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 31 毫秒
1.
X Qian  M A Weiss 《Biochemistry》1992,31(33):7463-7476
Solution structures of mutant Zn fingers containing aromatic substitutions in the hydrophobic core are determined by 2D-NMR spectroscopy and distance-geometry/simulated annealing (DG/SA). The wild-type domain (designated ZFY-6) is derived from the human male-associated protein ZFY and represents a sequence motif (Cys-X2-Cys-X-Ar-X7-Leu-X2-His-X4-His) that differs from the consensus (Cys-X2,4-Cys-X3-Phe-X5-Leu-X2-His-X3-His) in the location ("aromatic swap") and diversity (Ar = tyrosine, phenylalanine, or histidine) of the central aromatic residue (underlined). In a given ZFY domain the choice of a particular aromatic residue is invariant among vertebrates, suggesting that alternative "swapped" aromatic residues are functionally inequivalent. 2D-NMR studies of analogues containing tyrosine, phenylalanine, or histidine at the swapped site yield the following results. (i) The three DG/SA structures each retain the beta beta alpha motif and exhibit similar staggered-horizontal packing between the variant aromatic residue and the proximal histidine in the hydrophobic core. (ii) The structures and stabilities of the tyrosine and phenylalanine analogues are essentially identical, differing only by local exposure of polar (Tyr p-OH) or nonpolar (Phe p-H) surfaces. (iii) The dynamic stability of the histidine analogue is reduced as indicated by more rapid protein-deuterium exchange of hydrogen bonds related to secondary structure and amide-sulfur coordination (slowly exchanging amide resonances in D2O) and by more extensive averaging of main-chain dihedral angles (3J alpha NH coupling constants). An aspartic acid in the putative DNA recognition surface, whose configuration is well-defined as a possible helix N-cap in the tyrosine and phenylalanine analogues, exhibits multiple weak main-chain contacts in the NOESY spectrum of the histidine analogue; such NOEs are geometrically inconsistent and so provide complementary evidence for structural fluctuations. (iv) Because the three DG ensembles have similar apparent precision, the finding of reduced dynamic stability in the histidine analogue emphasizes the importance of experiments that directly probe fluctuations at several time scales. Our results provide insight into the design of biological metal-binding sites and the relationship of protein sequence to structure and dynamics.  相似文献   

2.
ZFY, a male-associated Zn-finger protein encoded by the human Y chromosome, exhibits a distinctive two-finger repeat: whereas odd-numbered domains fit a general consensus, even-numbered domains exhibit systematic differences. Do these odd and even sequences encode structurally distinct surfaces for DNA recognition? As a first step toward answering this question, we have recently described the sequential 1H NMR assignment of a representative nonconsensus Zn finger (designated ZFY-6T) based on 2D NMR studies of a 30-residue peptide [Kochoyan, M., Havel, T.F., Nguyen, D.T., Dahl, C.E., Keutmann, H. T., & Weiss, M.A. (1991) Biochemistry 30, 3371-3386]. Initial structural modeling by distance geometry/simulated annealing (DG/SA) demonstrated that this peptide retained the N-terminal beta-hairpin and C-terminal alpha-helix (beta beta alpha motif) observed in consensus Zn fingers. However, the precision of this initial structure was limited by resonance overlap, which led to ambiguities in the assignment of key NOEs in the hydrophobic core. In this paper these ambiguities are resolved by selective deuterium labeling, enabling a refined structure to be calculated by DG/SA and restrained molecular dynamics. These calculations provide a detailed view of the hydrophobic core and protein surface, which are analyzed in reference to previously characterized Zn fingers. Variant (even) and consensus (odd) aromatic residues Y10 and F12, shown in an "aromatic swap" analogue to provide equivalent contributions to the hydrophobic core [Weiss, M.A., & Keutmann, H.T. (1990) Biochemistry 29, 9808-9813], nevertheless exhibit striking differences in packing interactions: Y10--but not F12--contributes to a contiguous region of the protein surface defined by putative specificity-determining residues. Alternating surface architectures may have implications for the mechanism of DNA recognition by the ZFY two-finger repeat.  相似文献   

