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 265 毫秒
1.
Asymmetric Ni(II)-Fe(II) hybrid hemoglobin, XL[alpha(Fe)beta(Fe)][alpha(Ni)beta(Ni)], in which the alpha 1 beta 1 dimer containing ferrous protoporphyrin IX and the complementary alpha 2 beta 2 dimer containing Ni(II) protoporphyrin IX were cross-linked between Lys-82 beta 1 and Lys-82 beta 2 by reaction with bis(3,5-dibromosalicyl) fumarate, was synthesized and characterized. We have previously shown that (i) Ni(II) protoporphyrin IX, which binds neither oxygen nor carbon monoxide, mimics a fixed deoxyheme with respect to its effect on the oxygen equilibrium properties of the counterpart iron subunits in both symmetric Ni(II)-Fe(II) hybrid Hbs [Shibayama, N., Morimoto, H., & Miyazaki, G. (1986) J. Mol. Biol. 192, 323-329] and (ii) the cross-linking used in this study little affects the oxygen equilibrium properties of hemoglobin [Shibayama, N., Imai, K., Hirata, H., Hiraiwa, H., Morimoto, H., & Saigo, S. (1991) Biochemistry 30, 8158-8165]. These remarkable features of our model allowed us to measure the oxygen equilibrium curves for the first two steps of oxygen binding to the alpha 1 beta 1 dimer within the hemoglobin tetramer. At all pH values examined, the affinities of this asymmetric hybrid for the first oxygen molecule are as low as those of native hemoglobin. The hybrid did not show cooperative oxygen binding at pH 6.4, while significant cooperativity was observed with rising pH; i.e., the Hill coefficient was increased from 1.41 to 1.53 upon a pH change from 7.4 to 8.4. The electronic absorption spectrum of Ni(II) protoporphyrin IX in the alpha 2 subunit was changed upon carbon monoxide (or oxygen) binding to the alpha 1 beta 1 dimer.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
We have calculated the curvature of 504 eukaryotic promoters predicted by the bent A-tract model of Bolshoy et al. (Proc. Natl. Acad. Sci. USA, 88(6), pp. 2312-16) and the bent non-A-tract models of Calladine et al. (J. Mol. Biol., 201, pp. 127-37) and Satchwell et al. (J. Mol. Biol., 191, pp. 659-75) and found in each case a correlation between TBP binding sites and DNA curvature. Characterizing the TBP binding sites revealed that in addition to the classical TATA box (TATAAA) five more elements occur significantly often in the promoters, nearly all of them being one point mutations of the classical TATA box element. Separate curvature calculations for promoters with canonical and non-canonical TATA boxes have shown that in both cases the strong curvature of the helix axes in the domain of the binding sites is maintained (classical TBP binding sites: + 64-135%, non-classical TBP binding sites: + 27-49%). These results support the proposition that beside DNA flexibility and DNA-protein interactions intrinsic curvature of DNA is one further important criterion for the recognition of different DNA elements by TBP.  相似文献   

3.
Evidence that multiple, probably non-endocytic mechanisms are involved in the uptake into mammalian cells of the alpha-helical amphipathic model peptide FLUOS-KLALKLALKALKAALKLA-NH2 (I) is presented. Extensive cellular uptake of N-terminally GC-elongated derivatives of I, conjugated by disufide bridges to differently charged peptides, indicated that I-like model peptides might serve as vectors for intracellular delivery of polar bioactive compounds. The mode of the cellular internalization of I comprising energy-, temperature-, pH- and ion-dependent as well as -independent processes suggests analogy to that displayed by small unstructured peptides reported previously (Oehlke et al., Biochim. Biophys. Acta 1330 (1997) 50-60). The uptake behavior of I also showed analogy to that of several protein-derived helical peptide sequences, recently found to be capable of efficiently carrying tagged oligonucleotides and peptides directly into the cytosol of mammalian cells (Derossi et al., J. Biol. Chem. 269 (1994) 10444-10450; Lin et al., J. Biol. Chem. 270 (1995) 14255-14258; Fawell et al., Proc. Natl. Acad. Sci. USA 91 (1994) 664-668; Chaloin et al., Biochemistry 36 (1997) 11179-11187; Vives et al., J. Biol. Chem., 272 (1997) 16010-16017).  相似文献   

