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 484 毫秒
1.
To determine the interaction site for cytochrome c (Cc) on cytochrome c oxidase (CcO), a number of conserved carboxyl residues in subunit II of Rhodobacter sphaeroides CcO were mutated to neutral forms. A highly conserved tryptophan, Trp(143), was also mutated to phenylalanine and alanine. Spectroscopic and metal analyses of the surface carboxyl mutants revealed no overall structural changes. The double mutants D188Q/E189N and D151Q/E152N exhibit similar steady-state kinetic behavior as wild-type oxidase with horse Cc and R. sphaeroides Cc(2), showing that these residues are not involved in Cc binding. The single mutants E148Q, E157Q, D195N, and D214N have decreased activities and increased K(m) values, indicating they contribute to the Cc:CcO interface. However, their reactions with horse and R. sphaeroides Cc are different, as expected from the different distribution of surface lysines on these cytochromes c. Mutations at Trp(143) severely inhibit activity without changing the K(m) for Cc or disturbing the adjacent Cu(A) center. From these data, we identify a Cc binding area on CcO with Trp(143) and Asp(214) close to the site of electron transfer and Glu(148), Glu(157), and Asp(195) providing electrostatic guidance. The results are completely consistent with time-resolved kinetic measurements (Wang, K., Zhen, Y., Sadoski, R., Grinnell, S., Geren, L., Ferguson-Miller, S., Durham, B., and Millett, F. (1999) J. Biol. Chem. 274, 38042-38050) and computational docking analysis (Roberts, V. A., and Pique, M. E. (1999) J. Biol. Chem. 274, 38051-38060).  相似文献   

2.
We studied the effects of mutations of positively charged amino acid residues in the pore of X. tropicalis TMEM16A calcium-activated chloride channels: K613E, K628E, K630E; R646E and R761E. The activation and deactivation kinetics were not affected, and only K613E showed a lower current density. K628E and R761E affect anion selectivity without affecting Na+ permeation, whereas K613E, R646E and the double mutant K613E + R646E affect anion selectivity and permeability to Na+. Furthermore, altered blockade by the chloride channel blockers anthracene-9-carboxylic acid (A-9-C), 4, 4'-Diisothiocyano-2,2'-stilbenedisulfonic acid (DIDS) and T16inh-A01 was observed. These results suggest the existence of 2 binding sites for anions within the pore at electrical distances of 0.3 and 0.5. These sites are also relevant for anion permeation and blockade.  相似文献   

3.
We studied the effects of mutations of positively charged amino acid residues in the pore of X. tropicalis TMEM16A calcium-activated chloride channels: K613E, K628E, K630E; R646E and R761E. The activation and deactivation kinetics were not affected, and only K613E showed a lower current density. K628E and R761E affect anion selectivity without affecting Na+ permeation, whereas K613E, R646E and the double mutant K613E + R646E affect anion selectivity and permeability to Na+. Furthermore, altered blockade by the chloride channel blockers anthracene-9-carboxylic acid (A-9-C), 4, 4''-Diisothiocyano-2,2''-stilbenedisulfonic acid (DIDS) and T16inh-A01 was observed. These results suggest the existence of 2 binding sites for anions within the pore at electrical distances of 0.3 and 0.5. These sites are also relevant for anion permeation and blockade.  相似文献   

4.
Lehoux IE  Mitra B 《Biochemistry》2000,39(33):10055-10065
(S)-Mandelate dehydrogenase from Pseudomonas putida is an FMN-dependent alpha-hydroxy acid dehydrogenase. Structural studies of two homologous enzymes, glycolate oxidase and flavocytochrome b(2), indicated that a conserved arginine residue (R277 in MDH) interacts with the product carboxylate group [Lindqvist, Y., Branden, C.-I., Mathews, F. S., and Lederer, F. (1991) J. Biol. Chem. 266, 3198-3207]. The catalytic role of R277 was investigated by site-specific mutagenesis together with chemical rescue experiments. The R277K, R277G, R277H, and R277L proteins were generated and purified in active forms. The k(cat) for the charge-conserved mutation, R277K, was only 4-fold lower than wt-MDH, but its K(m) value was 40-fold lower; in contrast, k(cat)s for R277G, R277H, and R277L were 400-1000-fold lower than for wt-MDH and K(m) values were 5-15-fold lower compared to R277K. The K(d)s for negatively charged competitive inhibitors were relatively unaffected in all four R277 mutants. The k(cat) for R277G could be enhanced by the addition of exogenous guanidines or imidazoles; the maximum rescued k(cat) was approximately 70% of the wt-MDH value. Only reagents that were positively charged and could function as hydrogen bond donors were effective rescue agents. Our results indicate that R277 plays a major role in transition state stabilization through its positive charge-consistent with a mechanism involving a carbanion intermediate. The positive charge has a relatively small contribution toward substrate binding. R277 also forms a specific hydrogen bond with both the substrate and the transition state; this interaction contributes significantly to the low K(m) for (S)-mandelate.  相似文献   