3.
Phage T4 is among the best-characterized biological systems (S. Kanamaru and F. Arisaka, Seikagaku 74:131-135, 2002; E. S. Miller et al., Microbiol. Mol. Biol. Rev. 67:86-156, 2003; W. B. Wood and H. R. Revel, Bacteriol. Rev. 40:847-868, 1976). To date, several genomes of T4-like bacteriophages are available in public databases but without any APEC bacteriophages (H. Jiang et al., Arch. Virol. 156:1489-1492, 2011; L. Kaliniene, V. Klausa, A. Zajanckauskaite, R. Nivinskas, and L. Truncaite, Arch. Virol. 156:1913-1916, 2011; J. H. Kim et al., Vet. Microbiol. 157:164-171, 2012; W. C. Liao et al., J. Virol. 85:6567-6578, 2011). We isolated a bacteriophage from a duck factory, named HX01, that infects avian pathogenic Escherichia coli (APEC). Sequence and morphological analyses revealed that phage HX01 is a T4-like bacteriophage and belongs to the family Myoviridae. Here, we announce the complete genome sequence of phage HX01 and report the results of our analysis.  相似文献   

4.
Hydrogen-exchange rates were measured for RNase T1 and three variants with Ala --> Gly substitutions at a solvent-exposed (residue 21) and a buried (residue 23) position in the helix: A21G, G23A, and A21G + G23A. These results were used to measure the stabilities of the proteins. The hydrogen-exchange stabilities (DeltaG(HX)) for the most stable residues in each variant agree with the equilibrium conformational stability measured by urea denaturation (DeltaG(U)), if the effects of D(2)O and proline isomerization are included [Huyghues-Despointes, B. M. P., Scholtz, J. M., and Pace, C. N. (1999) Nat. Struct. Biol. 6, 210-212]. These residues also show similar changes in DeltaG(HX) upon Ala --> Gly mutations (DeltaDeltaG(HX)) as compared to equilibrium measurements (DeltaDeltaG(U)), indicating that the most stable residues are exchanging from the globally unfolded ensemble. Alanine is stabilizing compared to glycine by 1 kcal/mol at a solvent-exposed site 21 as seen by other methods for the RNase T1 protein and peptide helix [Myers, J. K., Pace, C. N., and Scholtz, J. M. (1997) Proc. Natl. Acad. Sci. U.S.A. 94, 3833-2837], while it is destabilizing at the buried site 23 by the same amount. For the A21G variant, only local NMR chemical shift perturbations are observed compared to RNase T1. For the G23A variant, large chemical shift changes are seen throughout the sequence, although X-ray crystal structures of the variant and RNase T1 are nearly superimposable. Ala --> Gly mutations in the helix of RNase T1 at both helical positions alter the native-state hydrogen-exchange stabilities of residues throughout the sequence.  相似文献   

5.
The electrostatic behavior of potentially titrating groups in reduced human thioredoxin was investigated using two-dimensional (2D) 1H and 15N nuclear magnetic resonance (NMR) spectroscopy. A total of 241 chemical shift titration curves were measured over the pH range of 2.1-10.6 from homonuclear 1H-1H Hartmann-Hahn (HOHAHA) and heteronuclear 1H-15N Overbodenhausen correlation spectra. Nonlinear least-squares fits of the data to simple relationships derived from the Henderson-Hasselbalch equation led to the determination of pKas for certain isolated ionizable groups, including the single histidine residue at position 43 (pKa = 5.5 +/- 0.1) and a number of aspartic and glutamic acid carboxylate groups. Many of the titration curves demonstrate complex behavior due to the effects of interacting titrating groups, the long range of electrostatic interactions through the protein interior, and, perhaps, pH-induced conformational changes on the chemical shifts. Unambiguous assignment of the pKas for most of the 38 potentially ionizing groups of human thioredoxin could therefore not be made. In addition, there was no clear evidence that Asp-26 titrates in a manner corresponding to that observed in the Escherichia coli protein [Dyson, H. J., Tennant, L. L., & Holmgren, A. (1991) Biochemistry 30, 4262-4268]. The pKas of the active site cysteines were measured, however, with Cys-32 having an anomalously low value of 6.3 +/- 0.1 and that of Cys-35 between 7.5 and 8.6. These pKas are in agreement with proposed mechanisms for redox catalysis of thioredoxin and previously measured pKas within the active site of E. coli thioredoxin [Kallis, G. B., & Holmgren, A. (1980) J. Biol. Chem. 255, 10261-10265]. The stabilization of a thiolate anion at physiological pH can be explained by the interaction of the S gamma of Cys-32 with the amide of Cys-35 observed in the previously determined high-resolution solution structure of reduced human thioredoxin [Forman-Kay, J. D., Clore, G. M., Wingfield, P. T., & Gronenborn, A. M. (1991) Biochemistry 30, 2685-2698].  相似文献   