4.
5.
In human hemoglobin (Hb) the beta37 tryptophan residue (betaW37), located at the hinge region of the alpha1beta2 interface, forms many contacts with alpha subunit residues of the opposite dimer, in both the T and R quaternary structures. We have carried out equilibrium O2 binding studies on a series of recombinant Hbs that have mutations at this residue site: betaW37Y, betaW37A, betaW37G, and betaW37E. Binding isotherms measured at high concentrations of these mutants were found to be shifted toward increased affinity and decreased cooperativity from that of the normal HbA0 tetramer. Analysis of these binding isotherms indicated that amino acid substitutions at the beta37 position could both destabilize the tetrameric form of the mutants relative to their constituent dimers and also alter cooperativity of the intact tetrameric species. These alterations from wild-type function are dependent on the particular side chain substituted, with the magnitude of change increasing as Trp is substituted by Tyr, Ala, Gly, and Glu. The dimer to tetramer assembly free energy of deoxy-betaW37E, the most perturbed mutant in the series, was measured using analytical gel chromatography to be 9 kcal/tetramer less favorable than that of deoxy HbA0. Stabilizing the betaW37E tetramer by addition of IHP, or by cross-linking at the alphaK99 positions, does not restore normal O2 binding behavior. Thermodynamic parameters of all the mutants were found to correlate with their CO binding rates and with their high-resolution X-ray crystal structures (see accompanying papers: Kwiatkowski et al. (1998) Biochemistry 37, 4325-4335; Peterson & Friedman (1998) Biochemistry 37, 4346-4357; Kavanaugh et al. (1998) Biochemistry 37, 4358-4373].  相似文献   

6.
The conserved residue Lys-34 in GroES was replaced by alanine and glutamic acid using site-directed mutagenesis. This residue is near the carboxy terminus of the mobile loop in GroES (residues 17-32) which becomes immobilized upon formation of the GroEL/GroES complex [Landry et al. (1993) Nature 364, 255-258]. Both charge neutralization (Lys-34-->Ala) and charge reversal (Lys-34-->Glu) at this position have little effect on the binding constant of GroES to GroEL, but they increase the enhancement by GroES of cooperativity in ATP hydrolysis by GroEL. This is reflected by a change in the Hill coefficient (at 10 mM K+) from 4.10 (+/- 0.22) in the presence of wild-type GroES to 5.17 (+/- 0.24) and 4.46 (+/- 0.14) in the presence of the GroES mutants Lys-34-->Ala and Lys-34-->Glu, respectively. The results are interpreted using the Monod-Wyman-Changeux (MWC) model for cooperativity [Monod et al. (1965) J. Mol. Biol. 12, 88-118]. They suggest that Lys-34 in GroES modulates the allosteric transition in GroEL by stabilizing a relaxed (R)-like state.  相似文献   

7.
A new framework for hemoglobin cooperativity was proposed by Ackers and colleagues on the basis of the hyper thermodynamic stability and deoxy (T) quaternary structure of one of diliganded deoxy-cyanomet hybrid hemoglobins, (alpha+CN-beta+CN-)(alpha beta), studied by hybridization of the equimolar mixture of deoxyhemoglobin and cyanomethemoglobin through a long (70-100 h) dimer exchange reaction [Daugherty et al. (1991) Proc. Natl. Acad. Sci. U.S.A. 88, 1110-1114]. Recently, we reported that the published hyperstability of (alpha+CN-beta+CN-)(alpha beta) is incorrect due to the occurrence of valency exchange between the heme sites of both parental hemoglobins during the long deoxy incubation [Shibayama et al. (1997) Biochemistry 36, 4375-4381]. We also noted a difficulty in maintaining both anaerobicity and excess free cyanide of the sample during the long incubation, which led to formation of cyanide-unbound aqometheme in the original deoxyhemoglobin resulting from the electron transfer to cyanometheme. This paper is a response to a recent argument against our work [Ackers et al. (1997) Biochemistry 36, 10822-10829]. Ackers et al. have claimed that no appreciable formation of aqomethemoglobin with their methods ensures their sample integrity, based on a supposition that our observed valency exchange may have occurred via aqometheme. In this paper, however, we demonstrate that appreciable (>27%) valency exchange really occurs between deoxy and cyanometheme sites during 72 h incubation under conditions where both anaerobicity and excess free cyanide of the sample solution are maintained by a continuous flow of humidified N2 with HCN. This confirms our view that previous experimental data on (alpha+CN-beta+CN-)(alpha beta) obtained by the long incubations should be subject to reexamination while our earlier estimation of a lower limit of free energy of (alpha+CN-beta+CN-)(alpha beta) (i.e., >/= -10.1 kcal/mol) by a rapid method (35 min) is still valid. We also suggest a possibility that the T quaternary structure of (alpha+CN-beta+CN-)(alpha beta) assigned by Ackers and colleagues using the long incubations is an artifact arising from the valency exchange. These results suggest that the putative mechanistic picture for hemoglobin cooperativity inferred from studies on deoxy-cyanomet hybrids is without foundation.  相似文献   

