首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Desulforedoxin (Dx) is a simple homodimeric protein isolated from Desulfovibrio gigas (Dg) containing a distorted rubredoxin-like center with one iron coordinated by four cysteinyl residues (7.9 kDa with 36 amino acids per monomer). In order to probe the geometry and the H-bonding at the active site of Dx, the protein was reconstituted with 113Cd and the solution structure determined using 2D NMR methods. The structure of this derivative was initially compared with the NMR solution structure of the Zn form (Goodfellow BJ et al., 1996, J Biol Inorg Chem 1:341-353). Backbone amide protons for G4, D5, G13, L11 NH, and the Q14 NH side-chain protons, H-bonded in the X-ray structure, were readily exchanged with solvent. Chemical shift differences observed for amide protons near the metal center confirm the H-bonding pattern seen in the X-ray model (Archer M et al., 1995, J Mol Biol 251:690-702) and also suggest that H-bond lengths may vary between the Fe, Zn, and 113Cd forms. The H-bonding pattern was further probed using a heteronuclear spin echo difference (HSED) experiment; the results confirm the presence of NH-S H-bonds inferred from D2O exchange data and observed in the NMR family of structures. The presence of "H-bond mediated" coupling in Dx indicates that the NH-S H-bonds at the metal center have significant covalent character. The HSED experiment also identified an intermonomer "through space" coupling for one of the L26 methyl groups, indicating its proximity to the 113Cd center in the opposing monomer. This is the first example of an intermonomer "through space" coupling. Initial structure calculations produced subsets of NMR families with the S of C28 pointing away from or toward the L26 methyl: only the subset with the C28 sulfur pointing toward the L26 methyl could result in a "through space" coupling. The HSED result was therefore included in the structure calculations. Comparison of the Fe, Zn, and 113Cd forms of Dx suggests that the geometry of the metal center and the global fold of the protein does not vary to any great extent, although the H-bond network varies slightly when Cd is introduced. The similarity between the H-bonding pattern seen at the metal center in Dx, Rd (including H-bonded and through space-mediated coupling), and many zinc-finger proteins suggests that these H-bonds are structurally vital for stabilization of the metal centers in these proteins.  相似文献   

2.
Maclura pomifera agglutinin is a tetrameric plant seed lectin with high affinity for the tumor-associated T-antigen disaccharide, Galbeta1,3GalNAcalpha, and hence for many O-linked glycopeptide structures. Unlike members of most lectin families, it lacks both metal ions and Cys residues. The structure of its complex with Galbeta1,3GalNAc was determined to 2.2 by first using multiwavelength anomalous diffraction with a lead derivative of the native protein, and then using molecular replacement with the unrefined structure as a model to solve the structure of the complex. The subunits share the beta-prism architecture and three-fold pseudo-symmetry of the related lectin jacalin, with the 21-residue beta-chains in the center of the tetramer. Interactions with the GalNAc predominate in the binding of the disaccharide. It forms a network of H-bonds with only one side chain, from an Asp residue, the amino group of the N-terminal Gly of the alpha-chain, and peptide backbone atoms of two aromatic residues. The Gal moiety does not H-bond directly with residues in the same monomer, i.e. there is no true subsite for it, but there are interactions through two water molecules. In the crystal, it interacts with residues in the binding site of an adjacent tetramer. The minimum energy conformation expected for the disaccharide is retained, despite its mediating the tetramer-tetramer interactions in the crystal packing. The resulting lattice is comparable to those seen for complexes of other lectins with branched glycopeptides.  相似文献   