6.
Continuous-wave and pulsed electron paramagnetic resonance have been applied to the study of the Cu(II) site of the copper-resistance protein PcoC from Escherichia coli and certain variant forms. Electron spin echo envelope modulation (ESEEM) experiments confirm the presence of two histidine ligands, His1 and His92, at the Cu(II) site of wild-type PcoC, consistent with the available X-ray crystallographic data for the homolog CopC (67% sequence identity) from Pseudomonas syringae pv. tomato. The variants H1F and H92F each lack one of the histidine residues close to the Cu(II) site. The ESEEM data suggest that the surviving histidine residue remains as a ligand. The nA variant features an extra alanine residue at the N terminus, which demotes the His1 ligand to position 2. At least one of the two histidine residues is bound at the Cu(II) site in this form. Simulation of the (14)N superhyperfine structure in the continuous-wave spectra confirms the presence of at least three nitrogen-based ligands at the Cu(II) sites of the wild-type, H92F and nA forms, while the H1F variant has two nitrogen ligands. The spectra of wild-type form can be fitted adequately with a 3N or a 4N model. The former is consistent with the crystal structure of the CopC homolog, where His1 acts as a bidentate ligand. The latter raises the possibility of an additional unidentified nitrogen ligand. The markedly different spectra of the H1F and nA forms compared with the wild-type and H92F proteins further highlight the integral role of the N-terminal histidine residue in the high-affinity Cu(II) site of PcoC.  相似文献   

7.
The molecular structures of three phosphorus-based peptide inhibitors of aspartyl proteinases complexed with penicillopepsin [1, Iva-L-Val-L-Val-StaPOEt [Iva = isovaleryl, StaP = the phosphinic acid analogue of statine [(S)-4-amino-(S)-3-hydroxy-6-methylheptanoic acid] (IvaVVStaPOEt)]; 2, Iva-L-Val-L-Val-L-LeuP-(O)Phe-OMe [LeuP = the phosphinic acid analogue of L-leucine; (O)Phe = L-3-phenyllactic acid; OMe = methyl ester] [Iva VVLP(O)FOMe]; and 3, Cbz-L-Ala-L-Ala-L-LeuP-(O)-Phe-OMe (Cbz = benzyloxycarbonyl) [CbzAALP(O)FOMe]] have been determined by X-ray crystallography and refined to crystallographic agreement factors, R ( = sigma parallel to F0 magnitude of - Fc parallel to/sigma magnitude of F0), of 0.132, 0.131, and 0.134, respectively. These inhibitors were designed to be structural mimics of the tetrahederal transition-state intermediate encountered during aspartic proteinase catalysis. They are potent inhibitors of penicillopepsin with Ki values of 1, 22 nM; 2, 2.8 nM; and 3, 1600 nM, respectively [Bartlett, P. A., Hanson, J. E., & Giannousis, P. P. (1990) J. Org. Chem. 55, 6268-6274]. All three of these phosphorus-based inhibitors bind virtually identically in the active site of penicillopepsin in a manner that closely approximates that expected for the transition state [James, M. N. G., Sielecki, A.R., Hayakawa, K., & Gelb, M. H. (1992) Biochemistry 31, 3872-3886]. The pro-S oxygen atom of the two phosphonate inhibitors and of the phosphinate group of the StaP inhibitor make very short contact distances (approximately 2.4 A) to the carboxyl oxygen atom, O delta 1, of Asp33 on penicillopepsin. We have interpreted this distance and the stereochemical environment of the carboxyl and phosphonate groups in terms of a hydrogen bond that most probably has a symmetric single-well potential energy function. The pro-R oxygen atom is the recipient of a hydrogen bond from the carboxyl group of Asp213. Thus, we are able to assign a neutral status to Asp213 and a partially negatively charged status to Asp33 with reasonable confidence. Similar very short hydrogen bonds involving the active site glutamic acid residues of thermolysin and carboxypeptidase A and the pro-R oxygen of bound phosphonate inhibitors have been reported [Holden, H. M., Tronrud, D. E., Monzingo, A. F., Weaver, L. H., & Matthews, B. W. (1987) Biochemistry 26, 8542-8553; Kim, H., & Lipscomb, W. N. (1991) Biochemistry 30, 8171-8180].(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

