocket (Figure 3.7). Among the 21 residues that constitute the RA-binding pocket (Kleywegt etal., 1994), 16 residues exhibited weak or no detectable cross-peaks in the 'H-‘sN HSQC spectrum. The cluster of the weak or missed lH-‘SN correlations suggested that the RA- binding pocket of CRABPII is rather dynamic in the absence of RA. It is unlikely that these peaks were affected by water saturation, because the lH-‘SN HSQC spectrum was recorded with a “water flip-back” pulse sequence that minimized water saturation and dephasing (Grzesiek & Bax, 1993). Furthermore, for several weak or missing IH-‘SN correlations, the amide protons remained two days after the lyophilized protein was dissolved in D20 and showed cross-peaks in the fingerprint regions of the homonuclear 2D spectra recorded in D20 (Figure 3.3), indicating that these amide protons exchange slowly with water and should not be perturbed by water saturation. Rather, the lH-‘SN correlations were most likely weak or missing as the result of broadening of the amide proton and/or nitrogen resonances due to conformational exchange at rates intermediate on the NMR time scale. For Lys8, Ile10, Leu19, Leu121 and Argl32-Vall35, clearly, line broadening of their nitrogen resonances was responsible for the weak peaks in the lH-‘SN HSQC spectrum because their amide protons exhibited sharp signals and showed cross-peaks with a-protons in 2D homonulcear spectra recorded in D20 (Figure 3.3). A 130 Figure 3.7. The relative peak intensity of the 2D lH—‘SN HSQC spectrum of apo- CRABPII color-coded along the C“ trace of the solution structure. The strong peaks are in white, the weak peaks in pink and the missing peaks in cyan. RA (green) is positioned on the basis of the superposition of the solution structure of apo-CRABPII and the crystal structure of holo-CRABPII. l3l 132 slow conformational exchange with a lifetime of at least 40 ms was observed as described in Chapter 2, resulting in two sets of cross-peaks for Trp87. It was noted that Trp87 is located at the N-terminus of BF far away from the RA-binding pocket. The residues in the RA-binding pocket likely undergo intermediate conformational exchange, probably on the microsecond to millisecond time scale, because only one set of NMR signals was observed for these residues. The clustering of the mobile residues and the relatively low frequency of conformational exchange suggest that the RA-binding pocket may undergo collective motions in the absence of the RA ligand. Apart from the RA-binding pocket, the structure of apo-CRABPII appears well ordered and less dynamic. Most residues, including those at the turns and loops, showed intense cross-peaks in the 2D lH-‘SN HSQC spectrum. Many of them also showed cross- peaks in the fingerprint regions of the 2D homonuclear spectra recorded in D20 (Figure 3. 3). Remarkably, the hydroxyl protons of Ser83 and Thr107 were observed in 2D homonuclear spectra recorded in H20, and the indole e'-proton of Trp7 was detected in 2D homonuclear spectra recorded in D20, suggesting that these side-chains are not only buried but also immobile. The dynamic nature of the RA-binding pocket is consistent with biochemical and NMR studies of CRABPI. Limited proteolysis showed that helix 012 of CRABPI is significantly more susceptible to proteolysis in the apo-form than in the holo-form (Jamison et al., 1994), suggesting that RA binding induces a conformational change or that helix 012 is more mobile in the apo-form. Hydrogen exchange measurements revealed that helix a2 and the BC-BD loop of CRABPI have much higher exchange rates in the apo-form than in the holo-fonn (Rizo, et al., 1994), indicating that these parts of the 133 molecule are more dynamic in the apo-form. Other iLBPs also have been found to exhibit similar dynamic properties. For example, it was shown by NMR that the second helix of rat intestinal F ABP is disordered in solution (Hodsdon and Cristola, 1997a, b). Is CRABPII dimeric in solution? Unlike other iLBPs, apo-CRABPI was crystallized in a dimeric form (Thompson et al., 1995). The crystalline dimer is held together by an intermolecular B-sheet formed by the BD strands of two CRABPI molecules, resulting in a 20-stranded double B-barrel with slightly more open RA-binding pockets. It was suggested that dimerization may be the mechanism by which CRABPI opens the ligand entrance so that RA can enter or exit the binding pocket without steric hindrance. In light of the potential significance of the dimerization to the ligand entrance problem, we carefully examined the possibility of the formation of the intermolecular B- sheet. We concluded that apo-CRABPII is predominately monomeric in solution for the following reasons. (i) On a Sephadex G-50 gel filtration column, apo-CRABPII eluted with a volume characteristic of monomeric CRABPII. (ii) For the most part, the line widths of the NMR signals of CRABPII were consistent with a monomeric structure. (iii) None of the predicted NOEs indicative of formation of the intermolecular B-sheet was observed. For example, according to the dimeric crystal structure of apo-CRABPI, the distance between H“ of Thr61 of molecule A and H“ of Thr60 of molecule B is 2.5 A but the distance between H“ of Thr61 and H“ of Thr60 of the same molecule is 5.9 A. Therefore, if there is a stable dimer in solution, we should see a NOE cross-peak between the two protons. On the other hand, if there is no stable dimer, there would be no NOE between the two protons. The two protons were sequentially assigned, but we did not see any NOE cross-peak between them. Although the formation of transient dimers could not 134 be ruled out, the transient dimers, if they exist, can only be a minor population under the conditions of the NMR or gel filtration experiments. What is the likelihood of the dimerization in viva? Since the cellular concentration of CRABPII is much lower than the concentrations of the NMR samples, the concentration of dimeric species would be even lower in vivo. Implications for RA binding. The crystal structures of holo-CRABPI and CRABPII suggested that RA cannot enter or exit the ligand binding pocket of the proteins in the absence of significant conformational changes. On the basis of limited proteolysis of several intracellular lipid binding proteins, it was proposed that a rigid body movement of the helix-turn-helix motif may serve as the mechanism by which the family of proteins opens up their ligand binding pocket (Jamison et al., 1994). In light of the solution structure of apo-CRABPII, it is unlikely that the two helices move as rigid rods as previously proposed. Rather, the second helix of the helix-tum-helix motif is flexible and undergoes conformational exchange. The dynamical nature of the second helix may be the cause of its high susceptibility to proteolysis. Based on the comparisons of the crystal structures of apo- and holo-CRABPI, it was suggested that movement of the BC—BD loop is responsible for the opening of the ligand entrance (Thompson et al., 1995). Furthermore, it was also proposed that the movement of the BC—BD loop is dependent on the formation of an intermolecular B-sheet as mentioned above. Comparison of the solution structure of the apo-CRABPII and the crystal structure of holo-CRABPII confirms the hypothesis that BC—BD loop moves when RA is bound. However, the movement of the BC—BD loop that opens the binding pocket 135 is shown by NMR not to involve dimerization of the protein because CRABPII is predominately monomeric in solution. The solution structure of apo-CRABPII suggests that the ligand entrance is widely open and readily accessible to RA. The enlargement of the ligand entrance of apo- CRABPII compared with holo-CRABPII is mainly due to a concerted conformational change in three structural elements, namely the second helix, the BC—BD loop and the BE—BF loop. A widely opened entrance may be essential for binding of RA, because RA is a long, relatively rigid, negatively charged molecule. A narrow entrance lined with the hydrophobic side-chains would result in unfavorable interactions with RA. The entry of RA into the deep binding pocket may be further facilitated by the positively charged potentials generated by Arg29, Arg59, Argl 11 and Arg132 (Chen et al., 1998). Binding of RA induces significant conformational changes in CRABPII. Furthermore, the interactions between RA and CRABPII may stabilize the structural elements constituting the RA-binding pocket, especially the second helix, the BC—BD loop and the BE—BF loop. 136 Conclusion Using multidimensional NMR spectroscopy, we have determined the first solution structure of a CRABP. Comparison of the solution structure of apo-CRABPII with the crystal structure of holo-CRABPII indicates that the largest conformational differences between the two structures are localized at the ligand entrance. The ligand entrance of apo-CRABPII is greatly enlarged and readily accessible to RA, mainly due to a concerted movement of the second helix and the BC-BD and BE-BD loops. Furthermore, the ligand binding pocket of apo-CRABPII is rather dynamic as indicated by analysis of the cross- peaks intensities of the lH-‘SN HSQC spectrum. CRABPII is predominately monomeric in solution as determined from NMR and biochemical evidence. Although the formation of transient dimers could not be ruled out, dimerization apparently is not a prerequisite for entry of RA into the ligand binding pocket of CRABPII. 137 References Anil-Kumar, Ernst, R. R., & Wiithrich, K. (1981) Biochem. Biophys. Res. Commun. 95, 1-6. Banaszak, L., Winter, N., Xu, Z., Bernlohr, D. A., Cowan, S. W., & Jones, T. A. (1994) Adv. Protein Chem. 45, 89-151. Basus, V. J. (1989) Methods Enzymol. 1 7 7, 132-149. Bax, A., Clore, G. M., & Gronenbom, A. G. (1991) J. Magn. Reson. 88, 425-431. Bax, A., Griffey, R. H., & Hawkins, B. L. (1983) J. Magn. Reson. 55, 301-315. Bodenhausen, G., & Ruben, D. J. (1980) Chem. Phys. Lett. 69, 185-188. Brooks, B. R., Bruccoleri, R. E., Olafson, B. D., States, D. J ., & Karplus, M. (1983).]. Comput. Chem. 4, 187-217. Briinger, A. T. (1992) X-PIOR: A System for Crystallography and NMR. Yale University, New Haven, CT. Chambon, P., Olson, J. A., & Ross, A. C. (coordinators) (1996) The retinoid revolution. FASEB. J 10, 938-1048. Chen, X., Tordova, M., Gilliland, G. L., Wang, L., Li, Y., Yan, H., and 11, X. (1998) J. Mol. Biol, in press. Clore, G. M., & Gronenbom, A. G. (1989) CRC Crit. Rev. Biochem. Mol. Biol. 24, 479- 564. Griesinger, C., Otting, G., Wiithrich, K., & Ernst, R. P. (1988).]. Am. Chem. Soc. 110, 138 7870-7872. Grzesiek, S., & Bax, A. (19923) J. Magn. Reson. 96, 432-440. Grzesiek, S., & Bax, A. (1992b) J. Am. Chem. Soc. 114, 6291-6293. Grzesiek, S., Anglister, J., & Bax, A. (1993) J. Magn. Reson. Ser. B 101, 114-119. Grzesiek, S., & Bax, A. (1993).]. Am. Chem. Soc. 115, 12593-12594. Hodson, M. E., & Cistola, D. P. (1997a) Biochemistry 36, 1450-1460. Hodson, M. E., & Cistola, D. P. (1997b) Biochemistry 36, 2278-2290. Jamison, R. S., Newcomer, M. E., & Ong, D. E. (1994) Biochemistry 33, 2873-2879. Jeener, J ., Meier, B. H., & Ernst, R. R. (1978) J. Chem. Phys. 71, 4546-4553. Kay, L. E. (1995) Prog. Biophys. Molec. Biol. 63, 277-299. Kay, L. E., & Bax, A. (1990) J. Magn. Reson. 86, 120-126. Kay, L. E., Ikura, M., Tschudin, R., & Bax, A. (1990) J. Magn. Reson. 89, 495-514. Kay, L. E., Keiffer, P., & Saarinen, T. (1992) J. Am. Chem. Soc. 114, 10663-10665. Kay, L. E., Xu. G. Y., Singer, A. U., Muhandiram, D. R., & Forman-Kay, J. D. (1993) J. Magn. Reson. BIO], 333-337. Kleywegt, G. J ., Bergfors, T., Senn, H., 1e Motte, P., Gsell, B., Shuao, K., & Jones, T. A. (1994) Structure 2, 1241-1258. Laskowski, R., MacArthur, M. W., Moss, D. S., & Thornton, J. M. (1993) J. Appl. Crystallogr. 26, 283-291. Laskowski, R., Rullmannn, J. A., MacArthur, M. W., Kaptein, R., & Thornton, J. M. (1997) J. Biomol. NMR 8, 477-486. Logan, T. M., Olejniczak, E. T., Xu, R. X., & Fesik, S. W. (1993) J. Biomol. NMR 3, 225-231. 