3.
The solution secondary structure of the highly active Y55F/Y88F "Tyr-14-only" mutant of delta 5-3-ketosteroid isomerase complexed with 19-nortestosterone hemisuccinate has been shown to consist of three helices, a six-stranded mixed beta-sheet, and five turns. The steroid binds near the general acid, Tyr-14, on helix 1, near the general base, Asp-38, on the first strand of the beta-sheet, and on the hydrophobic face of the beta-sheet [Zhao, Q., Abeygunawardana, C., & Mildvan, A. S. (1997) Biochemistry 36, 3458-3472]. On this hydrophobic face, Asp-99 is the only polar residue. Free isomerase shows a deshielded exchangeable proton resonance at 13.1 ppm assigned to the N epsilon H of neutral His-100. Its fractionation factor (phi = 0.79) and slow exchange with solvent suggest it to be buried or involved in an H-bond. The binding of dihydroequilenin or estradiol to isomerase induces the appearance of two additional deshielded proton resonances, one at 18.2 ppm assigned to the gamma-carboxyl proton of Asp-99, and the other, at 11.6 ppm, assigned to the zeta-OH proton of Tyr-14. While mutation of Asp-99 to Ala results in the disappearance of only the resonance near 18 ppm [Wu, R. W., Ebrahemian, S., Zwrotny, M. E., Thornberg, L. D., Perez-Alverado, G. C., Brothers, P., Pollack, R. M., & Summers, M. F. (1997) Science 276, 415-418], both of these resonances disappear in mutants lacking Tyr-14, suggesting an H-bonded catalytic diad, Asp-99-COOH--Tyr14-OH--O-steroid enolate. The catalytic diad is further supported by NOEs from the beta 1 and beta 2 protons of Asp-99 to the epsilon protons of Tyr-14, and from the zeta-OH proton of Tyr-14 to the gamma-carboxyl proton of Asp-99, indicating close proximity of these two residues, and by other data from the literature. A strong, low-barrier H-bond between Asp-99 and Tyr-14 is indicated by the 6.2 ppm deshielding, low fractionation factor (phi = 0.34) and slow exchange of the resonance at 18.2 ppm. A normal H-bond between Tyr-14 and the steroid is indicated by the 1.8 ppm deshielding, fractionation factor of 0.97 and the slow exchange of the resonance at 11.6 ppm. It is suggested that the 10(4.7)-fold contribution of Tyr-14 to catalysis is made possible by strong H-bonding from Asp-99 in the catalytic diad which strengthens general acid catalysis by Tyr-14. It is also noted that highly deshielded proton resonance on enzymes between 15 and 20 ppm, assigned to low-barrier H-bonds, generally involve carboxyl groups.  相似文献   

4.
We have used near-infrared Fourier transform (pre)resonance Raman spectroscopy to determine the protein interactions with the bacteriochlorophyll (BChl) dimer constituting the primary electron donor, P, in the reaction center (RC) from the thermophilic purple sulfur bacterium Chromatium tepidum. In addition, we report the alignment of partial sequences of the L and M protein subunits of C. tepidum RCs in the vicinity of the primary donor with those of Rhodobacter sphaeroides and Rhodopseudomonas viridis. Taken together, these results enable us to propose the hydrogen-bonding pattern and the H-bond donors to the conjugated carbonyl groups of P. Selective excitation (1064-nm laser radiation) of the FT (pre)-resonance Raman spectra of P in its neutral (P degree) and oxidized (P degree +) states were obtained via their electronic absorption bands at 876 and 1240 nm, respectively. The P degree spectrum exhibits vibrational frequencies at 1608, 1616, 1633, and 1697 cm-1 which bleach upon P oxidation. The P degree + spectrum exhibits new bands at 1600, 1639, and 1719 cm-1. The 1608-cm-1 band, which downshifts to 1600 cm-1 upon oxidation, is assigned to a CaCm methine bridge stretching mode of the P dimer, indicating that each BChl molecule possesses a single axial ligand (His L181 and His M201, from the sequence alignment). The 1616- and 1633-cm-1 bands correspond to two H-bonded pi-conjugated acetyl carbonyl groups of each BChl molecule. with different H-bond strengths: the 1616-cm-1 band is assigned to the PL C2 acetyl group which is H-bonded to a histidine residue (His L176), while the 1633-cm-1 band is assigned to the PM C2 acetyl carbonyl, H-bonded to a tyrosine residue (Tyr M196). Both PL and PM C9 keto carbonyls are free from interactions and vibrate at the same frequency (1697 cm-1). Thus, the H-bond pattern of the primary donor of C. tepidum differs from that of Rb. sphaeroides in the extra H-bond to the PM C2 acetyl carbonyl group; that of PL is H-bonded to a histidine residue in both primary donors (His L168 in Rb. sphaeroides and His L176 in C. tepidum). The P degree/P degree + redox midpoint potentials were measured to be +497 and +526 mV for isolated C. tepidum RCs with and without the associated tetraheme cytochrome c subunit, respectively, and +502 mV for intracytoplasmic membranes. The positive charge localization was estimated to be 69% in favor of PL, indicating a more delocalized situation over the primary donor of C. tepidum than that of Rb. sphaeroides (estimated to be 80% on PL). These differences in physicochemical properties are discussed with respect to the proposed structural model for the microenvironment of the primary donor of C. tepidum.  相似文献   