8.
Shiraishi Y  Imanishi M  Sugiura Y 《Biochemistry》2004,43(20):6352-6359
In the DNA recognition mode of C(2)H(2)-type zinc fingers, the finger-finger connection region, consisting of the histidine spacing (HX(3-5)H) and linker, would be important for determining the orientation of the zinc finger domains. To clarify the influence of spacing between two ligand histidines in the DNA binding, we exchanged the histidine spacing between Sp1 and GLI zinc fingers, which have an HX(3)H-TGEKK linker (typical) and an HX(4)H-SNEKP linker (atypical), respectively. A significant decrease in the DNA binding affinity and specificity is found in Sp1-type peptides, whereas GLI-type peptides show a mild reduction. To evaluate the effect of the linker characteristics, we further designed Sp1-type mutants with an SNEKP linker. As a result, the significant effect of the histidine spacing in Sp1-type peptides was reduced. These results demonstrate that (1) the histidine spacing significantly affects the DNA binding of zinc finger proteins and (2) the histidine spacing and the following linker regions are one effective target for regulating the DNA recognition mode of zinc finger proteins.  相似文献   

9.
Resonance Raman spectroscopy and step-scan Fourier transform infrared (FTIR) spectroscopy have been used to identify the ligation state of ferrous heme iron for the H93G proximal cavity mutant of myoglobin in the absence of exogenous ligand on the proximal side. Preparation of the H93G mutant of myoglobin has been previously reported for a variety of axial ligands to the heme iron (e.g., substituted pyridines and imidazoles) [DePillis, G., Decatur, S. M., Barrick, D., and Boxer, S. G. (1994) J. Am. Chem. Soc. 116, 6981-6982]. The present study examines the ligation states of heme in preparations of the H93G myoglobin with no exogenous ligand. In the deoxy form of H93G, resonance Raman spectroscopic evidence shows water to be the axial (fifth) ligand to the deoxy heme iron. Analysis of the infrared C-O and Raman Fe-C stretching frequencies for the CO adduct indicates that it is six-coordinate with a histidine trans ligand. Following photolysis of CO, a time-dependent change in ligation is evident in both step-scan FTIR and saturation resonance Raman spectra, leading to the conclusion that a conformationally driven ligand switch exists in the H93G protein. In the absence of exogenous nitrogenous ligands, the CO trans effect stabilizes endogenous histidine ligation, while conformational strain favors the dissociation of histidine following photolysis of CO. The replacement of histidine by water in the five-coordinate complex is estimated to occur in < 5 micros. The results demonstrate that the H93G myoglobin cavity mutant has potential utility as a model system for studying the conformational energetics of ligand switching in heme proteins such as those observed in nitrite reductase, guanylyl cyclase, and possibly cytochrome c oxidase.  相似文献   

10.
A new family of protein domains consisting of 50-80 amino acid residues is described. It is composed of nearly 40 members, including domains encoded by plastid and phage group I introns; mitochondrial, plastid, and bacterial group II introns; eubacterial genomes and plasmids; and phages. The name "EX1HH-HX3H" was coined for both domain and family. It is based on 2 most prominent amino acid sequence motifs, each encompassing a pair of highly conserved histidine residues in a specific arrangement: EX1HH and HX3H. The "His" motifs often alternate with amino- and carboxy-terminal motifs of a new type of Zn-finger-like structure CX2,4CX29-54[CH]X2,3[CH]. The EX1HH-HX3H domain in eubacterial E2-type bacteriocins and in phage RB3 (wild variant of phage T4) product of the nrdB group I intron was reported to be essential for DNA endonuclease activity of these proteins. In other proteins, the EX1HH-HX3H domain is hypothesized to possess DNase activity as well. Presumably, this activity promotes movement (rearrangement) of group I and group II introns encoding the EX1HH-HX3H domain and other gene targets. In the case of Escherichia coli restrictase McrA and possibly several related proteins, it appears to mediate the restriction of alien DNA molecules.  相似文献   