8.
The crystal structure of human T state hemoglobin crosslinked with bis(3,5-dibromo-salicyl) sebacate has been determined at 1.9 A resolution. The final crystallographic R factor is 0.168 with root-mean-square deviations (RMSD) from ideal bond distance of 0.018 A. The 10-carbon sebacyl residue found in the beta cleft covalently links the two betaLys82 residues. The sebacyl residue assumes a zigzag conformation with cis amide bonds formed by the NZ atoms of betaLys82's and the sebacyl carbonyl oxygens. The atoms of the crosslink have an occupancy factor of 1.0 with an average temperature factor for all atoms of 34 A2. An RMSD of 0.27 for all CA's of the tetramer is observed when the crosslinked deoxyhemoglobin is compared with deoxyhemoglobin refined by using a similar protocol, 2HHD [Fronticelli et al. J. Biol. Chem. 269: 23965-23969, 1994]. Thus, no significant perturbations in the tertiary or quaternary structure are introduced by the presence of the sebacyl residue. However, the sebacyl residue does displace seven water molecules in the beta cleft and the conformations of the beta1Lys82 and beta2Lys82 are altered because of the crosslinking. The carbonyl oxygen that is part of the amide bond formed with the NZ of beta2Lys82 forms a hydrogen bond with side chain of beta2Asn139 that is in turn hydrogen-bonded to the side chain of beta2Arg104. A comparison of the observed conformation with that modeled [Bucci et al. Biochemistry 35:3418-3425, 1996] shows significant differences. The differences in the structures can be rationalized in terms of compensating changes in the estimated free-energy balance, based on differences in exposed surface areas and the observed shift in the side-chain hydrogen-bonding pattern involving beta2Arg104, beta2Asn139, and the associated sebacyl carbonyl group.  相似文献   

9.
10.
Fascin is an actin-bundling protein that was first isolated from cytoplasmic extracts of sea urchin eggs [Kane, 1975: J. Cell Biol. 66:305-315] and was the first bundling protein to be characterized in vitro. Subsequent work has shown that fascin bundles actin filaments in fertilized egg microvilli and filopodia of phagocytic coelomocytes [Otto et al., 1980: Cell Motil. 1:31-40; Otto and Bryan, 1981: Cell Motil. 1:179-192]. Fifteen years later, the molecular cloning of sea urchin fascin [Bryan et al., 1993: Proc. Natl. Acad. Sci. U.S.A. 90:9115-9119] has led to the identification and characterization of homologous proteins in Drosophila [Cant et al., 1994: J. Cell Biol. 125:369-380], Xenopus [Holthuis et al., 1994: Biochim. Biophys. Acta. 1219:184-188], rodents [Edwards et al,. 1995: J. Biol. Chem. 270:10764-10770], and humans [Duh et al., 1994: DNA Cell Biol. 13:821-827; Mosialos et al., 1994: J. Virol. 68:7320-7328] that bundle actin filaments into structures which stabilize cellular processes ranging from mechanosensory bristles to the filopodia of nerve growth cones. Fascin has emerged from relative obscurity as an exotic invertebrate egg protein to being recognized as a widely expressed protein found in a broad spectrum of tissues and organisms. The purpose of this review is to relate the early studies done on the sea urchin and HeLa cell fascins to the recent molecular biology that defines a family of bundling proteins, and discuss the current state of knowledge regarding fascin structure and function.  相似文献   

11.
When electrospray ionization mass spectrometry (ESMS) was used to analyze purified bovine gamma E (gamma IVa)-crystallin, it yielded a relative molecular mass (M(r)) of 20.955 +/- 5. This mass is significantly different from that calculated from the published sequence (M(r) 20.894) (White HE et al., 1989, J Mol Biol 207:217-235). Further, ES-MS analysis of the protein after it had been reduced and carboxymethylated indicated the presence of five cysteine residues, whereas the published sequence contains six (Kilby GW et al., 1995, Eur Mass Spectrom 1:203-208). The entire protein sequence of gamma E crystallin has therefore been studied via a combination of ES-MS, ES-MS/MS, and Edman amino acid sequencing. The corrected sequence gives an M(r) of 20.955.3, which matches that obtained by ES-MS analysis of the purified native protein. The corrected sequence is also in agreement with a recent cDNA sequence obtained for a bovine gamma-crystallin by R. Hay (pers. comm.).  相似文献   