5.
The infrared spectra of parallel-stranded (ps) hairpin duplexes with mixed A*T/G*C composition and either isolated or sequential G*C pairs were studied in comparison with antiparallel-stranded (aps) duplexes and a corresponding set of molecules with hypoxanthine as a G base analogue lacking the exocyclic amino group. The ps duplexes showed the characteristic bands for the C2=O2 and C4=O4 stretching vibrations of thymine residues in trans-Watson-Crick A*T pairing at 1683 and 1668 cm-1. The latter band was superimposed on the stretching vibration of the free C6=O6 group of guanine. Substitution of guanine by hypoxanthine inhibited the formation of ps hairpin duplexes whatever the sequence, demonstrating that in the H-bonding between G and C the 2-NH2 group is necessary for stabilizing all of the investigated ps duplexes. This result is in agreement with a model of trans-Watson-Crick G*C base pairs with two H-bonds [N2H2(G)-N3(C) and N1H(G)-O2(C)]. However, trans-Watson-Crick A*T and G*C base pairs with two H-bonds are not isomorphous, which may explain the decreased stability of the ps, but not the aps, duplexes upon increasing the number of A*T/G*C steps. Molecular modeling studies performed on two of the ps duplexes reveal the existence of propeller twist for avoiding a clash between the N2(G) and N4(C) amino groups, and favorable stacking of sequential G*C base pairs. The optimized hairpin ps duplexes invariably incorporated G*C base pairs with two H-bonds, regardless of the initial structures adopted for the force field calculations.  相似文献   

6.
The influence of an amino acid on the stability of alpha-helical structure depends on the position of the residue in the helix with respect to the ends. Short alpha helices in proteins are stabilized both by H-bonding of the main-chain NH and CO groups and by capping interactions between side chains and unfulfilled peptide groups at the N and C termini. Peptide models based on consensus position-dependent helix sequences allow one to model capping effects in isolated helices and to establish a base line for these interactions in proteins. We report here an extended series of substitutions in the cap positions of our peptide models and the solution structure of peptide S3, with serine at the N-cap position defined as the N-terminal residue with partly helix and partly coil conformation. The resulting model, determined by 2D 1H NMR, is consistent with a structure at the N-cap involving H-bonding between the serine gamma oxygen and the peptide NH of the glutamic acid residue three amino acids toward the C terminus. A bifurcated H-bond of Ser O gamma with the NH of Asp5 is possible also, since this group is within interacting distance. This provides direct evidence that specific side-chain interactions with the main chain stabilize isolated alpha-helical structure.  相似文献   

7.
Comparative sequence analysis and ERNA-3D software were used to model the three-dimensional structure of the small domain of signal recognition particle RNA. RNA secondary structures were established by allowing only phylogenetically-supported base pairs. The folding of the RNA molecules was constrained further to include a well-supported pseudoknot. Helical sections were oriented coaxially where a continuous helical stack was formed in the RNA of another species. Finally, RNA helices were placed at distances that preserved the connectivity of the molecule with the smallest number of single-stranded nucleotide residues as identified from the aligned sequences. We show that the comparative three-dimensional structure modeling approach is an extremely powerful tool as it requires only a critical number of carefully aligned sequences.  相似文献   

8.
The question of whether interchanges of spatially neighboring residues are coupled, or whether they change independently of each other, has been addressed repeatedly over the last few years. Utilizing a residue order-independent structural comparison tool, we investigated interchanges of spatially adjacent residue pairs in conserved 3D environments in globally dissimilar protein structures. We define spatially adjacent pairs to be non-local neighboring residues which are in spatial contact, though separated along the backbone, to exclude backbone effects. A dataset of unrelated structures is extensively compared, constructing a matrix of all 400x400 interchanges of residue pairs. Our study indicates that (i) interchanges of residues which are spatial neighbors are independent of each other. With the exception of a few pairs, the pattern of interchanges of pairs of adjacent residues resembles that expected from interchanges of single residues. However, clustering residues of similar characteristics, serves to enhance secondary trends. Hence, (ii) clustering the hydrophobic, aliphatic and, separately, the aromatic, and comparing them with the charged, and the polar, indicates that hydrophobic pairs are favorably replaced by hydrophobic, and charged/ polar by charged/polar. The most strongly conserved are the charged. Interestingly, the type of charge (like or opposite) plays no role. Interchanges between the hydrophobic and hydrophilic classes are unfavorable. (iii) Clustering by volume indicates that the most highly conserved are the (Small, Small) pairs. The least favorable are interchanges of the type (Small, Small) <--> (Large, Large). Interchanges of the type (Large, Small) <--> (Large, Large) are less favorable than (Large, Small) <--> (Small, Small). Compensatory interchanges of the type (Large, Small) <--> (Small, Large) are unfavorable. (iv) Inspection of the trends in the interchanges of the clustered small residues versus clustered large rigid, and separately versus clustered large flexible, illustrates clear differences. Consistently, within all hydrophobic, large and small, the flexible aliphatic differ from the more rigid aromatic. The flexible aliphatic residue pairs are unfavorably replaced by other residue types. Furthermore, (v) the unique properties of the aromatics, conferred by the electronic configuration of their benzene rings, are transformed into clear trends. Replacements of polar residues by aromatics, while unfavorable, are nevertheless consistently more favorable than into aliphatics. We address these issues and their direct implications to protein design and to fold recognition.  相似文献   