11.
Two variants of the cytochrome c1 component of the Rhodobacter capsulatus cytochrome bc1 complex, in which Met183 (an axial heme ligand) was replaced by lysine (M183K) or histidine (M183H), have been analyzed. Electron paramagnetic resonance (EPR) and magnetic circular dichroism (MCD) spectra of the intact complex indicate that the histidine/methionine heme ligation of the wild-type cytochrome is replaced by histidine/lysine ligation in M183K and histidine/histidine ligation in M183H. Variable amounts of histidine/histidine axial heme ligation were also detected in purified wild-type cytochrome c1 and its M183K variant, suggesting that a histidine outside the CSACH heme-binding domain can be recruited as an alternative ligand. Oxidation-reduction titrations of the heme in purified cytochrome c1 revealed multiple redox forms. Titrations of the purified cytochrome carried out in the oxidative or reductive direction differ. In contrast, titrations of cytochrome c1 in the intact bc1 complex and in a subcomplex missing the Rieske iron-sulfur protein were fully reversible. An Em7 value of -330 mV was measured for the single disulfide bond in cytochrome c1. The origins of heme redox heterogeneity, and of the differences between reductive and oxidative heme titrations, are discussed in terms of conformational changes and the role of the disulfide in maintaining the native structure of cytochrome c1.  相似文献   

12.
A 13C NMR study is reported of switch variant anti-dansyl antibodies, which possess the identical VH, VL, and CL domains in conjunction with highly homologous but not identical heavy-chain constant regions. Each of these antibodies has been selectively labeled with 13C at the carbonyl carbon of Trp, Tyr, His, or Cys residue by growing hybridoma cells in serum-free medium. Spectral assignments have been made by following the procedure described previously for the switch variant antibodies labeled with [1-13C]Met [Kato, K., Matsunaga, C., Igarashi, T., Kim, H., Odaka, A., Shimada, I., & Arata, Y. (1991) Biochemistry 30, 270-278]. On the basis of the spectral data collected for the antibodies and their proteolytic fragments, we discuss how 13C NMR spectroscopy can be used for the structural analyses of antigen binding and also of domain-domain interactions in the antibody molecule.  相似文献   

13.
The first crystallographic structure of human type 3 3alpha-hydroxysteroid dehydrogenase (3alpha-HSD3, AKR1C2), an enzyme playing a critical role in steroid hormone metabolism, has been determined in complex with testosterone and NADP at 1.25-A resolution. The enzyme's 17beta-HSD activity was studied in comparison with its 3alpha-HSD activity. The enzyme catalyzes the inactivation of dihydrotestosterone into 5alpha-androstane-3alpha,17beta-diol (3alpha-diol) as well as the transformation of androstenedione into testosterone. Using our homogeneous and highly active enzyme preparation, we have obtained 150-fold higher 3alpha-HSD specificity as compared with the former reports in the literature. Although the rat and the human 3alpha-HSDs share 81% sequence homology, our structure reveals significantly different geometries of the active sites. Substitution of the Ser(222) by a histidine in the human enzyme may compel the steroid to adopt a different binding to that previously described for the rat (Bennett, M. J., Albert, R. H., Jez, J. M., Ma, H., Penning, T. M., and Lewis, M. (1997) Structure 5, 799-T812). Furthermore, we showed that the affinity for the cofactor is higher in the human 3alpha-HSD3 than the rat enzyme due to the presence of additional hydrogen bonds on the adenine moiety and that the cofactor is present under its reduced form in the active site in our preparation.  相似文献   

14.
Ribonuclease H (RNase H) from Escherichia coli is an endonuclease that specifically degrades the RNAs of RNA:DNA hybrids. The enzyme is a single polypeptide chain of 155 amino acid residues, of which 4 are methionines. To solve the crystallographic three-dimensional structure of E. coli RNase H by the multi-wavelength anomalous diffraction technique, we have constructed methionine auxotrophic strains of E. coli that overexpress selenomethionyl RNase H. MIC88 yields about 10 mg of selenomethionyl RNase H per liter of culture, which is comparable to the overexpression of the natural recombinant protein. We have purified both proteins to homogeneity and crystallized them isomorphously in the presence of sulfate. These are Type I crystals of space group P2(1)2(1)2(1) with the cell parameters a = 41.8 A, b = 86.4 A, c = 36.4 A, one monomer per asymmetric unit, and approximately 36% (v/v) solvent. Crystals of both proteins diffract to beyond 2-A Bragg spacings and are relatively durable in an x-ray beam. On replacement of sulfate with NaCl, crystals of natural RNase H grow as Type I' (very similar to Type I) at pH between 7.0 and 8.0; at pH 8.8, crystals of Type II are obtained in space group P2(1)2(1)2(1) with a = 44.3 A, b = 87.3 A, and c = 35.7 A. Type II crystals can be converted to Type I by soaking in phosphate buffer. RNase H crystals of Type II have also been reported by Kanaya et al. (Kanaya, S., Kohara, A., Miyakawa, M., Matsuzaki, T., Morikawa, K., and Ikehara, M. (1989) J. Biol. Chem. 264, 11546-11549).  相似文献   