5.
In the crystal structure of a calmodulin (CaM)-bound FMN domain of human inducible nitric oxide synthase (NOS), the CaM-binding region together with CaM forms a hinge, and pivots on an R536(NOS)/E47(CaM) pair (Xia et al. J Biol Chem 284:30708-30717, 2009). Notably, isoform-specific human inducible NOS S562 and C563 residues form hydrogen bonds with the R536 residue through their backbone oxygens. In this study, we investigated the roles of the S562 and C563 residues in the NOS FMN-heme interdomain electron transfer (IET), the rates of which can be used to probe the interdomain FMN/heme alignment. Human inducible NOS S562K and C563R mutants of an oxygenase/FMN (oxyFMN) construct were made by introducing charged residues at these sites as found in human neuronal NOS and endothelial NOS isoforms, respectively. The IET rate constant of the S562K mutant is notably decreased by one third, and its flavin fluorescence intensity per micromole per liter is diminished by approximately 24?%. These results suggest that a positive charge at position 562 destabilizes the hydrogen-bond-mediated NOS/CaM alignment, resulting in slower FMN-heme IET in the mutant. On the other hand, the IET rate constant of the C563R mutant is similar to that of the wild-type, indicating that the mutational effect is site-specific. Moreover, the human inducible NOS oxyFMN R536E mutant was constructed to disrupt the bridging CaM/NOS interaction, and its FMN-heme IET rate was decreased by 96?%. These results demonstrated a new role of the isoform-specific serine residue of the key CaM/FMN(NOS) bridging site in regulating the FMN-heme IET (possibly by tuning the alignment of the FMN and heme domains).  相似文献   

6.
Recent studies have shown that selection of proteins for degradation by the ubiquitin system occurs most probably by binding to specific sites of the ubiquitin-protein ligase, E3. A free alpha-NH2 residue of the substrate is one important determinant recognized by the ligase. Selective binding sites have been described for basic and bulky-hydrophobic NH2 termini (Reiss, Y., Kaim, D., and Hershko, A. (1988) J. Biol. Chem. 263, 2693-2698) and for alanine, serine, and threonine at the NH2-terminal position (Gonda, D. K., Bachmair, A., Wünning, I., Tobias, J. W., Lane, W. S., and Varshavsky, A. (1989) J. Biol. Chem. 264, 16700-16712). Proteins with acidic NH2-terminal residues are degraded by the ubiquitin system only following conversion of the acidic residue to a basic residue by the addition of an arginine moiety (Ferber, S., and Ciechanover, A. (1987) Nature 326, 808-811). Although the enzymes involved in this post-translational modification have been characterized, the underlying mechanism has been obscure. By using a chemical cross-linking technique, we demonstrate that proteins with acidic NH2 termini do not bind to E3 without prior modification of this residue by the addition of arginine. In contrast, proteins with a basic NH2-terminal residue bind to the ligase without any modification. The recognition of acidic NH2-terminal substrates by E3 is dependent upon the addition of all the components of the modifying machinery, arginyl-tRNA-protein transferase, arginyl-tRNA synthetase, tRNA, and arginine. The ligase-bound modified proteins are converted to ubiquitin conjugates in a "pulse-chase" experiment, indicating that the binding is functional and that the enzyme-substrate complex is an obligatory intermediate in the conjugation process. Chemical modification of the carboxyl groups, which results in their neutralization, generates substrates that bind to E3 without modification. This finding suggests that the amino-terminal binding site of E3 is negatively charged, and only positively charged amino-terminal residues may bind to it. Negatively charged (acidic) NH2-terminal residues will bind only following neutralization or reversal of the charge.  相似文献   