139 Mangelsdorf, D. J ., Umesono, K., & Evans, R. M. (1994) In The Retinoids: Biology, Chemistry, and Medicine (Sporn, M. B., Roberts, A. B., & Goodman, D. S., eds.), pp. 319-349, 2nd ed., Raven Press, New York. Marion, D., Ikura, M., Tschudin, R., & Bax, A. (1989a) J. Magn. Reson. 85, 393-399. Marion, D., Ikura, M., & Bax, A. (1989b) J. Magn. Reson. 84, 425-430. McCoy, M. A., & Mfieller, L (1992a) J. Am. Chem. Soc. 114, 2108-2112. McCoy, M. A., & Miieller, L (1992b) J. Magn. Reson. B98, 674-679. Muchmore, D. C., McIntosh, L. P., Russell, C. B., Anderson, D. E., & Dahluist, F. W. (1989) Methods Enzymol. I 7 7, 44-73. Muhandiram, D. R., & Kay, L. E. (1994) J. Magn. Reson. 3103, 203-216. Nilges, M., Clore, G. M., & Gronenbom, A. M. (1988) FEBS Lett. 229, 129-136. Ong, D. 13., Newcomber, M. E., & Chytil, F. (1994) In The Retinoids: Biology, Chemistry, and Medicine (Spom, M. B., Roberts, A. B., & Goodman, D. S., eds.), pp. 283-317, 2“d ed., Raven Press, New York. Piantini, U., Sorensen, O. W., & Ernst, R. P. (1982) J. Am. Chem. Soc. 104, 6800-6801. Rizo, J ., Liu, Z. P., & Gierasch, L. M. (1994) J. Biomol. NMR 4, 741-760. Sacchettini, J. C., Gordan, J. 1., & Banaszak, L. J. (1988) J. Biol. Chem. 263, 5815-5819. Shaka, A. J ., Keeler, J., Frenkiel, T., & Freeman, R. (1983).]. Magn. Reson. 52, 335-338. Shaka, A. J., Lee, C. J., & Pine, A. (1988) J. Magn. Reson. 77, 274-293. Shaka, A. J ., Barker, P. B., & Freeman, R. (1985) J. Magn. Reson. 64, 547-552. Smallcombe, S. H. (1993) J. Am. Chem. Soc. 115, 4776-4785. Sporn, M. B., Roberts, A. B. & Goodman, D. S. (1994) The Retinoids: Biology, Chemistry, and Medicine, 2"d ed., Raven Press, New York. 140 States, D. J ., Haberkon, R. A., & Ruben, D. J. (1982) J. Magn. Reson. 48, 286-292. Thompson, J. R., Bratt, J. M., & Banaszak, L. J. (1995) J. Mol. Biol. 252, 433-446. Wittekind, M., & Miieller, L. (1993) J. Magn. Reson. 101, 201-205. Wiithrich, K., Billletter, M., & Braiin, W. (1983) J. Mol. Biol. 169, 949-961. Zhang, J ., Liu, Z., Jones, T. A., Gierasch, L. M., & Sambrook, J. F. (1992) Proteins: Structure, Function, and Genetics 13, 87-99. Zhang, 0., Kay, L. E., Olivier, J. P., & Forman-Kay, J. D. (1994) J. Biomol. NMR 4, 845-858. Zhu, G., & Bax, A. (1992).]. Magn. Reson. 90, 405-410. Zuiderweg, F. R. P., & F esik, S. W. (1989) Biochemistry 28, 2387-2391. CHAPTER 4 Solution Structure of a Site-Directed Mutant (R1 11 M) of Human CRABPII Introduction The crystal structures of holo-CRABPs reveal that RA is buried in the binding pocket with its carboxyl group interacting with Argl l 1, Arg132 and Tyrl34 (CRABPII numbering) at the bottom of the pocket (Kleywegt et al., 1994). Arg132 and Tyrl34 interact directly with the carboxyl group of RA, whereas the interaction between Argl 11 and the carboxyl group of RA is mediated by a water molecule. Electrostatic calculations based on a modeled structure of CRABPI have suggested that the electrostatic repulsion between the guanidinium groups of the above two conserved arginine residues provides the conformational flexrbility to CRABPs for RA to enter the binding pockets (Zhang et al., 1992). Site-directed mutagenesis and competitive binding studies have shown that both Argl 11 and Argl32 are important for CRABPII to bind RA but Argl ll contributes more to the binding energy than Argl32 (Chapter 2). This chapter describes the structure determination of the CRABPII mutant protein R1 1 1M in solution by NMR spectroscopy. 141 142 The results suggest that Argl 11 play a critical role in determining the structure and dynamical properties of CRABPII. 143 Experimental Procedures Sample preparation. Unlabeled amino acids, nucleotides, vitamins and D- glucose were purchased from Sigma (St. Louis, Missouri). ['5N]-labeled amino acids and 'SNH4C1 were purchased from Cambridge Isotopic Laboratories. Site-directed mutagenesis and expression, and the purification of the recombinant mutant protein R11 1M have been described in Chapter 2. Uniform 15N-labeling and ”N- labeling of leucine were performed as detailed in Chapter 3. The procedure for 15N- labeling of isoleucine and phenylalanine was the same as that for isoleucine and leucine. For 15N-labeling of lysine, the expression strain BL21(DE3)/pLysS containing the plasmid pET-l7b/R111M was used. The strain was grown in M9 medium supplemented with 110 mg/L [‘SN]-lysine added ten minutes before the induction with IPTG. D20 sample was prepared by dissolving the lyophilized protein in PBS buffer (20 mM sodium phosphate, 100 mM sodium chloride in 99.9% DZO, pD 7.5, uncorrected). The protein solution was incubated for six hours at 37°C and lyophilized again. The lyophilized protein was then dissolved in 99.96% D20. HZO samples were prepared by directly concentrating the protein fractions eluted with the PBS buffer (pH 7.3) from a gel- filtration column (G-50) at the final step of the purification procedure (Chapter 2). About 10% D20 was added into the concentrated protein solutions. The protein concentrations of the NMR samples were ~ 2 mM. 144 NMR spectroscopy. All NMR spectra were recorded at 27°C. Homonuclear 2D spectra of unlabeled R11 1M in D20 and heteronuclear 2D lH-‘SN HMQC (Miieller, L. 1979; Bax et al., 1983) spectra of selectively l5N-labeled proteins in H20 were acquired on a Varian VXR-SOOS spectrometer. All 2D experiments were recorded in the phase- sensitive mode with F1 quadrature detection achieved by the hypercomplex method (States et al., 1982). The relaxation delay was 1.5 second, and the residual water resonance was suppressed by low-power presaturation. The spectral width was 7238 Hz for 'H and 2000 Hz for l5N. The following 2D homonuclear spectra were recorded: one DQF-COSY (Piantini et al., 1982) spectrum, one E.COSY (Griesinger et al., 1985) spectrum, three NOESY (Jeener et al., 1978; Kumar et al., 1981) spectra with mixing times of 50, 100 and 150 ms, and three clean-TOCSY (Griesinger et al., 1988) spectra with mixing times of 20, 45 and 75 ms. The time domain data were composed of 2048 X 320 complex points (128 transients) for the E.COSY and DQF-COSY experiments and 2048 X 256 complex points (80 transients) for all other homonuclear 2D experiments. The 2D lH-‘SN HMQC spectra were recorded with 1024 (‘H) X 64 (”N) complex points (32 transients). The data were processed with the program VNMR v. 5.1 (Varian Associates). A ~30°-shified sine-bell was applied in both dimensions for processing the DQF-COSY and E.COSY data, and a ~45°-shified sine-bell for the lH-‘SN HMQC data. A gaussian and minus LB combination was applied in F2 dimension and a ~75°-shifted sine-bell in F1 dimension for processing the NOESY and TOCSY data. The time domain data were then zero-filled to 2048 X 512, 4096 X 2048, 4096 X 2048, and 8192 X 4096 real points for the 2D HMQC, TOCSY, NOESY, DQFCOSY and E.COSY, respectively. 145 A five-order polynomial was applied for baseline correction in the F2 dimension after Fourier transformation. Heteronuclear NMR experiments of the uniformly ”N labeled protein were carried out on a Briiker DMX-600 spectrometer. The carriers for 1H and ”N were set at 4.70 ppm and at 115.7 ppm, respectively. A 2D lH-”N HSQC spectrum (Bodenhausen & Ruben 1980; Cavanagh et al., 1991; Palmer et al., 1991; Kay et al., 1992) was acquired with coherent selection by pulsed field gradients and sensitivity enhancement. The acquisition times and complex data points were IH (F2) 113.9 ms, 1024 and ”N (F1) 56.9 ms, 256 (32 transients). Both 3D lH-”N TOCSY-HMQC (Marion et al., 1989b) and NOESY- HMQC (Zuiderweg and Fesik, 1989; Marion et al., 1989c) spectra were recorded with the following acquisition times and complex data points: lH (F3) 113.9 ms, 1024; ”N (F2) 11.9 ms, 32; 'H (Fl) 14.3 ms, 128 (8 transients and 45 ms mixing time for TOCSY- HMQC, and 16 transients and 150 ms mixing time for NOESY-HMQC). In both 3D experiments the solvent resonance was suppressed by low-power presaturation. Quadrature detection in indirectly detected dimensions were accomplished by the States- TPPI method (Marion et al., 1989a). Homonuclear isotropic mixing in TOCSY-HMQC experiment was achieved by a DIPSI-3 sequence (Shaka et al., 1987). Two dimensional IH-”N HMQC-J (Key and Bax, 1990) data was recorded on a Briiker DMX-750 spectrometer at the National Magnetic Resonance Facility at Madison with the following acquisition times and complex data points: lH 136.4 ms, 2048 and ”N 232.7 ms, 512 (16 transients). A low power GARP-I sequence (Shaka et al., 1985) was applied for ”N broad-band decoupling during the data acquisition periods of these experiments. The spectra were processed with the program NMRPipe (Delaglio et al., 1995). A ~45°- 146 shifted sine-bell was applied in both F2 ('H) and F l (”N) dimensions for both 2D lH-”N HSQC and HMQC-J spectra. The lH-”N HSQC and HMQC-J data were zero-filled to 2048 (F2) x 512 (F1) and 2048 (F 2) x 1024 (F1) real points, respectively. A ~75°-shifted sine-bell was applied in both Fl ('H) and F3 ('H) dimensions and a cosine-bell in F2 (”N) dimension for processing of the 3D data. The first points of F l and F2 dimensions were calculated by linear prediction. Both 3D data were zero-filled to 2048 (F3) x 128 (F2) x 512 (F1) real points. The aliphatic part in the F3 dimension was removed after Fourier transformation. Resonance assignment. The sequential assignments were carried out following the standard procedure developed by Wiithrich and others (Wiithrich, 1986; Clore & Gronenbom, 1989 and 1991). Briefly, the spin systems were identified by the analysis of the homonuclear 2D TOCSY and DQF-COSY spectra and the 3D lH-”N TOCSY- HMQC spectrum or by selective labeling. Sequential linking was obtained from NOE connectivities observed in 3D lH-”N NOESY-HMQC experiment. The stereospecific assignments of B-mythylene protons were obtained by the joint analyses of the qualitative “JG, coupling constants derived from the E.COSY and DQF-COSY spectra and the relative volumes of the NOE cross peaks between H“—H“ and HN—H“ from the 2D homonuclear NOESY and 3D lH-”N NOESY-HMQC spectra (Basus, 1984). The methyl protons of valine and leucine residues and the ring protons of aromatic residues were stereospecifically assigned as detailed in Chapter 3. Secondary structural elements were identified according to their characteristic NOE patterns. Derivation of structural restraints. Three types of structural restraints were derived from the experimental NMR data: interproton distance restraints, torsion angle 147 restraints and hydrogen bond restraints. Approximate interproton distance restraints were derived from the homonuclear 2D NOESY with 100 and 150 ms mixing times, and the 3D ”N-edited NOESY-HMQC with 150 ms mixing time. The same procedure for establishing the upper limits of interproton distances was followed as described in Chapter 3. The only difference was that the NOE cross peaks from the 3D lH-”N NOESY-HMQC were picked by the programs CAPP and PIPP (Garret et al., 1991). The NOEs were classified as strong, medium, weak and very weak corresponding to interproton distance 1.8—2.7 A (1.8—2.9 A for NOEs involving NH protons), 1.8—3.3 A (1.8—3.5 A for NOEs involving NH protons), 1.8—5.0 A and 2.5—5.0 A, respectively. Pseudoatom corrections were made for non-stereospecifically assigned methylene protons and methyl protons (Wiithrich et al., 1983). Additional 0.5 A was added to the upper bounds for methyl protons. The backbone torsion angle 11) was restrained according to the value of “JNHQ obtained from 2D HMQC-J experiment. It was restrained —120° i 40° for UN”, 2 7.5 Hz, and —60° i 20° for 3.1M,CL .<_ 5.0 Hz and when NOE data indicated the presence of a helix. Backbone hydrogen bond restraints were derived from the cross peaks remaining in the fingerprint regions of the homonuclear 2D TOCSY and NOESY spectra recorded in D20. The HN---O and N---O were restrained to be 2.0 i 0.5 A and 3.0 i 0.3 A, respectively. Structure calculation. The structures were calculated as detailed in Chapter 3 except that an additional square-well potential fimction with force constant of 200 kcal.mol".rad'2 was applied for the dihedral angle restraints. 