15.
It is now generally accepted that protoporphyrinogen oxidase is the target-enzyme for diphenyl-ether-type herbicides. Recent studies [Camadro, J-M., Matringe M., Scalla, R. & Labbe, P. (1991) Biochem. J. 277, 17-21] have revealed that in maize, diphenyl ethers competitively inhibit protoporphyrinogen oxidase with respect to its substrate, protoporphyrinogen IX. In this study, we show that, in purified pea etioplast, [3H]acifluorfen specifically binds to a single class of high-affinity binding sites with an apparent dissociation constant of 6.2 +/- 1.3 nM and a maximum density of 29 +/- 5 nmol/g protein. [3H]Acifluorfen binding reaches equilibrium in about 1 min at 30 degrees C. Half dissociation occurs in less than 30 s, indicating that the binding is fully reversible. The specificity of [3H]acifluorfen binding to protoporphyrinogen oxidase is examined. [3H]Acifluorfen binding is inhibited by all the peroxidizing molecules tested. The phthalimide derivative, N-(4-chloro-2-fluoro-5-isopropoxy)phenyl-3,4,5,6-tetra hydrophthalimide, exerts a mixed-competitive inhibition on this binding. The effects of all these molecules on the binding of [3H]acifluorfen are tightly linked to their capacity to inhibit pea etioplast protoporphyrinogen oxidase activity. Furthermore, protoporphyrinogen IX, the substrate of the reaction catalyzed by protoporphyrinogen oxidase, was able to competitively inhibit the binding of [3H]acifluorfen. In contrast, protoporphyrin IX, the product of the reaction, did not inhibit this binding. All these results provide clear evidence that in pea etioplasts, [3H]acifluorfen exclusively binds to protoporphyrinogen oxidase, that the protoporphyrinogen oxidase inhibitors tested so far bind to the same region of the enzyme and that this region overlaps the catalytic site of the enzyme.  相似文献   

16.
The solution conformation of uniformly labeled 15N human thioredoxin has been studied by two-dimensional heteronuclear 15N-1H nuclear magnetic resonance spectroscopy. Assignments of the 15N resonances of the protein are obtained in a sequential manner using heteronuclear multiple quantum coherence (HMQC), relayed HMQC-correlated (COSY), and relayed HMQC-nuclear Overhauser (NOESY) spectroscopy. Values of the 3JHN alpha splittings for 87 of the 105 residues of thioredoxin are extracted from a variant of the HMQC-COSY experiment, known as HMQC-J, and analyzed to give accurate 3JHN alpha coupling constants. In addition, long-range C alpha H(i)-15N(i + 1) scaler connectivities are identified by heteronuclear multiple bond correlation (HMBC) spectroscopy. The presence of these three-bond scaler connectivities in predominantly alpha-helical regions correlates well with the secondary structure determined previously from a qualitative analysis of homonuclear nuclear Overhauser data [Forman-Kay, J. D., Clore, G. M., Driscoll, P.C., Wingfield, P. T., Richards, F. M., & Gronenborn, A. M. (1989) Biochemistry 28, 7088-7097], suggesting that this technique may provide additional information for secondary structure determination a priori. The accuracy with which 3JHN alpha coupling constants can be obtained from the HMQC-J experiment permits a more precise delineation of the beginnings and ends of secondary structural elements of human thioredoxin and of irregularities in these elements.  相似文献   