9.
The hydrogen bond distributions in 123 protein structures with the atom coordinates established at a resolution of less than 2 A were analyzed. The peculiarities of hydrogen bond distributions with respect to the lengths and remoteness of contacting residues in the primary structure were established. A hierarchy of H-bond energy distribution in the spatial structure of protein globules was demonstrated. The role of hydrogen bonds in the formation of domain structure and their special properties in proteins with different types of secondary structure are discussed.  相似文献   

10.
To gain more insight about Escherichia coli tmRNA structure, NiCR, a square planar macrocyclic nickel (II) complex, was used to probe guanine N7 exposure. On the basis of this additional structural information, a refined secondary structure of the molecule is proposed. In addition to its known specificity for guanine N7, we show here that the chemical probe can also cleave at specific uridine residues. In contrast to the alkaline-labile modification of guanine, the reactivity of NiCR at these uridine residues results in direct strand scission. To better characterize the uridine cleavage sites and assess the importance of the RNA structure for the reaction to occur, smaller RNA molecules derived from one pseudoknot (PK4) of E. coli tmRNA containing two uridine cleavage sites were engineered and probed. It is shown that this pseudoknot can fold by itself in solution and that the expected uridine residues are also cleaved by the nickel complex, suggesting that only a local sequence and/or structural context is required for cleavage. In E. coli tmRNA, the five uridine cleavage sites are located in double-stranded regions. These sites contain a G-U wobble base-pair and a downstream uridine which is cleaved. Using smaller RNAs derived from one stem of PK4, systematic changes in the proposed recognition motif indicate that the G-U pair is required for cleavage. Furthermore, there is no cleavage if the G-U pair is reversed. If the recognition motif is moved within the stem, the cleavage site moves accordingly. Additionally, if the recognition motif is changed such that the G-U pair is flanked by two uridine residues, the reactivity occurs only at the 3' uridine. Radical quenching studies have indicated that sulfate radical, as in the case of guanine oxidation, is involved in uridine oxidation. Although additional studies are required to better characterize the reaction, this paper reports a novel specificity for a chemical probe which may be useful for investigating structural motifs involving G-U pairs in folded RNAs.  相似文献   

11.
We report the crystal structure of nucleoside diphosphate kinase (NDP kinase) from Dictyostelium discoideum with thymidine diphosphate (dTDP) and Mg2+ bound at the active site. The structure has been refined to an R-factor of 18.3% at 2-A resolution. The base stacks on the aromatic ring of Phe 64 near the protein surface and is wedged between the side chains of Phe 64 and Val 116. The sugar and the pyrophosphate are deeper inside the protein and make numerous H-bonds with protein side chains. There is no backbone interaction with the nucleotide. A Mg2+ ion bridges the alpha- and beta-phosphates and interacts with the protein via water molecules. NDP kinase shows little specificity toward ribonucleotides and deoxyribonucleotides. This property, required by the enzyme biological function, can now be analyzed by comparing the crystal structures of free, ADP-ligated, and dTDP-ligated enzymes. The most significant differences are located in residues 60-64, which adapt their conformation to allow Phe 64 to stack on both types of bases. Nonspecific binding is achieved by the absence of polar interaction between the base and protein atoms. The ribose of ADP and the deoxyribose of dTDP occupy similar positions, their hydroxyl groups interacting with Lys 16 and Asn 119. The H-bond between Lys 16 and the O2' hydroxyl of ADP is replaced by a similar interaction with a water molecule in the dTDP complex. The beta-phosphate position is the same for ADP and dTDP, suggesting that the mechanism of phosphate transfer is the same for all substrates ofNDP kinase.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
We introduce a new algorithmic method for identifying the geometrical core of proteins that does not require the usual superposition of structures. A geometrical core is defined as the set of residues such that the C alpha (I) - C alpha (J) atom distances are identical in all structures of the protein family under study, where I and J are secondary structure positions in the structural units--strands, loops, or parts of them. The result of applying the algorithm to 53 Ig structures leads to the identification of two geometrical core sets of C alpha atom positions for the VL and VH domains. Applications of the core sets are described.  相似文献   