148 Results Isotopic enrichment. The level of ”N enrichment in the uniform l“N-labeling was high as judged by the signal to noise ratio of the 2D 'H-”N HSQC spectrum (Figure 4.1). No l5N isotopic scrambling to other residues was observed for selective labeling of isoleucine, leucine, lysine and phenylalanine. For example, the number of the cross peaks in the 2D 'H-‘SN HMQC spectrum of the [”N]-leucine-labeled protein was exactly equal to the total number of leucine residues in R11 1M (Figure 4.2). Spin system identification. Aliphatic spin systems were first identified by the analysis of the homonuclear 2D DQF-COSY and TOCSY spectra in D20 and then extended to amide resonances by the 3D lH-”N NOESY-HMQC experiment. Most of the threonine, alanine, valine and glycine residues could be identified easily by their unique cross-peak patterns. The majority of the residues belonging to either AMX or AM(PT)X spin system could also be identified. However, except for a few, the residues with long side-chains (lysine, leucine, isoleucine, arginine and proline) could not be distinguished from each other because of severe spectral overlapping in the upfield region (1.00 ppm- 2.00 ppm) and the inefficiency of the TOCSY experiments for long-relayed coherence transfer. Their spin system identifications were achieved or assisted by selective l“N- labeling. All the leucine, isoleucine and lysine residues could be identified by selective ”N-labeling. Furthermore, proline and arginine could also be distinguished from each other. In general, the H“ protons of prolines have larger down field chemical shifts than 149 Figure 4.1. Gradient- and sensitivity-enhanced 2D lH—‘SN HSQC spectrum of uniformly ”N enriched of apo-R1 1 1M. Sequential assignments are indicated with one-letter amino acid code and residue number. The side-chain amides of Asn and Gln are indicated by #. 150 105 623 I 0 6370454 04514 610200 G78 G104 ‘ - —110 ON,“ N14» 0 T110 T g 074* N115“ ‘ Q N 1 wean N “337 '\' 0N“ .889 3.4 8108. N9 0356;; N28!“ :4# N9”! N1 077 N115#IN25# N25“ 698 T‘s 90'9" 69.771 ’115 812 o L 9 4 N14 0 7‘2 oovvo109 T49 0116 T122 V339 0g T114 °.E16° v 0 F65 513°01‘10" K330913118 C8129163 E1476 2 110. T61 T17ROE72A126°5N°M027L38K9R59031h120 ”"5 M111 F15 K66 .K 2 419R110L 55 W87 E7370012,,“26 “1569“" L1 15 7 M13 :3; .133 V1. 14%,: 6197 V 8. T60E627Y15 100 434 (ppm) . 0126 00208? 1 F3 “3' T:E1 74136 1228 492 “’7’ F50 . . —125 9 W7N‘0 ‘513 L119 "29 T131 K11 1120 583 9 .0 0 0 A1 5 L113 Y134 "31321193 v 7 L99 '9 .04“ t, A40/N25 5137 596 K34 1538 (:95 45 Q 0 0542 E46 —130 warm C13“ V85” E70 0 K53 -'135 0 L121 1 1 1 1 1 1 1 10 5 10 0 9.5 .5 8 0 7 5 7 0 o 1 a H(ppm) 151 Figure 4.2. 2D 'H—‘SN HMQC spectrum of the [‘SN]-leucine selectively labeled apo- R1 1 1M. Sequential assignments are indicated with one-letter amino acid code and residue number. 152 0 L19 0 L28 L100 0 0 L18 1.119 a 1.99 0 L113 L121 [:30 — 126.0 120.0 115.0 DZ (ppm) 130.0 135.0 9.6 910 8.4 7.8 D1 (ppm) 153 those of arginines. For example, Prol was identified in this way. The cross peaks corresponding to H“ protons of four arginines could be identified in the 2D lH-‘SN HSQC spectra. Consequently, the H‘ was linked to the HY and H‘3 protons, and eventually to H“ protons by the 3D lH-‘SN TOCSY-HMQC experiment. The residues belonging to AMX spin system were further distinguished from each other based on distinct chemical shifis and NOEs. Four of the seven serines could be identified by the down field positions of their HB protons. All four asparagine residues could be identified by the NOE cross-peaks between Hfl and H6 protons observed in the 3D 'H-‘SN NOESY-HMQC spectrum. All five phenyalanines were identified by selective labeling, and confirmed by the homonuclear 2D NOESY experiments. One tyrosine and two tryptophan residues could be identified after the assignments of their aromatic ring protons. Two glutamines could also be differentiated from the other residues belonging to the same AM(PT)X spin system by the NOE cross-peaks between H7 and H‘ protons. All remaining residues were classified into several groups, and identified later along with sequential assignment. The positional lH resonance assignments of the side-chains were obtained mainly through the ZD DQF- COSY spectrum, and three TOCSY spectra with different mixing times. Remarkably, the hydroxyl protons of both Ser83 and Thr107 were observed in both homonuclear and heteronuclear NMR experiments performed in H20 (data not shown), and were assigned by the homonuclear TOCSY and NOESY experiments. The indole 8' proton of Trp7 was observed in the NMR spectra recorded on D20 sample. The ring protons of aromatic residues. The assignments of aromatic ring protons were derived from the 2D DQF-COSY spectrum, and three TOCSY spectra with different mixing times. The ring protons of phenylalanines and tyrosines were linked to their HB, 154 HCl and HN protons by NOE cross peaks between H‘3 and H5 protons and between H"1 and H5 protons observed in the 2D homonuclear NOESY. The H" and H51 protons of tryptophans could be assigned by the 3D lH-‘SN NOESY-HMQC experiment, and H83 protons by NOEs between H“1 and H‘13 protons and between HB and H£3 protons observed in the homonuclear 2D NOESY experiments. Sequential assignment. The sequential assignment (Table 4.1) was primarily based on sequential NOE connectivities such as daN, dNN and day The assignment was initiated after over sixty percent of the total 137 residues had been identified. Most residues from the ten B-strands and the first a-helix could be easily assigned (Figure 4.3). The sequential assignments of ValZ6—Lys38 posed a challenge. Val26, Arg29, Ile31 and Lys38 were weak or very weak in the 2D 1H-‘sN HSQC spectrum. No correlations between their amide and aliphatic protons could be observed in the 3D 'H-‘SN TOCSY- HMQC spectrum. Only a few NOEs useful for sequential assignments were observed in the 3D lH-‘SN NOESY-HMQC spectrum. The 1H chemical shifts of the four alanines (Ala32, Ala34, Ala35 and Ala36) were quite similar. The 'H and 15N resonances of both Va126 and Arg29 were assigned tentatively by comparison with the assignments of wild- type holo-CRABPII (Chapter 5), and a few weak sequential NOE peaks. The amide resonances of both residues became stronger upon RA binding. The aliphatic protons of Va126 were assigned by spin system identification. The assignments of Ile3l and Lys38 were helped greatly by selective labeling, and by comparison with the assignments of wild-type holo-CRABPII. The spin system of Asnl 15 could be easily identified by the homonuclear NMR experiments but its sequential assignment was achieved only at the later stage because of the weakness of its peaks in all l5N-edited spectra. The H"l and HB 155 Table 4.1. 'H and '5N resonance assignments of apo-R1 1 1M in PBS at 27°C Residue P 1 N2 F3 S4 G5 N6 W7 K8 I9 110 R1 1 S 12 E13 N14 F15 E16 E17 L18 L19 K20 V21 L22 623 V24 N25 V26 M27 L28 R29 K30 131 A32 V33 A34 A35 A36 S37 K38 P39 A40 V41 E42 143 K44 15N 123.2 112.6 110.7 119.0 124.4 125.2 126.1 118.9 120.6 115.6 124.1 117.5 120.5 117.8 118.5 121.8 115.3 118.7 121.4 118.0 106.7 121.8 127.4 120.7 119.6 119.1 120.9 118.1 119.3 123.5 117.1 122.0 119.0 120.2 112.2 120.1 127.4 120.8 129.4 123.7 129.1 HN 9.76 8.39 9.36 8.17 9.31 10.06 8.62 9.23 7.43 8.32 8.83 9.10 8.54 8.80 8.20 7.62 7.48 8.07 7.81 7.27 7.65 7.62 8.59 8.20 7.91 7.74 8.66 7.66 7.56 8.01 8.41 7.58 8.19 8.20 7.48 7.68 8.60 7.78 9.19 9.20 9.48 HO. 4.14 5.11 4.38 4.31 4.13, 3.97 5.24 5.08 5.27 3.67 4.59 4.60 5.28 4.83 4.96 4.60 3.32 3.74 3.75 3.64 3.89 3.79 3.97 4.00, 3.62 4.60 4.46 3.51 4.15 4.11 3.92 4.07 4.10 3.72 4.20 4.35 4.54 4.60 4.90 4.90 4.98 5.40 5.30 4.63 1.13 2.21, 1.75 2.99, 2.83 3.40, 3.22 3.99, 3.99 2.74, 2.56 3.08, 2.91 1.85, 1.85 1.39 1.87 1.72, 1.72 3.70, 3.61 1.93, 1.93 3.38, 3.27 3.14, 2.97 1.88, 1.88 1.97, 1.78 1.56, 1.56 1.61, 0.81 1.96, 1.69 2.19 1.70, 1.59 1.88 3.15, 2.80 2.05 2.09, 2.03 1.82, 1.66 1.93, 1.93 2.02 1.37 2.10 1.47 1.40 1.40 3.92, 3.82 1.82, 1.40 1.59 1.29 2.06, 2.10 2.25 1.84, 1.84 Others 1171.92, 1.94;H‘5 3.13, 3.11 H’ 7.64, 6.89; N7 112.5 2, 6H 7.43; 3, 5H 7.20; 4H 7.27 2H 6.80; 4H 7.04; 5H 6.81; 6H 7.21; 7H 7.11 H‘ 9.28; Nc 125.2 I-P 1.65, 1.41; H6 0.86, 0.86; Hc 2.86, 2.86 H1 0.92, 0.92; H15 0.60; 1-17'“ 0.69 1171.21 1.21;H51.00;HV“‘ 0.81 H" 1.59; H‘5 3.07; H‘ 7.29; N‘ 83.8 H7 2.14, 2.21 H1 7.29, 6.91; N1 109.8 2, 6H 7.21; 3,5H 6.94; 4H 7.02 H” 1.98, 2.18 H" 2.42, 2.24 H" 0.84; 1-15 0.79, 0.79 H" 1.09; H‘5 0.38, 0.22 HY 1.41, 1.41;H51.20, 0.95; H‘ 3.22, 2.96 H1 1.10, 0.94 H" 0.96; H‘5 0.86, 0.69 H" 0.92, 0.82 H" 0.95, 0.90 H1 2.60, 2.45 H7 1.35; H’5 0.90, 0.79 H7 1.74, 1.55; H‘5 1.47, 1.47; HE 2.92, 2.92 1171.85, 1.85;H‘"“ 1.01 1170.95, 0.95 H7 1.07, 0.93; Hz5 3.47, 3.47 H" 0.43, 0.41 H" 2.30, 2.30 W 1.66, 1.28; H‘5 0.69; H1m 0.91 1171.54, 1.35; H‘ 2.86, 2.78 156 Residue Q45 E46 G47 D48 T49 F50 Y5 l 152 K53 T54 SSS T56 T57 V58 R59 T60 T6 1 E62 163 N64 F 65 K66 V67 G68 E69 E70 F71 E72 E73 Q74 T75 V76 D77 G78 R79 P80 C8 1 K82 S83 L84 V85 K86 W87 E88 889 E90 N91 K92 M93 lSN 127.4 128.9 117.9 126.7 116.9 124.7 122.6 123.8 131.9 124.6 121.2 115.5 110.9 113.0 119.3 122.9 120.0 122.9 117.9 119.9 117.1 120.7 127.5 114.8 120.4 130.5 121.5 119.7 121.8 117.9 111.8 122.4 113.1 108.4 121.8 118.7 120.0 124.9 120.9 130.3 123.7 120.5 127.7 113.2 116.6 111.1 120.2 126.5 8.59 8.55 9.04 8.51 8.06 9.00 8.32 8.42 9.26 8.96 9.16 8.36 8.34 7.43 7.38 8.76 8.95 8.62 8.50 8.00 8.29 9.20 8.83 8.77 7.90 9.18 8.99 8.61 8.42 7.78 9.18 9.26 8.75 7.93 6.86 8.78 8.83 9.02 7.33 9.21 8.39 8.54 9.25 8.63 8.69 7.98 7.74 9.27 HO. 4.50 4.40 3.88, 3.62 5.19 4.75 4.60 5.20 4.72 5.17 5.10 5.34 4.85 4.05 4.19 4.60 5.23 4.65 5.13 4.85 5.59 4.67 5.31 3.05 4.40, 3.63 4.81 4.91 5.15 5.09 4.63 5.35 4.62 3.79 4.40 4.23, 3.58 4.63 4.70 5.08 4.79 4.93 4.24 4.40 4.52 5.02 4.38 4.61 3.77 4.01 5.71 5.16 HB 1.86, 1.86 1.84, 1.68 3.16, 2.74 4.18 1.74, 1.74 3.03, 2.77 1.59 1.77, 1.69 3.89 3.75, 3.75 4.72 4.27 2.02 1.72, 1.72 3.94 4.01 1.93, 1.93 1.86 2.51, 2.51 2.28, 2.28 1.67, 1.54 1.67 2.00, 2.00 2.02, 1.94 3.51, 3.22 2.06, 1.97 1.98, 1.90 1.79, 1.79 4.18 1.98 2.74, 1.98 1.82, 1.58 2.42, 1.90 2.64, 2.13 1.58, 1.51 2.80, 1.55 1.61, 1.29 1.32 1.64, 1.51 3.28, 2.99 1.95, 1.76 4.18, 4.01 1.87, 1.87 3.12, 2.92 1.96, 1.96 1.22, 1.22 Others 11* 1.03, 1.03; H5 7.71, 6.98; N5 106.9 H1 2.03, 2.03 HY 1.02 2, 6H 6.50; 3, 5H 7.06; 4H 6.81 2, 6H 6.94; 3, 5H 6.56 H1 1.38, 1.28; H8 0.59; H“ 0.43 H" 1.24, 1.16; H‘ 2.86, 2.86 H" 0.96 H" 1.16 11’ 1.31 Hy 0.82, 0.82 H" 1.40, 1.40 H1 1.11 H1 1.07 H" 2.10, 2.04 H" 1.36, 0.79; H5 0.91; I-P'"n 1.11 H1 7.51, 6.81;N7113.1 2, 6H 6.20; 3, 5H 6.52; 4H 6.26 H1 1.32, 1.32; H‘ 2.96, 2.96 H" 0.46, 0.25 117 2.22, 2.22 H" 2.24, 2.15 2, 6H 7.26; 3, 5H 7.81; 4H 6.55 H" 2.34, 2.16 H" 2.14, 2.14 H” 2.20, 2.03; H‘5 7.35, 6.58; N‘5 110.3 H" 1.06 H" 0.86, 0.86 H7 1.36, 0.69; H5 3.21, 3.12; H‘ 8.59; N‘ 87.9 H1 2.05, 2.17; H5 4.03, 3.66; H1 1.28; H8 2.82; m 6.26 H1 1.19; 1150.71, 0.71 HY 0.22, —0.03 H7 1.31, 1.22 2H 7.38; 4H 7.98; 5H 6.86; 6H 7.21; 7H 7.58 H‘ 10.50 (1057):; Na 129.6 H7 2.22, 2.12 117 2.12, 2.12 H" 7.37, 6.70; N7 112.6 H" 1.50, 1.50; Hz5 1.23, 0.93; H‘ 2.71, 2.56 H‘1 1.75, 1.75; H“ 0.31 Residue V94 C95 E96 Q97 K98 L99 L100 K101 6102 E103 G104 P105 K106 T107 S108 W109 T110 M111 E112 L113 T114 N115 D116 G117 E118 L119 1120 L121 T122 M123 T124 A125 D126 D127 V128 V129 C130 T131 R132 V133 Y134 V135 R136 E137 15N 118.5 127.6 127.4 121.0 123.6 126.1 121.7 117.6 109.0 117.1 109.1 123.6 118.9 112.9 117.1 110.4 119.7 124.3 126.7 117.2 119.6 115.8 107.7 117.8 124.8 125.8 135.4 117.8 121.2 116.0 125.6 122.5 121.0 123.7 125.7 129.5 124.8 126.3 122.0 126.9 111.8 123.8 127.4 HN 8.64 8.91 7.70 7.31 8.97 8.75 7.70 7.97 8.18 8.08 8.10 8.54 8.75 7.84 8.29 9.44 8.62 8.85 8.17 9.29 9.14 7.75 7.94 7.40 8.64 9.38 9.69 8.85 9.20 8.42 8.44 9.23 8.44 8.39 8.35 9.57 8.06 9.43 8.42 9.73 9.19 8.56 8.15 Ha 4.24 4.46 4.32 4.92 4.60 4.00 4.10 4.40 4.06, 3.75 4.63 4.12, 3.95 4.54 4.42 5.38 4.75 5.72 5.15 5.46 5.78 5.29 4.62 4.46 4.77 4.21, 3.73 5.03 5.04 4.85 5.54 5.75 5.20 5.50 4.88 4.17 4.54 3.99 4.55 5.25 5.00 5.04 4.92 5.54 5.34 3.89 3.93 157 1.13 1.77 0.51,—0.10 1.86, 1.80 2.03, 1.85 1.57, 1.39 1.67, 1.51 1.47, 1.47 1.82, 1.82 1.79, 1.79 2.08, 2.08 1.85, 1.85 4.50 3.83, 3.83 3.58, 3.33 4.19 1.86, 1.83 1.93, 1.93 1.31, 0.77 4.52 2.82 2.82, 2.46 2.04, 1.92 0.95, 0.84 1.84 1.92, 1.36 4.08 1.87, 1.84 3.97 1.06 2.77, 2.73 2.70, 2.70 2.18 1.87 3.03, 2.74 3.77 1.87, 1.84 1.70 3.30, 3.17 2.38 1.31, 1.09 1.85, 1.75 Others H" 0.75, 0.69 H" 2.17, 1.97 H" 2.50, 2.15; Hz5 7.51, 6.68; N15 114.8 H" 1.29; H‘ 2.85 H" 1.39; H5 0.69, 0.69 H" 0.78 H" 1.66, 1.66; H5 1.29, 1.29 H’2.12, 1.99 H" 1.87, 1.80; H5 3.63, 3.35 H" 1.56, 1.40; H5 1.81, 1.81; Hc 2.98, 2.98 H" 1.17, 1-1"I 6.04 2H 6.87; 4H 6.83; 5H 6.41; 6H 6.41; 7H 6.71 H‘ 11.17; N‘c 136.4 H" 1.07 H“ 2.41, 2.25 H" 2.09, 2.09 H" 1.14; H5 0.51, 0.48 H" 1.08 H" 7.54, 6.83; N" 112.7 H" 2.35, 2.20 H" 0.30; H5 —0.3 1, —0.38 H"1.10,1.10;H51.04;H"'" 0.79 H" 1.60; H6 0.77, 0.71 H" 1.08 H" 2.45, 2.37 H" 0.99 H" 0.96, 0.74 H" 0.81, 0.74 H" 1.02 H" 1.75, 1.75; H13 3.02, 1.96; Hc 7.17; N“ 84.4 H" 0.89, 0.84 2, 6H 6.82; 3, 5H 6.68 H" 1.08, 0.91 H" 0.62, 0.23; Hz5 2.82, 2.70; Hc 6.86; N‘ 85.2 H" 2.09, 2.09 "' Two sets of resonances. 