12.
Gene V protein of bacteriophage f1 is a dimeric protein that binds cooperatively to single-stranded nucleic acids. In order to determine whether a monomer-dimer equilibrium has an appreciable effect upon the thermodynamics of gene V protein binding to nucleic acids, the dissociation constant for the protein dimer was investigated using size-exclusion chromatography. At concentrations ranging from 5 x 10(-10) to 1.2 x 10(-5) M, the Stokes radius of the protein was that expected of the dimer of the gene V protein. The Stokes radius of the protein was also independent of salt concentration from 0.2 to 1.0 M NaCl in a buffer containing 10 mM Tris-HCl, pH 7.4, and 1 mM EDTA. The binding of the dimeric gene V protein to poly(dA) was studied using a simplified lattice model for protein-protein interactions adapted for use with a dimeric protein that binds simultaneously to two strands of nucleic acid. Interpretation of the salt dependence, C = [d log(Kint omega)]/[d log(NaCl)], of binding of such a dimeric protein to nucleic acid using the theory of Record et al. (Record, M. T., et al. (1976) J. Mol. Biol. 107, 145-158) indicates that C is a function of the numbers of cations and anions released from protein and nucleic acid upon binding of the dimer, not of the monomer. Cooperativity of gene V protein binding to poly(dA) was studied with titration experiments that are sensitive to the degree of cooperativity of binding. The cooperativity factor omega, defined as the ratio of the binding constant for a site adjacent to a previously bound dimer to that for an isolated site, was found to be relatively insensitive to salt, with a value in the range of 2000-7000 for binding to poly(dA) at 3 degrees C and at 23 degrees C. This high cooperativity factor supports the suggestion that protein-protein contacts play a major role in the formation of the superhelical gene V protein-single-stranded nucleic acid complex.  相似文献   

13.
Ca2+ binding to the N-domain of skeletal muscle troponin C (sNTnC) induces an "opening" of the structure [Gagné, S. M., et al. (1995) Nat. Struct. Biol. 2, 784-789], which is typical of Ca2+-regulatory proteins. However, the recent structures of the E41A mutant of skeletal troponin C (E41A sNTnC) [Gagné, S. M., et al. (1997) Biochemistry 36, 4386-4392] and of cardiac muscle troponin C (cNTnC) [Sia, S. K., et al. (1997) J. Biol. Chem. 272, 18216-18221] reveal that both of these proteins remain essentially in the "closed" conformation in their Ca2+-saturated states. Both of these proteins are modified in Ca2+-binding site I, albeit differently, suggesting a critical role for this region in the coupling of Ca2+ binding to the induced structural change. To understand the mechanism and the energetics involved in the Ca2+-induced structural transition, Ca2+ binding to E41A sNTnC and to cNTnC have been investigated by using one-dimensional 1H and two-dimensional {1H,15N}-HSQC NMR spectroscopy. Monitoring the chemical shift changes during Ca2+ titration of E41A sNTnC permits us to assign the order of stepwise binding as site II followed by site I and reveals that the mutation reduced the Ca2+ binding affinity of the site I by approximately 100-fold [from KD2 = 16 microM [sNTnC; Li, M. X., et al. (1995) Biochemistry 34, 8330-8340] to 1.3 mM (E41A sNTnC)] and of the site II by approximately 10-fold [from KD1 = 1.7 microM (sNTnC) to 15 microM (E41A sNTnC)]. Ca2+ titration of cNTnC confirms that cNTnC binds only one Ca2+ with a determined dissociation constant KD of 2.6 microM. The Ca2+-induced chemical shift changes occur over the entire sequence in cNTnC, suggesting that the defunct site I is perturbed when site II binds Ca2+. These measurements allow us to dissect the mechanism and energetics of the Ca2+-induced structural changes.  相似文献   