13.
The metals of the cytochrome c oxidase structures of the bovine heart mitochondrion (PDB code 1occ) and of the soil bacterium Paracoccus denitrificans (1arl) include a dicopper center (CuA), magnesium, two proximal hemes, a copper (CuB) atom, and a calcium. The mitochondrial structure also possesses a bound distant zinc ion. The extended environments of the metal sites are analyzed emphasizing residues of the second shell in terms of polarity, hydrophobicity, secondary structure, solvent accessibility, and H-bonding networks. A significant difference in the CuA metal environments concerns D-51 I in 1occ, absent from 1arl. The D-51 I appears to play an important role in the proton pumping pathway. Our analysis uncovers several statistically significant residue clusters, including a cysteine-histidine-tyrosine cluster overlapping the CuA-Mg complex; a histidine-acidic cluster enveloping the environment of Mg, the two hemes, and CuB; and on the protein surface a mixed charge cluster, which may help stabilize the quaternary structure and/or mediate docking to cytochrome c. These clusters may constitute possible pathways for electron transfer, for O2 diffusion, and for H2O movement. Many hydrogen bonding relations along the interface of subunits I and II demarcate this surface as a potential participant in proton pumping.  相似文献   

14.
Three-dimensional structures of a representative set of more than 30 hydrogen-bonded nucleic acids pairs have been studied by reliable ab initio quantum mechanical methods. We show that many hydrogen-bonded nucleic acid base pairs are intrinsically nonplanar, mainly due to the partial sp3 hybridization of nitrogen atoms of their amino groups and secondary electrostatic interactions. This finding extends the variability of intermolecular interactions of DNA bases in that i) flexibility of the base pairs is larger than has been assumed before, and ii) attractive proton-proton acceptor interactions oriented out of the base pair plane are allowed. For example, all four G-A mismatch base pairs are propeller twisted, and the energy preferences for the nonplanar structures range from less than 0.1 kcal/mol to 1.8 kcal/mol. We predict that nonplanarity of the amino group of guanine in the G(anti)...A(anti) pair of the ApG step of the d(CCAAGATTGG)2 crystal structure is an important stabilizing factor that improves the energy of this structure by almost 3 kcal/mol. Currently used empirical potentials are not accurate enough to properly cover the interactions associated with amino-group and base-pair nonplanarity.  相似文献   

15.
The betabellin target structure is a beta-sandwich protein consisting of two 32 residue beta-sheets packed against one another by interaction of their hydrophobic faces. The 32 residue chain of betabellin-15S (HSLTAKIpkLTFSIAphTYTCAV pkYTAKVSH, where p=DPro, k=DLys, and h=DHis) did not fold in water at pH 6.5. Air oxidation of betabellin-15S provided betabellin-15D, the 64 residue disulfide bridged two-chain molecule, which also remained unfolded in water at pH 6.5. By circular dichroic spectropolarimetry, the extent of beta structure observed for betabellin-15D increased with the pH and ionic strength of the solution and the betabellin-15D concentration. By electron microscopy, in 5.0 mM MOPS and 0.25 M NaCl at pH 6.9, betabellin-15D formed long narrow multimeric fibrils. A molecular model was constructed to show that the dimensions of these betabellin-15D fibrils are consistent with a single row of beta-sandwich molecules joined by multiple intersheet H-bonds.  相似文献   

16.
The pores of ion channel proteins are often modeled as static structures. In this view, selectivity reflects rigidly constrained backbone orientations. Such a picture is at variance with the generalization that biological proteins are flexible, capable of major internal motions on biologically relevant time scales. We tested for motions in the sodium channel pore by systematically introducing pairs of cysteine residues throughout the pore-lining segments. Two distinct pairs of residues spontaneously formed disulfide bonds bridging domains I and II. Nine other permutations, involving all four domains, were capable of disulfide bonding in the presence of a redox catalyst. The results are inconsistent with a single fixed backbone structure for the pore; instead, the segments that line the permeation pathway appear capable of sizable motions.  相似文献   