158 Figure 4.3. Strip plots extracted from the 3D 15N-edited NOESY-HMQC spectrum of apo-R1 11M showing the sequential and medium-range NOE connectivies of a-helix (01A) (A) and B-strand ([1]) (B). 159 7 7 7 7 7 7777 7 7 7 7 ”cream 0°90 $0 9 0 "9 “9‘: ° 1' "5‘3 e 000097690 3 =15 '6 2:, '00 E éi-‘i '25; j 6 “9‘90 3 ,OW 5.?3‘ 60 ‘ o 6.2—04-? 9.. ea E99 17 0 0 EL 0 "é 'Qoewo 00:11 A; Q «0'00 '§ 0 997.15 77.? ,No .e.. -2 9'0 r1560 0 co 0 :1 9%: 1 ° L? as: o - o 0 g? =10 '° ° 0 —§ m 0 0 E? I“ Q _§ mon 00mm 0 0 Joe 1 10° 7:5 w a o 3 3"0‘ '9 .0“ o" - gpk ._‘z :7. 13 09110990 °° EH» 9° ”717?: 9 ‘7 H GEDGQGO ,, ,gfia 751$ o..- _ no ea 0090 $74 0'1: '3 1 °°®°° “10% O: 1.31-'07- '3 u 9 @009” 70° 0 .2 21¢-- ’2 -E 0% 9° “07 0° ' .2" 7b‘ ° hi < . .0 “mo 0 9.0 on o cmflk'o LE Residue Residue 160 protons of Thr114, whose chemical shifts were close to the water signal, were not observed in the two sets of the 3D spectra due to water presaturation. They were assigned by the analysis of the 2D homonuclear spectra recorded in D20 helped by the comparison with their assignments in wild-type holo-CRABPII. Aan and Thr57 were the only two residues not observed in the 2D 1H—‘SN HSQC spectrum. Their aliphatic protons were assigned by spin system identification after the sequential assignments. The side-chain assignments were complete except for Argl 1, Leu22, Arg29, Ile31, Arg59 and eight lysine residues. Stereospecific assignments were made for the B-methylene protons of 51 residues and the methyl protons of ten valine and six leucine residues. Secondary structure determination. Regular secondary structure elements were characterized by their distinct NOE patterns. The following ten strands forming two anti- parallel B-sheets could be identified (Kleywegt et al., 1994) (Figure 4. 4): Ser4—Glu13 (BA), Pro39—Glu46 (BB), Thr49-Thr56 (BC), Thr60—Lys66 (BD), Glu69—Thr75 (BE), ProSO—Glu88 (BF), Lys92—Lys98 (BG), K106—L113 (BH), E118—A125 (BI) and V128— R136 (BJ). The second strand BB began at Pro39, one residue shorter than the corresponding BB in the crystal structure of holo-CRABPII, where it began at Lys38. Lys38 was very weak in the 2D lH-‘SN HSQC spectrum. Long-range NOEs were observed between the backbone protons of all adjacent B-strands except between BD and BE. There were also NOEs between the backbone protons of the N-terrninal half of BA and the C-terminal half of BJ. The first two strands BA and BB were connected by two anti-parallel (it-helices. The first (01A), PhelS—Leu22, showed NOE connectivity and hydrogen bonding patterns typical of a regular a-helical structure. The N-terminal half 161 Figure 4.4. Summary of the sequential and medium-range NOEs involving backbone HN and H“ atoms, slow-exchange backbone amide protons and the deduced secondary structures of apo-R1 1 1M. Line thickness for daN and dNN sequential NOE distances reflects the intensities of the cross peaks. Asterisks indicate the residues whose amide protons remained afier 24 h of exchange with D20 in PBS buffer at room temperature and pD 7.5 as determined by 2D homonuclear NMR experiments. 162 10 20 30 l | 1 PNFSGNWKIIFlSENFEELLKVLGVNVMLRKIAVAAA Slow NH exchange " * " " ‘ " " daN(i,i+1) -—_-— — ‘— dNN(i,i+l) I _ --_-— daN(i,i+2) — _ L———_— __ — .— dNN(i,i+2) fi=——=__ ”‘3— '— daN(i,i+3) a 1 dNN(i,i+3) Secondary structures 1 1 . r L133: Ljfi’ . (1A “3 fl 4'0 5'0 60 7 I SKPAVEIKQEGDTFYIKTSTTVRTTEINFKVGEEFEE SlowNH exchange ‘ * " ‘ ' "I t t t t :- 9 t e daN(i,i+1) “ .— dNN(1,1+1) - - 4 — — _ — daN(i,i+2) dNN(',i+3) _ 7 7 "1° QTVDG RPCKSLVKWESENKMVCEQKLLKGEGPKT Slow NH exchange “ ‘ " ‘ ' ‘ ' * ' * d01N(i,i+1) - -— u - dNN(iri+l) - * g i L - daN(i,i+2) — — dNN(i,i+2) —— 1— __ dNN (1.1+3) ‘— BF "_ [3G Secondarystrucnnescrp ............ p ”1.1.1Lr, a: 1:0 120 130 SWTMELTN DGELILTMTADDVVCTHVYVRE 510w NH exchange a e e e t e e e e e e e e t e e e d111N(i,i+1) - -_ dNN(1,1+1) — — — — — _ daN(i,i+2) — dNN(i,i+2) ——- —— dNN (1.1+3) pH pl — 3.1 Secondary structures cm 1:1:r::r==::> u xxxxxxx ¢ 163 (Leu28—Ile31) of the second helix appeared to be partially unwound. The strength and number of the NOEs observed for residues Leu28-Ile31were well below what is expected for a regular a-helix. However, the chemical shifts of these residues indicated that they are predominately in a helical conformation. No slow exchanging amide protons were observed for the residues in the second helix. The loop (Gly23—Va126) between the two helices could be detected but the link (Ser37—Pro39) connecting dB and BB appeared to undergo conformational exchange in solution. Structural restraints. A total of 2302 structurally useful distance restraints were obtained by the analysis of the homonuclear 2D NOESY and heteronuclear 3D lH-‘sN NOESY-HMQC data. Of the 2302 distance restraints, 704 were from long-range NOEs (Table 4.2). The distributions of the distance restraints along the amino acid sequence were quite uniform except for the two N-terminal residues, the residues in turns or loops and in the second a-helix (Figure 4.5). The number of NOE restraints for each of them was well below the average. A total of 77 dihedral angle restraints were obtained from the 2D 'H-‘SN HMQC-J data. A total of 98 hydrogen bond restraints from 49 hydrogen bonds were included in the structural refinement. Quality of the structures. Figure 4.6A shows the superposition of 28 refined structures with no NOE violation exceeding 0.25 A and no dihedral angle violation larger than 5.0° The statistics of the NMR structures is summarized in Table 4.2. The stereochemical qualities of the structures were checked by the program PROCHECK— NMR as described in Chapter 3. The Ramachandran plots and overall G-factors showed that the stereochemical qualities of the structures were comparable to those X-ray structures with 2.5 A resolution and R-factors less than 20 % (Table 4.2). 164 Table 4.2. Restraint and structural statistics of apo-R1 1 1M Restraint Statistics Numbers of experimental NOE restraints lntra Sequential Medium Long Total Number of experimental ¢ dihedral restraints Number of hydrogen bonds Structural Statistics RMSD from experimental distance restraints (A) (2302) RMSD from experimental dihedral restraints (deg) (77) Deviation fiom idealized covalent geometry Bonds (A) Angles (deg) Impropers (deg) 711 Measures of Structural Quality (By Procheck) % residues in most favorable regions of the Ramachandran plot % residues in additional allowed regions of the Ramachandran plot No. of bad contacts H-bond energy Overall G-factor Coordinate Precision RMSD for Ca trace (A) RMSD for backbone atoms (A) RMSD for all heavy atoms (A) 635 252 704 2302 77 98 al )" 0.032 i 0.001 0.024 0.926 t 0.102 0.836 0.004 i 0.001 0.003 0.621 d: 0.028 0.520 0.495 i 0.004 0.411 71.8 i 2.4 71.0 25.8 i 0.8 25.8 6 i 2 6 0.40 i 0.10 0.50 -0. 19 :h 0.01 -0.20 0.43 d: 0.14 0.54 :t 0.26 0.98 :t 0.23 “ are the final 28 simulated annealing structures. b r is the restrained minimized mean structure obtained by restrained regularization of the mean structure, which is obtained by averaging the coordinates of the individual SA structures best fitted to each other. 165 Figure 4.5. The distribution of NOEs along the amino acid sequence of CRABPII. For intra-residue NOEs only those that are structurally useful are included. No. of NOEs No. of NOEs 166 8 a 0 Long Medium Sequential lntraresidue 1 * 1 1111 1111111 1 PN‘SGNWKIIRSENFEELLKVLGVNVMLRKIAVAAASKPAVEIKOEGDTFVIKYSTYVRYTEINFKVOE 5 10 15 2o 25 30 35 4o 45 so 55 so 65 Residue Number .1 " i a I 2 a. a ,1: .. 11; - u ‘ I 3;! mar oilinixsluil 1u1111.11111511111..-.1h 1.1115.--1131.h!-.1111271111 EFEEDTVDGRPCKSLVKWESENKMVCEOKLLKGEGPKTSWTMELYNDGELILTMYADDVVCYRVVVRE 70 75 80 85 9O 95 100 105 110 115 120 125 130 135 Residue Number 167 Figure 4.6. (A) Stereoview of the C“ traces of the superimposed 28 refined solution structures of apo-R11 1M. (B) Stereoview of the superimposed C“ traces of the restrained minimized mean structures of apo-R1 1 1M (thin line) and apo-CRABPII (thick line). (C) Stereoview of the C“ traces of the restrained minimized mean structure of apo-R1 11M (thin line) superimposed with the C“ traces of the crystal structure of holo-CRABPII (thick line). 168 169 Description of R] I 1M solution structure. The tertiary structure of apo-R1 1 1M was very similar to the crystal structure of CRABPs (Figure 4.6C) (Kleywegt et al., 1994; Thompson et al., 1995). The protein formed a compact, single domain structure consisting of two antiparallel B-sheets and two short a-helices. Both B-sheets consisted of five B-strands and were twisted into a nearly orthogonal flattened barrel. Inside the protein a binding pocket was formed by the side-chains from eight of the ten B-strands and lined by both hydrophobic and hydrophilic residues. It was closed off from the exteriors except for a small tunnel-like opening. The entrance was formed mainly by the hydrophobic side-chains from one side of helix 01B, the BC-BD loop and the BE-BF loop, as well as Leu19 and Va124. The loop connecting (1A and (18 extended into the solvent. In solution the ligand entrance was slightly more flexible than the other parts of the protein, but it was still relatively well defined in the solution structure. 170 Discussion Assignment. The sequential assignment of the apo-R1 1 1M posed a challenge for amide-directed sequential assignment strategy based on a series of homonuclear 2D and heteronuclear 2D and 3D 'SN-edited NMR experiments. The problem was caused by the presence of three prolines in the middle of the amino acid sequence, the low percentage of the residues whose spin systems could be identified unambiguously by homonuclear 2D NMR experiments, and the relatively low intensities of a dozen cross peaks in the 2D lH-‘SN HSQC spectrum. The intensities of these weak peaks were even lower in the 3D lH-‘SN NOESY-HMQC spectrum and the lowest in the 3D lH-‘sN TOCSY-HMQC spectrum. Moreover, eight residues had H“ protons signals and one residue had an HB proton signal quite close to the solvent resonance. These protons were not observed in the two sets of 3D lSN-edited data. Consequently, their sequential assignments could not be achieved by the 3D 'H-‘SN NOESY-HMQC experiment. Most of these difficulties were overcome by selective 15N-labeling. Lysine, leucine, isoleucine and phenylalanine were chosen because of the abundance of these residues in the protein, and the difficulty in identifying them by the standard procedure (Wiithrich, 1986). After their identifications by selective 'SN-labeling, the spin systems of over sixty percent of the total residues were known, which made it possible to achieve the total sequential assignment. Comparison of the solution structures of apo-R11 1M and wild-type apo- CRABPII. The backbone fold of apo-R11 1M is quite similar to that of apo-CRABPII. 