14.
NMR spectra of the downfield region of normal adult hemoglobin are reported as a function of oxygenation and temperature. Spectra were run in D2O at pD 7.4. A specially made NMR tube insert allowed precise measurement of the degree of oxygenation and of methemoglobin formation before and after taking the NMR spectrum. Plots of the estimated intensity of the most downfield prominent NMR peak, identified as arising from a deoxy-beta subunit by Davis et al. ((1971) J. Mol. Biol. 60, 101-111), versus the average degree of oxygenation y, measured optically, yield a nearly straight line within experimental error, for samples stripped of organic phosphates and for samples containing 2,3-diphosphoglycerate or inositol hexaphosphate. Intensities of peaks further upfield than this peak, previously attributed to deoxy-alpha subunits, are difficult to measure directly especially for samples containing inositol hexaphosphate. The latter samples show broadening in these alpha peaks as the degree of oxygenation increases. This extra broadening appears to increase with temperature. Linearity of the beta peak intensity with oxygenation is expected if there is no large oxygen affinity difference between alpha and beta subunits. However, the cooperativity of binding, and inaccuracy of the data, make it impossible to make accurate estimates of affinity differences.  相似文献   

15.
A strategy has been developed for quantitatively "translating" the distributions of cooperative free energy between different oxygenation analogs of hemoglobin (Hb). The method was used to resolve the cooperative free energies of all eight carbon monoxide binding intermediates. These parameters of the FeCOHb system were determined by thermodynamic transformation of corresponding free energies obtained previously for all species of the Co/FeCO system, i.e., where cobalt-substituted hemes comprise the unligated sites [Speros, P. C., et al. (1991) Biochemistry 30, 7254-7262]. Using hybridized combinations of normal and cobalt-substituted Hb, ligation analog systems Co/FeX (X = CO, CN) were constructed and experimentally quantified. Energetics of cobalt-induced structural perturbation were determined for all species of both the "mixed metal" Co/Fe system and also the ligated Co/FeCN system. It was found that major energetic perturbations of the Co/Fe hybrid species originate from a pure cobalt substitution effect on the alpha subunits. These perturbations are transduced to the beta subunit within the same dimeric half-tetramer, resulting in alteration of the free energies for binding at the nonsubstituted (Fe) sites. Using the linkage strategy developed in this study along with the determined energetics of these couplings, the experimental assembly free energies for the Co/FeCO species were transformed into cooperative free energies of the 10 Fe/FeCO species. The resulting values were found to distribute according to predictions of a symmetry rule mechanism proposed previously [Ackers, G. K., et al. (1992) Science 255, 54-63]. Their distribution is consistent with accurate CO binding data of normal Hb [Perrella, M., et al. (1990b) Biophys. Chem. 37, 211-223] and also with accurate O2 binding data obtained under the same conditions [Chu, A. H., et al. (1984) Biochemistry 23, 604-617].  相似文献   

16.
A new, simple method for predicting transmembrane segments in integral membrane proteins has been developed. It is based on low-stringency dot-plots of the query sequence against a collection of non-homologous membrane proteins using a previously derived scoring matrix [Cserz? et al., 1994, J. Mol. Biol., 243, 388-396]. This so-called dense alignment surface (DAS) method is shown to perform on par with earlier methods that require extra information in the form of multiple sequence alignments or the distribution of positively charged residues outside the transmembrane segments, and thus improves prediction abilities when only single-sequence information is available or for classes of membrane proteins that do not follow the 'positive inside' rule.  相似文献   

17.
We find that specific oxidation for the Met-192 residue in delta-chymotrypsin to methionine sulfoxide results in a twofold increase in Km(app) and unchanged kcat in the hydrolysis of N-acetyl mono(amino acid) amide substrates. However, the catalyzed hydrolyses of N-acetyl dipeptide amide substrates by (methionine sulfoxide)-192-delta-chymotrypsin (MS-delta-Cht) shows a four- to fivefold decrease in kcat and unchanged Km(app) with respect to delta-chymotrypsin. Hydrolysis of alpha-casein by MS-delta-Cht shows a similar 4.2-fold decrease in kcat. These results imply that the Met-192 acts differently with substrates that bind only in the primary, S1, binding site (i.e., AcPheNH2) from those that bind to more extended regions of the enzyme active site. In the binding of c+AcPheNH2 and AcTrpNH2, the results support a mechanism in which the Met-192 acts to slow the rate of sustrate dissociation from the Michaelis complex to free substrate and enzyme. This is in agreement with the x-ray crystallographic structure of dioxane inhibited alpha-chymotrypsin (Steitz, T., et al. (1969), J. Mol. Biol. 46, 337). However, this mechanism is not apparent when peptide and protein substrates bind. The decrease in kcat on Met-192 modification of approximately fivefold in the hydrolysis of polypeptide substrates show a small, but significant, catalytic contribution of the Met-192 toward the lowering of the energy of activation polypeptide substrate hydrolysis by chymotrypsin. This may support the crystallographic model of Fersht et al. (Fersht, A., et al. (1973), Biochemistry 12, 2035) in which it is proposed that the Met-192 participates in the distortion of bound polypeptide substrates toward the reaction transition-state configuration and, thus, plays a role in catalysis. However, if this mechanism occurs, the effect is small, only contributing about 1 kcal/mol to the lowering of the reaction activation energy.  相似文献   