17.
18.
Frameshift and readthrough sites within retroviral messenger RNAs are often followed by nucleotide sequences that have the potential to form pseudoknot structures. In the work presented here, NMR methods were used to characterize the base-pairings and structural features of the RNA pseudoknot downstream of the gag-pro frameshift site of simian retrovirus type-1 (SRV-1) and a functional mutant of the SRV-1 pseudoknot. Evidence is presented that these pseudoknots contain two A-form helical stems of six base-pairs each, connected by two loops, in a classic H-type pseudoknot topology. A particularly interesting feature is that the shorter of the two connecting loops, loop 1, consists of only a single adenosine nucleotide that spans the major groove of stem 2. In this respect, the frameshift-associated pseudoknots are structurally similar to the pseudoknot within the gene 32 mRNA of bacteriophage T2, previously characterized by NMR methods. Despite having similar nucleotide sequences, the solvent exchange rates of the imino protons at the junction of the helical stems in the wild-type and mutant frameshifting pseudoknots differ from each other and from the bacteriophage T2 pseudoknot. The implications of this finding are discussed.  相似文献   

19.
The three-dimensional (3-D) structure of a RNA pseudoknot that causes the efficient ribosomal frameshifting in the gag-pro region of mouse mammary tumor virus (MMTV) has been determined recently by nuclear magnetic resonance (NMR) studies. But since the structure refinement in the studies did not use metal ions and waters, it is not clear how metal ions participate in the stabilization of the pseudoknot, and what kind of ion-RNA interactions dominate in the tertiary contacts for the RNA pseudoknotting. Based on the reported structure data of the pseudoknot VPK of MMTV, we gradually refined the structure by restrained molecular dynamics (MD) using NMR distance restraints. Restrained MD simulation of the RNA pseudoknot was performed with sodium ions and water molecules. Our results are in good agreement with known NMR data and delineate the importance of the metal ion coordination in the stability of the pseudoknot. In the non-coaxially stacking pseudoknot, stem 1 (S1), stem 2 (S2), and the intervening A14 involves unconventional stacking of base pairs coordinated by Na+ and/or bridging water molecules. A6 and G7 of loop L1 make a perfect base stacking in the major groove and are further stabilized by coordinated Na+ ions and water molecules. The first 4-nucleotide (nt) ACUC of loop L2 form a sharp turn and the following 4-nt AAAA cross the minor groove of S1 and are steadied by interactions with the nucleotides of S , bridging water molecules and coordinated Na+ ions. Our studies suggest that the metal ion plays a crucial role in the RNA pseudoknotting of VPK. In the stacking interior of S1 and S2, the Na+ ion is positioned in the major groove and interacts directly with the carbonyl group O6 of G28 and carbonyl group O4 of U13 in the wobble base pair U13:G28. The ion-RNA interactions in MMTV VPK not only stabilize the RNA pseudoknot but also modify the electrostatic properties of the nucleotides at the critical parts of the pseudoknot VPK.  相似文献   

20.
The bivalve mollusc Lucina pectinata harbors sulfide-oxidizing chemoautotrophic bacteria and expresses a monomeric hemoglobin I, HbI, with normal O2, but extraordinarily high sulfide affinity. The crystal structure of aquomet Lucina HbI has revealed an active site with three residues not commonly found in vertebrate globins: Phe(B10), Gln(E7), and Phe(E11) (Rizzi, M., Wittenberg, J. B., Coda, A., Fasano, M., Ascenzi, P., and Bolognesi, M. (1994) J. Mol. Biol. 244, 86-89). Engineering these three residues into sperm whale myoglobin results in a triple mutant with approximately 700-fold higher sulfide affinity than for wild-type. The single crystal x-ray structure of the aquomet derivative of the myoglobin triple mutant and the solution 1H NMR active site structures of the cyanomet derivatives of both the myoglobin mutant and Lucina HbI have been determined to examine further the structural origin of their unusually high sulfide affinities. The major differences in the distal pocket is that in the aquomet form the carbonyl of Gln64(E7) serves as a H-bond acceptor, whereas in the cyanomet form the amido group acts as H-bond donor to the bound ligand. Phe68(E11) is rotated approximately 90 degrees about chi2 and located approximately 1-2 A closer to the iron atom in the myoglobin triple mutant relative to its conformation in Lucina HbI. The change in orientation potentially eliminates the stabilizing interaction with sulfide and, together with the decrease in size of the distal pocket, accounts for the 7-fold lower sulfide affinity of the myoglobin mutant compared with that of Lucina HbI.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号