171 The restrained minimized mean NMR structures of the two proteins can be superimposed with a RMSD of 3.2 A for all backbone C“ atoms (Figure 4.6B and Figure 4.7C), a RMSD of 2.3 A for the backbone C“ atoms of the ten B-strands and the first or-helix. However, there are significant conformational differences between the two structures, mainly localized to three segments (Leu19-Ala36, Glu73-Cys81 and Leu99-Pr0105) clustered around the ligand entrance. Surprisingly, the conformational differences occur in the region far away from the site of the point mutation. If these segments are excluded, the C“ RMSD between the two structures is reduced to 1.6 A. The first segment (Leu19- Ala36) encompasses the middle of the first helix to the end of the second helix, the second segment (Glu73-Cys81) the end of BE and the beginning of the BF, and the third segment (Leu99-Pr0105) the long BG—-—BH loop. In apo-R1 1 1M, all three segments move towards the center of the ligand entrance so that the opening of the ligand binding pocket is much smaller than that in wild-type apo-CRABPII. The binding pocket in apo-R1 11M is much less exposed to the external solvent than in the wild-type structure. For example, the distance between the C“ atoms of V3158 in the BC—BD loop and Va124 in the turn connecting two helices is 7.4 A shorter in apo-R1 11M than in wild-type apo-CRABPII. The distance between the C“ atoms of Va158 and Asp77 in the BE—BF loop is 10.8 A shorter in apo-R1 11M than in wild-type apo-CRABPII. Comparison of the solution structures of apo-RI I 1M and the crystal structure of wild-type halo-CRABPII. The restrained minimized mean NMR structure of apo-R1 11M and the crystal structure of wild-type holo-CRABPII could be superimposed very well (Figure 4.6C). The RMSD between the two superimposed structures is 0.96 A for the Cu atoms of all regular secondary structures and 1.38 A for all C,l atoms. The 172 Figure 4.7. C“ atom deviations between apo-R11 1M and wild-type holo-CRABPII (A) and between apo-R1 11M and wild-type apo-CRABPII (B). C, Deviations (A) C,ll Deviations (A) 173 > 14- j .3 N l A O 14 A 030 .I.J A J A ' . 1 . 1.1 1' .. ...111 11. 1.1.1111 11111111 1111 11111.111‘1. 1 10 20 30 40 50 60 70 80 90 100 110 120 130 N o 0— 14. 3 12- A O l n A 1 1111 .111 1 1 11 1 80 90 100 110 120 130 2 011 1111 .111.111 . 50 60 70 0 10 20 30 40 Residue Number 174 superposition also shows good alignments of the side-chains that are involved in binding of RA such as those of Arg132 and Tyrl34. It is obvious from examining Figure 4.7 that the solution structure of apo-R1 11M is more similar to the crystal structure of holo-CRABPII than to the solution structure of apo-CRABPII. The ligand entrance of apo-R1 1 1M is much smaller than that of the apo- CRABPII, but is still larger than that of holo-CRABPII. The distance between the C“ atoms of Va158 in the BC—BD loop and Va124 in the turn connecting two helices is 15.0, 21.3 and 28.8 A in holo-CRABPII, apo-RlllM and apo-CRABPII, respectively. The distance between the C“ atoms of Va158 and Asp77 in the BE—BF loop is 12.4, 13.4 and 24.4 A in holo-CRABPII, apo-R1 1 1M and apo-CRABPII, respectively. It appears that the point mutation and the binding of RA cause similar conformational changes, although the magnitudes of the changes induced by the mutation are not as large as those caused by the binding of RA. Dynamical properties of apo-R11 1M. Apo-Rl 1 1M was quite stable in solution. The NMR sample lasted for several months with almost no changes in the spectral properties and quality. More than 50 % backbone amide protons in the first a-helix and two B-sheets remained after a day of exchange with D20 (Figure 4.8). The hydroxyl protons of Thr107 and Ser83 were observed in the homonuclear 2D spectra recorded in H20, and the indole 8‘ proton of Trp7 was observed in the homonuclear 2D spectra recorded in DZO, indicating that these protons had very low solvent accessibility. All these observation suggested that R1 11M is for the most part fairly stable and rigid. The peak intensity of a 2D lH-‘SN HSQC spectrum is, in general, modulated by both amide proton exchange and conformational exchange. Amide proton exchange 175 Figure 4.8. The fingerprint regions of the homonuclear 2D TOCSY spectra of apo- CRABPII (A) and apo-R1 11M (B) recorded with 40 ms mixing time. The TOCSY experiment was initiated more than 24 h after dissolving a lyophilized protein sample in PBS, pD 7.5, in 13,0. 176 .23... .75... 3 N... . 3. ca . to _ 2. 2: 3 S 3. 2. ed a... 2: 15.3.. _...... 1.. :..::..:;..:_::r_:_..:_::_:_35:37.:_.. . _..f-.._1.::1r-P.r.E TREE _ _ _ . :FELE. ._. :3. .. __ ._ FL. .. __E._:._ H. BEFFPIF. .59: _ 2:3 SE .1 z .33. O 0255; a”? O 29; “New . 3:32.. :3. 32> . Es. o o _ c m . . ° New N3. . a...— men a n1. . . 93> 3:2 .0 o 23 n: 28 3. m I . S. O 82.5. :5. 2:. . _ 2%.”6 98.295 2:. .r 2 _ _ _ E m _ 85 OR? 02: _ Sm 3. 22¢ on H . mo...— °m~m . m2 mm; C cu: m. U a on“. o: 3.x :2 . ..,. 3. . o 0 . 2m 0. .. _ 03.... “we a; 2.3 an“. o: .3. . . N2 93 o o T . . X> . 0 “8 Q m f. . £52; 0 __ m 0 . 2.x 3> . 3. m £5 0 o w m o aim; 352:. W13. . r1. " o .W 1 1 -1- 1- w -3. m < 177 appeared to have little effects on the peak intensities of the HSQC spectra of apo-R1 11M and apo- CRABPII (Chapter 2), partly owing to the inclusion of a “water flip-back” pulse (Grzesiek & Bax, 1993) in the HSQC experiments that minimized water saturation and dephasing. In fact, the average peak intensity of the slow exchanging amide protons was the same as that of all amide protons. Thus, variation in peak intensity is primarily due to conformational exchange. As expected, the residues in the loops such as Gly47 and Asnl 15 had peak intensities lower than the average (Figure 4.1). In general, the turns or loops were more mobile than the regular secondary structures. Of the all secondary structural elements only helix aB, especially its N-terminal half Leu28—Ile31, were more flexible than the others. All the amide protons in this helix exchanged completely with the solvent in less than 12 hours. By contrast, most of the amide protons of the first helix remained after more than two days of exchange with D20. The peaks of Leu28, Arg29, Lys30 and Ile3l were much weaker than the average (Figure 4.1 and 4.7A). Va126 and Lys38 had very weak peaks, and Thr57 was not observed in 2D lH-‘sN HSQC spectrum. Furthermore, the number of NOEs observed for these residues was very low. Taken together, these results suggest that conformational exchanges are occurring in the aA-aB turn, the aB-BB link and the BC-BD loop. In other words, the ligand entrance was slightly more mobile than most of the other parts of the molecule. Did mutation change the dynamical properties of CRABPII? For the most part, apo-R1 11M and wild-type apo-CRABPII had similar dynamic properties. Thus, both proteins had about the same number of slow-exchanging backbone amide proteins. Furthermore, the identities of most of the amide protons were the same for the two proteins (Figure 4.7). However, apo-CRABPII appeared to be more mobile than apo- 178 R1 11M. The spectral qualities of the apo-CRABPII deteriorated in about two weeks as opposed to several months for apo-R11 1M. For most peaks, the wild-type protein had slightly larger line width than the mutant so that the DQF-COSY and TOCSY experiments did not work as efficiently for the wild-type as for the mutant. To compare the dynamic properties of the two proteins in more detail, we examined the intensities of the peaks in their 2D lH-‘SN HSQC spectra. The intensities of the 2D 'H-‘SN HSQC spectrum of the mutant were quite uniform (Figure 4.7A). As mentioned earlier, of all secondary structure elements, only three residues (Arg29-Ile3l) at the flexible N-terminal half of the second helix showed weak peaks. The rest of the weak peaks invariably belonged to the residues in the loops or turns. However, the wild- type protein displayed large variations in peak intensities (Chapter 3). Most residues in the RA-binding pocket and near the ligand entrance showed weak or no peaks in the 2D 'H-‘SN HSQC spectrum, indicating that the RA-binding pocket, especially the ligand entrance, is rather dynamic. Most of these peaks became much stronger afler the mutation (Figure 4.9C). In fact, most of the residues with significant intensity differences between the two spectra were clustered in these regions. Of the 21 residues that constitute the RA- binding pocket, all but three (Ile3l, Thr54 and Met123) showed significant increases in peak intensities after the point mutation. In particular, the mutation stabilized the B-bulge structure composed of residues Ile9-Arg11 and Thr13l-V133. This may be due to the attenuation of the flexibility of its neighboring residues Leu28-Ala36. The same phenomenon was observed for the residues Ser12-Glul7, which may be caused by the attenuation of the fluctuations of the two helices (01A and 013). The residues Lys8, Ala40, 179 Figure 4.9. Relative peak intensities of the 2D IH—‘sN HSQC spectra of apo-R1 1 IM (A) and wild-type apo-CRABPII (B) as a function of the residue number. (C) was obtained by subtracting (B) from (A). 180 80- C: 60. 40¢ 20- 0 20- 40. 60. 80- 14“ v 120- 1 . . . o o 0 0 o 8 6 4 2 100. £22... 252$. 140 120- l\ 100- 80- 60- 40- 20- O- 70 80 90 100 110 120 130 60 20 30 40 10 Residue Number 181 LysS3, Glu62, Ile63, Leull9 and Vall35 also became much stronger after the point mutation. Their implications will be discussed later. Binding of RA substantially stabilized CRABPI as measured by thermal denaturation. In particular, it stabilized the flexible parts of the ligand binding pocket in CRABPI as revealed by NMR (Rizo et al., 1994). Our NMR studies showed that the changes in dynamical properties caused by the point mutation were rather similar to those induced by RA binding (Chapter 5). Thus, the dynamic properties of apo-R1 11M were more similar to those of holo-CRABPI and holo-CRABPII than to apo-CRABPI and apo- CRABPII. Binding of RA and the mutation caused similar changes in the dynamical properties of CRABPII. Roles of Arg] 1 1 . Argl 11 interacts with the carboxyl group of RA via a hydrogen bond mediated by a water molecule. In Chapter 2 I have shown that Argl 11 contributes to the binding of RA by ~2.2 kcal/mol, indicating that Argl 11 is indeed important for the binding of RA. X-ray crystallographic and molecular modeling studies suggest that Argl 11 is one of the four arginine residues that generate a positive potential that may facilitate the binding of RA (Zhang et al., 1992). The NMR results presented in this chapter suggest an additional role for Argl 11 related to the structure and dynamics of CRABPII. As mentioned earlier, RA binding induces significant but localized conformational and dynamical changes in both CRABPI and CRABPII. The ligand entrance, in particular, is much more open and flexible in the apo-forms than in the holo- forms. The large opening of the ligand entrance and its high flexibility are probably essential for the unhindered entry of RA into the deep binding pocket and the subsequent . 