18.
Can the coupling effect among different amino acid components be used to improve the prediction of protein structural classes? The answer is yes according to the study by Chou and Zhang (Crit. Rev. Biochem. Mol. Biol. 30:275-349, 1995), but a completely opposite conclusion was drawn by Eisenhaber et al. when using a different dataset constructed by themselves (Proteins 25:169-179, 1996). To resolve such a perplexing problem, predictions were performed by various approaches for the datasets from an objective database, the SCOP database (Murzin, Brenner, Hubbard, and Chothia. J. Mol. Biol. 247:536-540, 1995). According to SCOP, the classification of structural classes for protein domains is based on the evolutionary relationship and on the principles that govern the 3D structure of proteins, and hence is more natural and reliable. The results from both resubstitution tests and jackknife tests indicate that the overall rates of correct prediction by the algorithm incorporated with the coupling effect among different amino acid components are significantly higher than those by the algorithms without using such an effect. It is elucidated through an analysis that the main reasons for Eisenhaber et al. to have reached an opposite conclusion are the result of (1) misusing the component-coupled algorithm, and (2) using a conceptually incorrect rule to classify protein structural classes. The formulation and analysis presented in this article are conducive to clarify these problems, helping correctly to apply the prediction algorithm and interpret the results.  相似文献   

19.
Reports an error in the original article by T. A. Goodall and W. K. Halford (Health Psychology, 1991, Vol 10[1], 1–8). On Page 2, the citation to Johnson et al (1986) should be deleted from the following statement: "However, most studies (e.g., Carney, Schechter, & Davis 1983; Christenson et al., 1983; Johnson, Silverstein, Rosenbloom, Carter, & Cunningham, 1986) have operationalized self-management as a single global index of compliance to treatment.' The Johnson et al study used a series of summary measures. (The following abstract of this article originally appeared in record 1991-25029-001.) Reviews the determinants of effective self-management and the methods of promoting better self-management. Demographic variables have been thought to affect self-management, but evidence suggests they have little impact. The important determinants of self-management are transient situational factors such as psychological stress and social support. Interventions to promote better self-management have reported initial improvements in blood glucose control, but the long-term effects are unclear. Self-management has been inadequately assessed and attempts to improve self-management have relied excessively on providing information… (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

20.
To elucidate the minimum requirement of amino acid residues for the active center in human adenylate kinase (hAK1), we carried out random site-directed mutagenesis of key lysine residues (K9, K21, K27, K31, K63, K131, and K194), which were conserved in mammalian AK1 species, with the pMEX8-hAK1 plasmid [Ayabe, T., et al. (1996) Biochem. Mol. Biol. Int. 38, 373-381]. Twenty different mutants were obtained and analyzed by steady-state kinetics, and all mutants showed activity loss by Km and/or k(cat) effects on MgATP2-, AMP2-, or both. The results have led to the following conclusions. (1) Lys9 would appear to interact with both MgATP2- and AMP2- but to a larger extent than with AMP2-. (2) Lys21 is likely to play a role in substrate binding of both MgATP2- and AMP2- but more strongly affects MgATP2-. (3) Lys27 and Lys131 would appear to play a functional role in catalysis by interacting strongly with MgATP2-. (4) Lys31 would appear to interact with MgATP2- and AMP2- at the MgATP2- site. (5) Lys63 would be more likely to interact with MgATP2- than with AMP2-. (6) Lys194 in the flanking C-terminal domain would appear to interact not only with MgATP2- but also with AMP2- at the MgATP2- site by stabilizing substrate binding. The loss of the positively charged epsilon-amino group of lysine affects both the affinity for the substrate and the catalytic efficiency. Hence, hydrophilic lysine residues in hAK1 would appear to be essential for substrate-enzyme interaction with the coordination of some arginine residues, reported previously [Kim, H. J., et al. (1990) Biochemistry 29, 1107-1111].  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