17.
We present a systematic investigation of the nature and strength of the hydrogen bonding in HX···HX and CH3X…HX (X = Br, Cl and F) dimers using ab initio MP2/aug-cc-pVTZ calculations in the framework of the quantum theory of atoms in molecules (QTAIM) and electron localisation functions (ELFs) methods. The electron density of the complexes has been characterised, and the hydrogen bonding energy, as well as the QTAIM and ELF parameters, is consistent, providing deep insight into the origin of the hydrogen bonding in these complexes. It was found that in both linear and angular HX…HX and CH3X…HX dimers, F atoms form stronger HB than Br and Cl, but they need short (~2 Å) X…HX contacts.  相似文献   

18.
A simple derivative of histamine, alpha-methylhistamine i.e. 4-(2-aminopropyl)-imidazole, was shown to potently inhibit the K+-induced release of [3H] histamine from slices of rat cerebral cortex previously incubated in the presence of [3H] histidine. The maximal inhibition elicited by alpha-methylhistamine was of about 60% i.e. similar to that elicited by exogenous histamine. The effect occurred with an EC50 value of 4.3 +/- 1.1 X 10(-9) M about 10 times lower than that of histamine and was reversed by a H3-receptor antagonist. Since alpha-methylhistamine is known to display negligible potency at H1- and H2-receptors, this compound appears to be the first highly potent and selective H3-receptor agonist to be identified.  相似文献   

19.
Site-directed mutagenesis studies of bovine pancreatic phospholipase A2 (PLA2, overproduced in Escherichia coli) showed that replacement of surface residue Lys-56 by a neutral or hydrophobic amino acid residue resulted in an unexpected and significant change in the function of the enzyme. The kcat for phosphatidylcholine micelles increases 3-4-fold for K56M, K56I, and K56F and ca. 2-fold for K56N and K56T but does not change for K56R. These results suggest that the side chain of residue 56 has significant influence on the activity of PLA2. In order to probe the structural basis for the enhanced activity, the crystal structures of wild-type and K56M PLA2 were determined by X-ray crystallography to a resolution of 1.8 A. The results suggest that the mutation has not only perturbed the conformation of the side chain of Met-56 locally but also caused conformational changes in the neighboring loop (residues 60-70), resulting in the formation of a hydrophobic pocket by residues Met-56, Tyr-52, and Tyr-69. Docking of a phosphatidylcholine inhibitor analogue into the active site of K56M, according to the structure of the complex of cobra venom PLA2-phosphatidylethanolamine inhibitor analogue [White, S.P., Scott, D. L., Otwinowski, Z., Gleb, M. H., & Sigler, P. (1990) Science 250, 1560-1563], showed that the choline moiety [N(CH3)3]+ is readily accommodated into the newly formed hydrophobic pocket with a high degree of surface complementarity. This suggests a possible interaction between residue 56 and the head group of the phospholipid, explaining the enhanced activities observed when the positively charged Lys-56 is substituted by apolar residues, viz., K56M, K56I, and K56F. Further support for this interpretation comes from the 5-fold enhancement in kcat for the mutant K56E with a negatively charged side chain, where there would be an attractive electrostatic interaction between the side chain of Glu-56 and the positively charged choline moiety. Our results also refute a recent report [Tomasselli, A. G., Hui, J., Fisher, J., Zürcher-Neely, H., Reardon, I.M., Oriaku, E., Kézdy, F.J., & Heinrikson, R.L. (1989) J. Biol. Chem. 264, 10041-10047] that substrate-level acylation of Lys-56 is an obligatory step in the catalysis by PLA2.  相似文献   

20.
Prior explicit allometric models are extended to predict the scaling relationship between the ability of plants with leafless stems to harvest sunlight H and total standing plant biomass M(T) (which equals the sum of standing stem and root biomass, M(S) and M(R)). Provided that H scales in a directly proportional manner (isometrically) with respect to either stem surface area (i.e.H proportional, variant SA(S) ) or total stem biomass (i.e. H proportional, variant M(S)), the allometric model presented here predicts that SA(S) proportional, variant M(T)(3/4) or M(S) proportional, variant M(T)(3/4), respectively. These alternative predictions are tested empirically using data for standing stem and root biomass gathered for the large columnar cactus species Pachycereus pringlei. Statistical comparisons between observed and predicted scaling relationships indicate that SA(S) proportional, variant M(T)(3/4), whereas M(S) proportional, variant M(T)(3/4) is mathematically inconsistent with the observation that stem biomass scales nearly isometrically with respect to root biomass. The contention that the H of leafless stems scales isometrically with respect to stem surface area is thus reasonable both theoretically and empirically.  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