182 closure of the ligand entrance. Our NMR results show that the point mutation also causes significant but localized changes in the conformation and dynamics of CRABPII. Strikingly, apo-RlllM is more similar to wild-type holo-CRABPII than to wild-type apo-CRABPII in both structure and dynamics, suggesting that Argl 11 may play a major role in maintaining the ligand entrance in an open and dynamic state for the entry of RA. The next question is how Argl ll affects the internal motions of CRABPII. At first thought, the interactions leading to the high flexibility of apo-CRABPII was the electrostatic repulsion between the positively charged guanidinium groups of two conserved arginines, Arglll and Arg132 in CRABPII and Arglll and Argl31 in CRABPI (Zhang et al., 1992). However, the two guanidinium groups in apo-CRABPII were ~7.6 A apart. Furthermore, if the electrostatic repulsion played a critical role substitution of Argl32 with methionine would most probably change the dynamical properties to the same extent. However, this was not supported by our NMR characterization of apo-R132M. The qualities of the homonuclear 2D spectra of apo- R132M were slightly better than those of the wild-type but worse than those of the apo- R1 1 1M, suggesting that apo-R132M is less flexible than apo-CRABPII but not as rigid as apo-R1 1 1M. For most peaks apo-R1 11M had the smallest line widths among all mutants that we generated (V24A, R59G, L121A, M123A and R132M). Thus the electrostatic repulsion between the two arginines is, at most, only partially responsible for the conformational flexibility of apo-CRABP. Other interactions contributing to the dynamical properties of apo-CRABPII were suggested by the solution structures of apo-R11 1M and apo-CRABPII (Chapter 3). In both structures, the guanidinium group of the Arg132 is partially exposed and surrounded 183 by the side-chains of Ser12, Pro39, Thr54 (y-methyl group), Cysl30 and Tyrl34. The hydroxyl groups of Ser12 and Tyrl34 and the thiol group of Cys130 provide the favorable polar interactions with the guanidinium group of the Argl32. In particular, the hydroxyl group of the phenolic ring of Tyrl34 is within the hydrogen bond distance to the Ne nitrogen of the guanidinium group of Arg132. By contrast, in the solution structure of apo-CRABPII, the guanidinium group of Argl 11 is almost completely buried and in an environment much more hydrophobic than that of Arg132. It is surrounded by the side-chains of Val4l, lle43, Ile52, Thr54, Ile63, Trp109, Leul l9 and Leu121. The hydrophobic side-chains of Val4l, Ile63, Leul l9 and Leu121 had close contacts with the guanidinium group of Argl l l. The only favorable interactions may be between the guanidinium group of Arglll and the aromatic ring of Trp109 and between the guanidinium group of Argl 11 and the hydroxyl group of Thr54. As mentioned earlier, the peaks of the residues Val41, LysS3, Glu62, Ile63, Leull9 and Leu121 became much stronger in the 2D lH-‘SN HSQC spectrum and/or 2D homonuclear spectra after the point mutation. It is tempting to speculate that the charged guanidinium group of Argl 11 is mobile because of its hydrophobic environment and possibly the repulsive interaction with Argl32. The mobility of the guanidinium group of Arglll drives the conformational flexibility of apo-CRABPII and results in a more open conformation for the ligand entrance. The hydrophobic side-chain of Metl 11 is preferred by the hydrophobic environment. Binding of RA neutralizes, at least partially, the positive charge of Argl 11 and therefore reduces the mobility of its guanidinium group. CRABPs are most stable in the holo-forms because there are additional hydrophobic interactions between the proteins and RA. It is noted that the guanidinium group of Argl 11 is more 184 than 17 A away from the ligand entrance, the most flexible region of the apo-CRABPII. It is not clear how the mobility of the guanidinium group propagates to the ligand entrance. 185 Conclusion The solution structure of apo-R1 1 1M, a site-directed variant of human CRABPII, has been determined by multidimensional heteronuclear NMR. The solution structure of apo-R1 1 1M was more similar to the crystal structure of wild-type holo-CRABPII than to the solution structure of wild-type apo-CRABPII. The size of the ligand entrance, the mobility of the RA-binding pocket and the ligand entrance were reduced greatly after the point mutation Argl 11 to Metl 11. The results suggest that Argl 11 plays a major role in maintaining the ligand entrance in an open and dynamic state for the entry of RA. 186 References Anil-Kumar, Ernst, R. R., & Wiithrich, K. (1980) Biochem. Biophys. Res. Commun. 95, 1-6. Basus, V. J. (1989) Methods Emzymol. 1 7 7, 132-149. Bax, A., Griffey, R. H., & Hawkins, B. L. (1983) J. Magn. Reson. 55, 301-315. Bodenhausen, G., & Ruben, D. J. (1980) Chem. Phys. Lett. 69, 185-188. Cavanagh, J., Palmer A. G., Wright, P. E., & Rance, M. (1991).]. Magn. Reson. 91, 429- 436. Charnbon, P., Olson, J. A., & Ross, A. C. (coordinators) (1996) The retinoid revolution, FASEB J. 10, 938-1048. Delaglio, F ., Grzesiek, S., Viiister, G. W., Zhu, G., Pfeifer, J ., & Bax, A. (1995) J. Biomol. NMR 6, 277-293. Garret, D. S., Powers, R., Gronenbom, A. M., & Clore, G. M. (1991) J. Magn. Reson. 95, 214-220. Clore, G. M., & Gronenbom, A. G. (1989) CRC Crit. Rev. Biochem. Mol. Biol. 24, 479- 564. Clore, G. M., & Gronenbom, A. G. (1991) Annu. Rev. Biophys. Chem. 21 , 29-63. Griesinger, C., Sorensen, O. W., & Ernst, R. P. (1985) J. Am. Chem. Soc. 107, 6394- 6396. Griesinger, C., Otting, G., Wiithrich, K., & Ernst, R. P. (1988).]. Am. Chem. Soc. 110, 7870-7872. 187 Grzesiek, S., & Bax, A. (1993) J. Am. Chem. Soc. 115, 12593-12594. Hodson, M. E., & Cistola, D. P. (1997) Biochemistry 36, 2278-2290. Jeener, J., Meier, B. H., & Ernst, R. R. (1978).]. Chem. Phys. 71, 4546-4553. Kay, L. E., & Bax, A. (1990) J. Magn. Reson. 86, 120-126. Kay, L. E., Keiffer, P., & Saarinen, T. (1992).]. Am. Chem. Soc. 114, 10663-10665. Kleywegt, G. J., Bergfors, T., Senn, H., 1e Motte, P., Gsell, B., Shuao, K., & Jones, T. A. (1994) Structure 2, 1241-1258. Marion, D., Ikura, M., Tschudin, R., & Bax, A. (1989a) J. Magn. Reson. 85, 393-399. Marion, D., Driscoll, P. C., Kay, L. E., Wingfeld, P. T., Bax, A., Gronenbom, A. M., & Clore, G. M. (1989b) Biochemistry 28, 6150-6156. Marion, D., Kay, L. E., Sparks, S. W., Torchia, D. A., Bax, A. (1989c) J. Am. Chem. Soc.111,1515-1517. Mueller, L. (1979) J. Am. Chem. Soc. 101, 4481-4484. Palmer, A. G., Cavanagh, J ., Wright, P. E., & Rance, M. (1991) J. Magn. Reson. 93, 151- 170. Piantini, U., Sorensen, O. W., & Ernst, R. P. (1982) J. Am. Chem. Soc. 104, 6800-6801. Rizo, J ., Liu, Z. P., & Gierasch, L. M. (1994) J. Biomol. NMR 4, 741-760. Shaka, A. J ., Lee, C. J ., & Pine, A. (1988) J. Magn. Reson. 77, 274—293. Shaka, A. J ., Barker, P. B., & Freeman, R. (1985) J. Magn. Reson. 64, 547-552. Spom, M. B., Roberts, A. B. & Goodman, D. S. (1994) The Retinoids: Biology, Chemistry, and Medicine, 2"“ ed. Raven Press, New York. States, D. J ., Haberkon, R. A., & Ruben, D. J. (1982).]. Magn. Reson. 48, 286-292. Wiithrich, K. (1986) NMR of Proteins and Nuclear Acids, Wiley, New York. 188 Wiithrich, K., Billletter, M., & Braiin, W. (1983).]. Mol. Biol. 169, 949-961. Zhang, J ., Liu, Z., Jones, T. A., Gierasch, L. M., & Sambrook, J. F. (1992) Proteins: Structure, Function, and Genetics 13, 87-99. CHAPTER 5 NMR Studies of Human CRABPII in Complex with All-trans-Retinoic Acid Introduction The crystal structures of holo-CRABPs suggest that RA cannot enter or exit the deep binding pocket in the absence of major conformational changes in the protein (Kleywegt et al., 1994). Indeed, significant conformational differences between the crystal structure of holo-CRABPII and the solution structure of apo-CRABPII have been observed, especially in the ligand entrance region (Chapter 3). In comparison to the holo crystal structure, apo-CRABPII exhibits a concerted movement of the second helix, the BC-BD loop and the BE-BF loop, so that the ligand entrance of apo-CRABPII is greatly enlarged and readily accessible to RA. Furthermore, the ligand-binding pocket of apo- CRABPII is rather disordered. The results suggest that binding of RA induce significant changes in both conformation and dynamics of CRABPII. This chapter describes the sequential resonance assignments and preliminary 189 190 results showed that RA binds to CRABPII in solution in the same manner as determined by crystallography and the ligand entrance of holo-CRABPII is much less flexible than that of apo-CRABPII. Relative motions between the bound RA and human CRABPII in solution, particularly at the ligand entrance, were detected. Such motions may provide a transient pathway for the exit of RA. 191 Experimental Procedures Sample preparation. All-trans-RA and D20 were purchased from Sigma (St. Louis, Missouri), dimethyl sulfoxide-d, (DMSO-d6 ) were purchased from Aldrich. Expression of CRABPII gene and purification of the recombinant protein have been described in Chapter 2. The NMR samples of holo-CRABPII were prepared as follows. CRABPII was first dissolved in ~50 mi PBS buffer (20 mM sodium phosphate, 150 mM sodium chloride with 99.6% D20, pD 7 .5, uncorrected). The protein concentration was ~40 uM. All-trans-RA was dissolved in DMSO-d6. The concentration of the RA solution was ~3 mM. The RA solution was then added to the protein solution by aliquots of ~5 11.1. The solution was quickly mixed by stirring after each addition. The entire procedure was carried out in a dark room. The final DMSO concentration was less than 1.5%. The sample was concentrated to ~06 ml. H20 sample of uniformly l“N-labeled holo-CRABPII was prepared similarly except that the D20 in PBS buffer was replaced by HZO. The final protein concentrations of the NMR samples were ~2 mM. NMR spectroscopy. Homonuclear 2D spectra of the unlabeled DZO sample and heteronuclear 2D and 3D spectra of the uniformly '“N-labeled HZO sample were acquired and processed as detailed in Chapter 3 and 4. The temperature was set at 25 “C. Homonuclear 2D spectra were acquired with a spectral width of 8000 Hz in both dimensions. The carrier frequency was 4.70 ppm (solvent resonance). The relaxation delay was 1.2 s. The residual water resonance was suppressed by low-power 192 presaturation. One DQF-COSY spectrum, three clean-TOCSY spectra with mixing times of 20, 33 and 50 ms, and two NOESY spectra with mixing times of 100 and 150 ms were acquired on the D20 sample. The time domain data were composed of 2048 X 320 complex points for the DQF-COSY (128 transients) and NOESY (64 transients) experiments and 2048 X 256 complex points for the TOCSY experiments (64 transients). The carrier frequencies were 4.70 ppm for 1H and 121.0 ppm for 15N for the heteronuclear experiments. A 2D lH-‘SN HSQC spectrum was acquired with the following acquisition times and complex data points: lH (F2) 121.9 ms, 1024 and 15N (F 1) 53.3 ms, 128 (32 transients). A 3D 1H-‘SN NOESY-HSQC spectrum (16 transients, 150 ms mixing time) and a 3D lH-‘SN TOCSY-HSQC spectrum (8 transients, 35 ms mixing time) were recorded with the same acquisition times and complex data points: IH (F3) 121.9 ms, 1024, lsN(132) 15.2 ms, 32, 1H (Fl) 15.2 ms, 128. 193 Results and Discussion. Protein ’ H and ”N resonance assignments and secondary structures. The protein 'H and 15N resonances (Table 5.1) were assigned by the same strategy as detailed previously (Chapter 4). Briefly, spin system identification was achieved by the homonuclear 2D and 3D l5N-edited TOCSY experiments. The sequential linking was accomplished by the analyses of the 2D and 3D NOESY spectra. The assignment was helped greatly by comparison with the sequential assignments of the wild type apo- CRABPII and the mutant apo-R1 1 1M (Chapters 3 and 4). Secondary structural elements were identified as described in Chapter 4. They were essentially the same as those observed in the crystal structure except that the first helix 01A appeared to be mobile in solution (Figure 5.1 and 5.2). Conformational and dynamical changes in CRABPII induced by the binding of RA. The majority of the NOEs observed between the bound RA and the protein were in agreement, at least qualitatively, with those predicated from the crystal structure of holo-CRABPII, suggesting that the tertiary structure of holo-CRABPII in solution is most likely very similar to that in the crystalline state. In particular, the contacts between the bound RA and the residues at the ligand entrance observed by crystallography were also detected by NOEs, indicating that the conformation of the ligand entrance of holo- CRABPII in solution is similar to that in the crystalline state. These NMR results further strengthened the conclusion based on the comparison of the holo crystal structure and the 194 Table 5.1. 1H and 15N chemical shifis of human CRABPII in complex with all-trans-RA in PBS at 25°C. No P1 N2 F3 S4 GS N6 W7 K8 19 110 R1 1 812 E13 N 14 F15 E16 E17 L18 L19 K20 V21 L22 G23 V24 N25 V26 M27 L28 R29 K30 131 A32 V33 A34 A35 A36 S37 K38 P39 A40 V41 E42 I43 K44 Q45 lSN 121.9 112.4 110.5 118.5 122.9 123.3 125.8 118.0 120.8 120.5 121.2 113.8 118.9 121.3 117.5 106.6 121.4 128.2 120.5 119.8 117.8 118.8 116.2 118.9 120.5 116.5 120.4 121.2 116.6 113.5 119.0 123.0 124.5 130.0 122.9 128.8 127.9 HN 9.86 8.48 9.22 8.19 9.31 10.17 8.60 9.24 7.49 8.52 7.59 7.25 8.24 7.84 7.15 7.68 7.60 8.80 8.57 7.94 7.80 8.81 7.63 7.00 8.56 8.85 7.19 8.18 8.69 7.38 7.05 8.77 7.95 9.29 9.29 9.57 8.60 HQ 4.35 5.08 4.35 4.70 4.04, 3.82 5.69 5.21 5.26 3.99 4.62 4.58 5.07 4.83 4.99 4.46 3.30 3.74 3.74 3.60 3.83 3.77 3.96 3.96, 3.61 3.74 4.42 3.49 4.13 4.09 3.58 3.90 3.75 3.88 3.50 4.15 4.24 3.96 4.22 4.80 5.09 5.03 5.22 5.35 5.38 4.65 4.51 H“ 2.39, 1.92 2.98, 2.90 3.34, 3.28 4.26, 3.99 2.72, 2.55 3.23, 2.99 1.98, 1.88 1.91 1.70, 1.70 3.59, 3.33 1.94, 1.94 3.35, 3.16 2.99, 2.72 1.85, 1.85 1.87, 2.16 1.86, 1.69 1.66, 1.60 1.93, 1.63 2.15 1.84, 1.58 1.47 3.17, 2.64 2.07 2.13, 1.98 1.80, 1.64 1.95 1.91, 1.91 1.91 1.18 2.07 1.47 1.51 0.86 3.98 1.90, 1.61 1.92, 1.87 1.37 1.88 2.08, 2.02 2.26 1.87, 1.76 1.84, 1.79 Others 1172.11, 2.11; H“ 3.35, 3.28 yN 112.5;yN11, 7.75, 7.02 2, 6H 7.43; 3, 5H 7.38; 4H 7.17 2H 6.80; 4H 7.07; 5H 7.26; 6H 6.81;7H 7.14 N“ 124.8; H“ 9.17 H" 1.68, 1.63; H“ 1.37, 1.33; H“ 2.87, 2.83 H""' 0.83 H" 2.09, 2.19 2, 6H 7.06; 3, 5H 6.64; 4H 6.59 H" 1.98, 2.17 H" 2.27, 2.27 H“ 0.81, 0.81 H" 1.17; H6 0.32, 0.22 H" 1.56, 1.35; H“ 2.82, 2.93 H" 1.10, 0.94 H“ 0.92, 0.70 H" 0.82, 0.82 yN 112.4;er12 7.71, 6.92 H" 0.97 H" 2.60, 2.45 H" 1.33;}1‘5 0.88, 0.77 H" 1.72 1.53; H“ 1.45, 1.45; H“ 2.90, 2.90 H"'“ 0.83 H" 1.03, 0.96 H" 1.42, 1.31; H" 2.95, 2.95 H“ 3.65, 3.55 H" 0.76, 0.44 H"2.27, 2.27 H" 1.67, 1.34; H“ 0.48; H""' 0.85 H" 1.49, 1.36 H"2.04, 2.04 195 No E46 G47 D48 T49 F50 Y51 152 K53 r54 SSS T56 T57 vss R59 T60 T61 E62 163 N64 F65 K66 V67 G68 E69 E70 F71 E72 E73 Q74 r75 v76 D77 G78 R79 P80 C81 K82 S83 L84 vss K86 W87 E88 SS9 E90 N91 K92 M93 iSN 128.7 117.5 126.5 116.6 124.1 122.5 123.3 133.4 124.7 123.1 113.5 109.5 119.0 122.7 117.2 122.9 116.8 120.2 116.8 120.8 127.3 114.8 120.1 130.3 121.3 119.4 120.3 118.1 111.7 122.4 112.3 107.0 120.7 118.8 119.8 124.7 120.5 130.0 123.6 120.1 127.7 112.4 117.1 111.0 119.9 127.0 HN 8.53 9.30 8.56 8.08 9.10 8.35 8.35 9.31 8.84 9.56 7.74 6.71 7.09 8.73 8.72 8.77 8.29 8.10 8.36 9.35 8.98 8.93 7.96 9.25 9.06 8.72 8.44 7.68 9.25 9.25 8.75 7.96 7.91 8.85 8.89 9.05 7.34 9.24 8.47 8.69 9.25 8.60 8.72 8.05 7.83 9.40 HG. 4.37 3.96, 3.61 5.21 4.66 4.63 5.33 4.62 5.14 4.79 5.25 4.93 3.98 4.48 4.76 5.32 4.64 4.85 4.93 5.59 4.70 5.30 3.24 4.42, 3.62 4.76 4.80 5.14 5.05 4.76 5.32 4.61 3.95 4.46 4.06, 3.59 4.70 4.65 5.09 4.75 4.86 4.25 4.38 4.51 5.14 4.37 4.57 3.68 4.97 5.52 5.21 H“ 1.80, 1.66 3.16, 2.72 4.15 2.57, 2.57 3.03, 2.69 1.89 1.71, 1.67 3.83 3.82, 3.77 4.33 4.18 2.18 2.02, 1.91 3.93 3.98 1.94, 1.94 1.90 2.58, 2.53 2.13, 2.13 1.65, 1.60 1.88 2.04, 1.92 2.00, 1.91 3.57, 3.18 1.96, 1.92 2.04, 1.92 1.69, 1.73 4.15 2.12 2.82, 2.13 180,156 2.62, 2.17 1.58, 1.50 2.78, 1.54 1.60, 1.30 1.37 1.64, 1.45 3.26, 3.00 1.98, 1.69 4.14, 3.99 1.87, 1.80 3.18, 2.88 2.01, 1.95 1.80, 1.80 Others H" 2.01, 2.01 H" 1.01 2, 6H 6.46; 3, 5H 6.81; 4H 7.09 2, 6H 6.93; 3, 5H 6.61 H" 1.35, 1.35; H""1 0.28; H“ 0.37 H" 1.60, 1.48; H" 1.17; H“ 2.81, 2.81 H" 0.85 H" 1.18 H" 1.32 H" 0.81, 0.81 H" 1.65, 1.49 H" 1.06 H" 0.97 H" 2.21, 2.10 H" 1.33, 0.75; H“ 0.84; H“ 1.06 yN 113.0; mi, 7.62, 6.93 2, 6H 6.20; 3, 5H 6.51; 4H 6.28 H" 1.53, 1.47; H“ 2.94, 2.94 H"0.66, 0.45 H" 2.16, 2.16 H" 2.22, 2.17 2, 6H 7.24; 3, 5H 6.83; 4H 6.47 H"2.33, 2.14 H" 2.16, 2.16 H"2.24, 1.99 H" 1.07 H" 1.18, 0.86 H" 1.27, 1.27; H“ 2.80, 2.80 H" 1.17; H“ 0.73, 0.69 H" 0.24, 0.00 H" 1.52, 1.27; H“ 1.15 2H 7.32; 4H 7.98; 5H 6.85; 6H 7.23; 7H 7.63 N“ 129.5; H“ 10.68 H" 2.16, 2.02 H" 2.09, 2.09 yN 112.6; yNH, 7.49, 6.84 H" 1.49, 1.49 H"2.07, 199; He 0.43 No V94 C95 E96 Q97 K98 L99 L100 K101 (3102 13103 6104 P105 K106 T107 8108 W109 T110 R111 E112 L113 T114 N115 D116 G117 E118 L119 1120 L121 T122 M123 T124 A125 D126 D127 V128 V129 C130 T131 R132 V133 Y134 V135 R136 E137 ISN 118.8 127.1 127.5 120.7 122.8 125.9 121.5 117.5 108.7 116.8 108.8 123.6 119.0 112.1 116.2 110.8 119.9 123.1 126.4 116.9 115.6 107.6 117.4 124.5 125.1 135.4 119.3 121.2 115.6 125.4 122.2 121.1 123.3 125.6 129.7 124.5 128.6 122.7 126.7 112.0 122.3 126.5 HN 8.80 8.87 7.67 7.41 8.96 8.84 7.74 8.03 8.27 8.17 8.16 8.58 8.80 7.80 8.26 9.58 8.81 8.91 8.23 9.38 7.74 7.98 7.42 8.92 9.51 9.98 9.15 9.28 8.27 8.40 9.26 8.49 8.41 8.42 9.59 8.45 9.85 8.56 9.79 9.52 8.44 8.10 HQ 4.30 4.51 4.35 4.93 4.59 4.02 4.10 4.42 4.06, 3.75 4.61 4.14, 3.93 4.51 4.47 5.39 4.70 5.79 5.24 5.62 5.80 5.31 4.60 4.46 4.76 4.19, 3.70 4.98 5.13 4.98 5.61 5.85 5.16 5.48 4.88 4.16 4.52 3.95 4.42 5.20 5.08 5.09 4.98 5.61 5.35 3.88 3.87 HB 1.80 0.41 —0.27 1.90, 1.80 1.77, 1.77 1.53, 1.53 1.64, 1.47 1.65, 1.48 1.80, 1.80 1.80, 1.80 2.03, 2.03 1.80, 1.80 4.49 3.93, 3.77 3.53, 3.33 3.96 1.80, 1.50 1.95, 1.95 1.28, 0. 75 4.761 2.82 2.81, 2.42 1.83, 1.72 0.87 1.82 1.90, 1.34 4.05 1.84, 1.80 3.96 1.04 2.75, 2.75 2.68, 2.68 2.14 1.84 2.73 3.76 1.75, 1.61 1.71 3.03, 3.03 2.39 1.32, 1.09 1.81 1.69 196 Others H" 0.86, 0.75 H" 2.14, 2.14 H" 2.43, 2.04 H" 1.26, 1.26; H‘ 2.79, 2.79 H" 1.38; H‘5 0.67, 0.67 H" 1.28; H’5 0.78 H" 1.53, 1.37 H" 2.13, 1.98 H" 1.85, 1.73; H‘5 3.61, 3.25 H" 1.53, 1.42 H" 1.15 2H 6.92; 4H 6.90; 5H 6.35; 6H 6.44; 7H 6.67 N‘ 138.1; H‘ 11.38 H" 1.00 H" 1.32 H"2.04, 2.04 H" 1.12, H‘5 0.46, 0.46 H" 1.26 yN 112.6; yNH2 7.65, 6.95 H" 2.04, 1.91 H" 0.46; H5 0.15, —0.25 H6 0.78, 0.71 H" 1.09 H"2.64, 2.33 H‘ 1.63 H" 1.00 H" 0.92, 0.69 H" 0.78, 0.70 11" 1.00 H" 0.86, 0.86 2, 6H 6.83; 3, 5H 6.46 H" 1.06, 0.86 H" 0.58, 0.04; H6 2.81, 2.73 H" 2.07 2.07 197 Figure 5.1. Strip plots extracted from the 150 ms mixing time 15N-edited NOESY- HSQC spectrum of the holo-form CRABPII showing the NOE connectivies for residues Met27-Lys38. 198 >16»); 3) («555 1&1 1; [All LE2 M‘ .‘ ERR». ‘4, 1 3 ~ ~ _ é" 5* =1— N‘ ., 1.11,? gt \‘1'51 m0, - saw v . i+1 i+1 i-2 f i-l 1+1 0.0 1.0 ”6.0 I I ‘T 1 T I 1 1 I I 1 1 M27 L28 R29 K30 131 A32 V33 A34 A35 A36 S37 K38 199 Figure 5.2. Summary of the sequential and medium-range NOEs involving backbone HN and Hcl atoms and the deduced secondary structures of holo-CRABPII. Line thickness for dGIN and dNN sequential NOE distances reflects the intensities of the cross peaks. 200 10 20 3o PNFSGNWKIIRSENFEELLKVLGVNVMLRKIAVAAA SbWNngo t t :1 10 o .3. t ‘ (W1) _— _ ‘—_——- daNth) __ _ -—-—__ _ (111110.02) Kai-P3) —__— ___-*1—__ 654-3) ' . — fl Wm rIIIEA1111> $ k _ 40 50 60 70 SKPAVEIKQEGDTFY‘IKTSTTVRTTEINFKVGEEFEE WWW OOO OOOO O O daN(i,i+1) - —-— ——- —1_-_-—. dNN(i,i+l) I __ I. _ ‘ - (161116.92) __ ___ __ —__ (W2) _. _ dNNG.’r*3) __ WW 0223;?» anpézzb 11111 PLOW”, czgcn 80 90 100 QTVDGRPCKSLVKWESENKMVCEQKLLKGEGPK SWNHW t 1t t a o o t . duN(i,'r1-1) — —-———-—-— - dNNCLi'tl) —_ _— —* — duN(i,i+2) -'_'-— ——-1_._ dNN(1,1+2) ___ —__ _ 911196.113) -.__ __ mm c’ ”HEP-Fm, IIIIIBGIIIII) E2: 110 '1 120 g 130 SWTHELTNDGELILTMTA‘DDVVCTRVYVRE ”mm O O O O OO OOOO O OO O O O OO O daNGJH) h.- nun—— —_- daNam) — “1010:1121 __ _ dNN(i.i+3) —— . BH 31 N mm m m fiT, _ _ _ rTLC. 201 apo solution structure that RA binding induces significant conformational changes at the ligand entrance (Chapter 3). Dynamical changes upon binding of RA were evidenced by the differences in peak intensity between the 2D lH-‘SN HSQC spectra of apo- and holo-CRABPII (Figure 5.3 and 5.4) and the differences in the number and intensity of the observed NOEs. Remarkable differences were observed in the second helix, which is part of the ligand entrance. It had a typical a-helical NOE pattern in the holo-form (Figure 5.1 and 5.2) but no such pattern was observed in the apo-form. Furthermore, the residues in the second helix had much more intense cross peaks in the 2D lH-‘SN HSQC spectrum of the holo form. The second helix is apparently disordered in the apo form (Chapter 3). Other residues at the ligand binding pocket also showed more intense cross peaks in the 2D 'H- 15N HSQC spectrum upon binding of RA, including Thr56-Ile63, Val76-Gly78, Leu121, Arg132 and Argl34. Thr56-Ile63 and Val76-Gly78 constitute the [SC-[3D loop and the BE-BF loop, respectively, at the ligand entrance. Leu121, Arg132 and Arg134 are located at the bottom of the binding pocket. Their side-chains interact with the carboxyl group of RA. The results suggest that while the ligand binding pocket of apo-CRABP undergoes conformational exchange, the conformational exchange is greatly reduced upon binding of RA. Similar ligand-induced dynamical changes have been observed in the NMR studies of CRABPI (Rizo et al., 1994) and intestinal fatty acid binding protein (I-FABP) (Hodsdon et al., 1995; Hodsdon et al., 1996; Hodsdon & Cistola, 1997a, b). The stability of the ligand entrance in holo-CRABPs may be brought about by the hydrophobic interactions between the isoprene tail of RA and the protein. However, not all HSQC cross peaks in CRABPII became stronger after the binding of RA. The cross 202 Figure 5.3. Gradient- and sensitivity-enhanced 2D lH—‘sN HSQC spectrum of uniformly l5N enriched of the wild type holo-CRABPII in complex with all-trans-retinoic acid. Sequential assignments are indicated with one-letter amino acid code and residue number. 120 ~1 "190112110 ‘ 125—— 1—<——--—- _... 11 ._.-.._..1_..-___._ .— O W87Ne . W 109Ne 203 W., -—.-——.. .— G780 & o 0117 ‘ 0102” 6104 19181 8108}. . 369331.19 T124 . E90 W109. T49 966' 1141?”. 0116 O V33 9 Tax-30.1538 m F15 0126 o w 043 El 2,140 555 9 1.1191(1 0 M93 '- N25 0 . c130 QV85 E70 Y134 0 L121 K1331 9&4 M2 6400192091214M4 G 87 °D 27EG99°~ V24 V263 M5523 3“ 0 E96 rose-.9131 .9. 11.5 ITIIIIHITTllilll 11.0 10.5 10.0 Tj11111111111111111111711111[11111] 9.5 8.5 '11 (99m) 8.0 7.5 6.5 204 Figure 5.4. Relative peak intensities of the 2D lH—-‘5N HSQC spectra of apo-CRABPII (A) and holo-CRABPII (B) as a fimction of the residue number. (C) was obtained by subtracting (B) from (A). Relative lntenslty 205 111 1 11 1 1 10 20 30 40 50 60 70 80 90 100 110 120 130 Resldue Number 206 peaks of Ser12, Glu13, Asn14, Glu16 and Glul7 could not be observed in the 2D lH-‘SN HSQC spectra. The cross peaks of IlelO, Argl l, PhelS, Leu19 were very weak. Leu18 probably was also weak. These residues became more flexible afier the binding of RA. Their high mobility compensates, at least partially, the decrease in entropy caused by ligand binding. ’H resonance assignments of CRABPII-bound RA. The 'H resonances of the bound RA were assigned by analysis of the homonuclear 2D DQF-COSY, TOCSY and NOESY spectra of holo-CRABPII recorded in D20. The olefinic protons were well resolved from the protein resonances. Through-bond correlations could be clearly identified for CH(7), CH(8), CH( 10), CH(l 1) and CH(12) protons (Figure 5.5). CH(l 1) was assigned by its distinct through-bond correlation pattern. The remaining olefinic protons were assigned based on their NOE patterns (Figure 5.6). The methyl protons of CH3(18), CH3(19) and CH3(20) were assigned based on their NOEs to olefinic protons. The assignments of the methyl protons of CH3(16) and CH3(17) were tentative. None of the methylene protons of CH2(2), CH2(3) and CH2(4) of B-ionone ring was assigned at present. The 1H chemical shifts of the RA bound to human CRABPII (Table 5.2) were almost identical to those reported for the RA bound to mouse CRABPII (Norris, et al., 1994) except for CH(12) which differed by ~0.2 ppm. The chemical shifts of the CRABPII-bound RA protons were also similar to those of free RA in CC], except for CH(8), CH(l l) and CH3(20) protons, which were 6.06 ppm, 6.96 ppm and 2.36 ppm in CC14, respectively (Waterhous & Muccio, 1990). Conformation and dynamics of bound RA. Most of the NOEs observed for the bound RA (Table 5.3, Figure 5.6 and 5.7) were consistent with the distances measured 207 Table 5.2. 'H resonance assignment of retinoic acid bound to human CRABPII in PBS at 25°C. Olefinic proton (CH) CH(7) CH(8) CH(10) CH(ll) CH(12) CH(14) Chemical shift (ppm) 6.08 6.36 6.22 6.47 6.14 5.62 Methyl protons (CH3) CH3(16)’ CH3(17)' CH3(18) CH3(19) CH3(20) Chemical shifi (ppm) 0.90 0.99 1.66 1.92 1.97 ' the individual assignments of CH3(16) and CH3(17) were assigned tentatively. 208 Table 5.3. The relative intramolecular NOE volumes 3 of RA bound to human CRABPII in PBS at 25°C (the mixing time was 100 ms). Proton pair Relative NOE volume Proton pair Relative NOE volume Proton pair Relative NOE volume Proton pair Relative NOE volume Proton pair Relative NOE volume 7-8 4.8 8-10 40.0 10-11 38.3 8-19 33.6 12-20 18.5 7-10 1.9 8-12 16.1 10-12 43.1 11-14 30.0 12-14 100.0 7-16 34.0* 8-14 10.7 10-14 36.7 11-12 9.1 14-20 strong" 7-17 40.1* 8-16 31.8* 10-18 33.2 11-19 strong“ 7-18 26.9 8-17 31.6* 10-19 30.5 11-20 strong" 7-19 45.0 8-18 45.6 10-20 6.0 12-19 12.3 a The NOEs were picked by the program NMRCOMPASS 2.5 (Molecular Simulations) from the homonuclear 2D NOESY spectrum with a mixing time of 100 ms. The relative volumes were obtained by integration using NMRCOMPASS. 'These intramolecular NOE peaks may be overlapped with intermolecular ones. " Strong peaks but no accurate values could be obtained because of overlapping. 209 Figure 5.5. Part of the 2D TOCSY spectrum with 33 ms mixing time of the RA bound to human CRABPII at 25°C. The CH(7), CH(8), CH(lO), CH(l 1), CH(12) and CH(14) protons are labeled. 210 ’1 p CH(14) V 1 1111 111:1 1? CH(10)-CH(11) Ct at 01 O O 0 fi N 0 0’1 O as 1111111111111111111111111111111111111111111111111:1 IIIIIIIIWTIIIIll[[11llllllllllllll[11'11111111111111IIIIIIIIPTIIIIIIIIITIIT 7.0 6.8 6.6 6.4 6.2 6.0 5.8 5.6 F2 (PM) 211 Figure 5.6. Part of the NOESY spectrum with 100 ms mixing time of the holo-CRABPII in complex all-trans-RA at 25°C showing its intramolecular NOEs and the intermolecular NOEs between RA and Phe15 of the protein. The identities of the NOEs are as follows. A, CH(14)-CH(11); B, CH(14)-CH(8); C, CH(14)-CH(10); D, CH(14)-CH(12); E, CH(10)-CH(12); F, CH(lO)-CH(7); G, CH(8)-CH(10); H, CH(8)-CH(12); I, CH(8)-CH(7); J, CH(1 l)-CH(12); K, CH(11)-CH(10); a, 2, 6H(F15)-CH(14); b, 4H(F15)-CH(14); c, 3, 5H(F15)-CH(14); d, 4H(F15)-CH(12); e, 3, 5H(F15)-CH(12); f, 4H(F15)-CH(10); g, 3, 5H(F15)-CH(10); h, 4H(FlS)—CH(8); i, 3, 5H(F15)-CH(8); j, 4H(F15)-CH(11); k, 3, 5H(F15)-CH(14). '1 H l-fi N 212 v a m. . U' o 6.6 1.3 v ' L14111111111111111111111111111111111111111411111111111111111 m m 1.» E: '9 at o h 01 O at m e m 7.0 ‘ 4"" ‘3'; ‘3l:" ‘. lllllllllllllllllllll[1111]lllllllllllllrlllllllll[ITTIlllllll 7.0 6.8 6.6 6.4 6.2 6.0 5.8 5.6 F2 (ppm) in 212 08 0A am. .O Q. . a TIAJFCJ4211143J m 8 O 1 5 r w 0 6 2 6 H 1 4 6 _~—ddd«q.—«ddfidqqud—d—qq—qq—q—qqfiuj 6 6 5.8 5.6 6.0 6.2 6.8 6.6 6.4 7.0 13 F2 (m) 213 Figure 5.7. The relative volumes (V "/6) of some of the NOEs observed for the bound RA vs. the internuclear distance (r) measured from the crystal structure of the holo- CRABPII in complex with all-trans-RA. The straight line represents a linear relationship between the relative volume and the distance. 214 Relative NOE Volume (V"’6) 2v eon—~35 520352:— QN o6 mé o4. m.m o.m mN CF.O . _ . p _ . _ . _ . . Iiimr. N; 1 - . 13-12 1.75: INF-”2 - 213+: F .. 12in 3.0 1 $2.5 .. I o F .o 1 622$: 3 5221-1... I . 5921-12 INTI: . 1 I .. 3 o $1-13 :m-E omd .. . 12+: I mud 1 215 from the crystal structure of holo-CRABPII (Kleywegt et al., 1994), suggesting that the conformation of the bound RA in solution is similar to that in the crystalline state. The few discrepancies between the observed NOEs and the distances measured from the holo crystal structure were probably caused by the internal motions of the bound RA in solution rather than a different conformation as discussed below. It is well known that NOE between an isolated pair of protons H,-H2 is proportional to cross-relaxation rate 0, which is determined by the internuclear distance r between the two protons and the spectral densities J‘q’(00) at two frequencies (Chapter 1). o (HlHHZ) = 74Hh2/4r" {6.10" (20)") - J ‘0’ (0)} where y” is the proton gyromagnetic ratio. However, for a pair of protons belonging to a protein or a ligand bound to a protein the situation is usually much more complicated. In particular, internal motions may have significant effects on the spectral density. Furthermore, cross-correlation with additional spins may also change the apparent cross- relaxation rate (Werbelow & Grant, 1977). While these factors may complicate the accurate measurement of inter-proton distances, they are also part of the basis for studying internal motions of biomacromolecules. If two pairs of protons in the same molecule have the same inter-nuclear distance but different NOE intensities, it suggests that their motions relative to other parts of the molecule are different. Analysis of the intramolecular NOEs of the bound RA indicated that all NOEs involving CH(7) are weaker than expected. For example, the distances between CH(7) and CH(8), between CH(lO) and CH(1 l) and between CH(1 l) and CH(12) protons all are ~3.0 A. Medium NOE cross-peaks are expected for all three pairs of protons if their 216 relaxation times (T, and T 2) are much longer than the mixing time (100 ms) and their positions in the binding pocket are fixed. However, the volume of the NOE cross-peak between CH(7) and CH(8) was significantly lower that those of the other two pairs of protons (Table 5.3 and Figure 5.6 and 5.7). By contrast, the corresponding peaks in both 2D DQF-COSY and TOCSY spectra were all relative strong (Figure 5.5), indicating that the relaxation times of these protons were similar to the others, and were longer than the 100 ms mixing time. The weakness of the NOE cross-peaks involving CH(7) was most likely due to the relative motions of the proton. The nature of such relative motions could be either intramolecular or intermolecular. For RA, if disregarding the methyl groups, the possible intramolecular motion is about the 6-s bond (Honig etal., 1971), which affects the NOEs of the protons, among the others, CH(7) and CH(8). The intermolecular motions could be either the whole RA molecule moves as a unit relative to the protein or local motions restricted to certain parts of the RA molecule. The motions of the whole RA molecule would attenuate the intensities of all the NOEs involving the RA protons to similar extent. Thus, the attenuation of the peak intensity is most likely due to the local motions relative to a portion of RA. The weakness of the CH(7)-CH(8) cross-peak can be caused by either intramolecular motions about 6-s bond or local intermolecular motions restricted to their locations. In either case it suggested that the ionone ring of the bound RA is not fixed as snugly in the protein as suggested by the crystal structure of holo-CRABPII. The NOEs involving CH(14), on the other hand, were all stronger than the average. In fact, the NOE peaks between CH(8) and CH(14) and between CH( 10) and CH(14) protons were stronger than that between CH(7) and CH(8) protons (Table 5.3). In 217 the crystal structure of holo-CRABPII, the distances between the CH(10) and CH(14) protons, between the CH(8) and CH(14) protons and between the CH(7) and CH(8) protons were ~4.8 A, ~6.8 A and ~3.0 A, respectively. Obviously, spin diffusion contributed to the appearance of the NOE cross-peak between CH(8) and CH(14) protons and to the strength of the NOE cross-peak between CH(lO) and CH(14) protons. The spin diffusions could be mediated by protons on the protein and/or RA. Since only the methyl group of Met123 is within 3 A distance of CH(14) and the methyl group is ~7.5 A away from CH(8), the spin diffusions are most likely through CH(12). Since only a structure rigid on the time scale relevant to the NOESY experiment can provide efficient pathways for spin diffusion, it suggests that the carboxylate end of the isoprene tail was essentially immobile relative to the protein. Very interestingly, the rigidity of the carboxylate end of the isoprene tail and the flexibility of the B-ionone of the bound RA molecule correlate well with the binding affinities and specificity of CRABPs for the retinoids. It has been shown that the carboxylic group is critical for RA to bind to CRABPS, while the modifications at the B- ionone ring and the C-7 and C-8 positions can be tolerated. Our conclusion of the flexibility of the bound RA relative to the protein appears to be inconsistent with what was reported for mouse holo-CRABPII (Norris, et al., 1994). Their conclusion of a single, static conformation of the bound RA at the 6-s bond was based solely on intramolecular NOEs between the olefinic protons and the methyl protons. It was derived under the assumptions that both the protein and RA are rigid, and the NOE volume is proportional to 1". Furthermore, only three NOEs picked from an NOESY spectrum with 30 ms mixing time had values much higher than the noise. All 218 others have intensities similar to the noise. In fact, most NOEs involving CH(7) were also lower than expected by their model, suggesting that there are also motions in the RA bound to mouse CRABPII. Interestingly, similar intermolecular motions of the bound RA relative to the protein was observed in mouse holo-CRABPI. A mechanism for the exit of RA. The intensities of most of the intermolecular NOEs between the bound RA and the protein agreed, at least qualitatively, with those predicted from the crystal structure (Table 5 .3). However, differences between the solution and crystalline state were also observed. The predicted NOEs between the H‘5 protons of Arg59 and CH(7) of the bound RA, between the H5 protons of Arg59 and CH(8) and between the H5 protons of Leul9 and CH(8) were not observed, and the intensity of the NOE between the H‘3 protons of Ala32 and CH(7) was much lower than predicted. The intensities of these NOEs must be attenuated by intermolecular motions. In the holo crystal structure, the side-chain of Arg59 extended from the BC-BD loop to the Glu74 at the C-terminus of BB. The motion of this side-chain relative the bound RA could temporarily open up the ligand entrance. Such flexibility may be necessary for the exit of the bound RA. 219 Conclusion. The 'H and [5N resonance assignments of holo-CRABPII have been completed. Most of the observed intermolecular NOEs between the protein and bound RA and the intramolecular NOEs of the bound RA are, at least qualitatively, consistent with those predicted from the crystal structure of holo-CRABPII. The results suggested that RA binds to CRABPII in solution in the same manner as determined by crystallography. Comparison of the intensities of backbone amide resonances between apo- and holo- CRABPII indicated that the ligand entrance of holo—CRABPII is much less flexible than that of apo-CRABPII. Relative motions between the bound RA and human CRABPII in solution, particularly at the ligand entrance, were observed. Such motions may provide a transient pathway for the exit of RA. 220 References Chambon, P., Olson, J. A., & Ross, A. C. (coordinators) (1996) The retinoid revolution, FASEB J. 10, 938-1048. Hodson, M. E., Toner, .1. J ., & Cistola, D. P. (1995).]. Biomol. NMR 6, 198-210. Hodson, M. E., Ponder, J. W., & Cistola, D. P. (1996) J. Mol. Biol. 264, 585-602. Hodson, M. E., & Cistola, D. P. (1997a) Biochemistry 36, 1450-1460. Hodson, M. E., & Cistola, D. P. (1997b) Biochemistry 36, 2278-2290. Honig, B., Hudson, 3., Sykes, B. D., & Karplus, M. (1971) Proc. Natl. Acad. Sci. U. S. A. 68, 1289-1293. Jamison, R. S., Newcomer, M. E., & Ong, D. E. (1994) Biochemistry 33, 2873-2879. Kleywegt, G. J., Bergfors, T., Senn, H., 1e Motte, P., Gsell, B., Shuao, K., & Jones, T. A. (1994) Structure 2, 1241-1258. Norris, A. W., Rong, D., d’Avignon, D. A., Rosenberger, M., Tasaki, K., & Li. B. (1995) Biochemistry 34, 15564-15573. Rizo, J ., Liu, Z. P., & Gierasch, L. M. (1994) J. Biomol. NMR 4, 741-760. Spom, M. B., Roberts, A. B. & Goodman, D. S. (1994) The Retinoids: Biology, Chemistry, and Medicine, 2"d ed. Raven Press, New York. Sporn, M. B., Roberts, A. B. & Goodman, D. S. (1994) The Retinoids: Biology, Chemistry, and Medicine, 2"d ed. Raven Press, New York. Waterhous, D. V., & Muccio, D. D. (1990) Magn. Reson. Chem. 28, 223-226. Werbelow, L. G., & Grant, D. M. (1977) Adv. Magn. Reso. 9, 189-299. 111011an STATE UNIV. LIBRARIES 111111111111111111111111111111111111111111111111111111111111111 31293016900684