7.
M Inoue  H Yamada  T Yasukochi  T Miki  T Horiuchi  T Imoto 《Biochemistry》1992,31(42):10322-10330
The "right-sided" and "left-sided" substrate binding modes at the lower saccharide binding subsites (D-F sites) of chicken lysozyme were investigated by utilizing mutant lysozymes secreted from yeast. We constructed the following mutant lysozymes; "left-sided" substitution of Asn46 to Asp, deletion of Thr47, and insertion of Gly between Thr47 and Asp48 and "right-sided" substitution of Asn37 to Gly. Analyses of their activities and substrate binding abilities showed that Asn46 and Thr47 are involved in the initial enzyme-substrate complex and Asn37 is involved in the transition state. These results support an earlier proposal that interactions between substrate and residues at the left side of lysozyme stabilize a catalytically inactive enzyme-substrate complex, while interactions between substrate and residues at the right side stabilize the catalytically active complex [Pincus, M. R., & Scheraga, H. A. (1979) Macromolecules 12, 633-644]. These results are also consistent with the proposed kinetic mechanism for lysozyme reaction that the rearrangement of an initial enzyme-substrate complex (beta-complex) to another complex (gamma-complex) is required for catalytic hydrolysis [Banerjee S. K., Holler, E., Hess, G. P., & Rupley, J. A. (1975) J. Biol. Chem. 250, 4355-4367].  相似文献   

8.
The high-potential iron-sulfur protein (HiPIP) from Rhodospirillum tenue (strain 3761) shows only a weak (20-25%) sequence similarity to HiPIPs from Chromatium vinosum, Ectothiorhodospira halophila and Ectothiorhodospira vacuolata, including the strict conservation of only two of the twelve residues assumed to be in the 4Fe-4S cluster packing region [Tedro, S. M., Meyer, T. E. and Kamen, M. D. (1979) J. Biol. Chem. 254, 1495-1500]. In spite of these differences, the general range and distribution of hyperfine-shifted 1H-NMR peaks of oxidized and reduced R. tenue HiPIP resemble those of E. halophila HiPIP I [Krishnamoorthi, R., Markley, J. L., Cusanovich, M. A., Pryzycieki, C. T. and Meyer, T. E. (1986) Biochemistry 25, 60-67]. Temperature- and pH-dependence and longitudinal relaxation behavior were determined for hyperfine-shifted peaks of the oxidized protein. Tentative assignments of peaks to ligands and aromatic residues suggest the presence of common apoprotein-active-site interactions in these proteins. Differences occur in the pattern of paramagnetically shifted peaks attributed to hydrogens bonded to the 4Fe-4S cluster. Hyperfine-shifted peaks of R. tenue HiPIP are not perturbed by pH changes in the range 5-9. In contrast, those of the C. vinosum protein exhibit a pH-dependence of chemical shifts that has been attributed to the titration of His42 [Nettesheim, D. G., Meyer, T. E., Feinberg, B. A. and Otvos, J. D. (1983) J. Biol. Chem. 258, 8235-8239]. Since R. tenue HiPIP contains no histidine, the present observation confirms the above hypothesis.  相似文献   

9.
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   

10.
11.
A C Anusiem  M Kelleher 《Biopolymers》1984,23(7):1147-1167
Interest in the thermodynamics of the iron-binding site in hemoproteins has increased in recent years due to refinements in x-ray crystallographic studies of hemoproteins [see Deathage, J. F., Lee, R. S., Anderson, C. M. & Moffat, K. (1976) J. Mol. Biol. 104 , 687–706; Heidner, E. J., Ladner, R. C. & Perutz, M. F. (1976) J. Mol. Biol. 104 , 707–722; Deathage, J. F., Lee, R. S. & Moffat, K. (1976) J. Mol. Biol. 104 , 723–728; Ladner, R. C., Heidner, E. J. & Perutz, M. F. (1976) J. Mol. Biol. 114 , 385–414; Fermi, G. & Perutz, M. F. (1977) J. Mol. Biol. 114 , 421–431; Takano, T. (1977) J. Mol. Biol. 110 , 537–568 and 569–589], the synthesis and x-ray analysis of model heme compounds [see Scheidt, W. R. (1977) Acc. Chem. Res. 10 , 339–345; Kastner, M. E., Scheidt, W. R., Mashino, T. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 666–667; Mashiko, T., Kastner, M. E., Spartalian, K., Scheidt, W. R. & Reed, C. A. (1978) J. Am. Chem. Soc. 100 , 6354–6362; Hill, H. A. O., Skite, P. P., Buchler, J. W., Luchr, H., Tonn, M., Gregson, A. K. & Pellizer, G. (1979) Chem. Commun. 4 , 151–152; and Scheidt, W. R., Cohen, I. A. & Kastner, M. E. (1979) Biochemistry 18 , 3546–3556], and the numerous data on heme–protein interactions that account for the differences observed in ligand binding between the various species of animals. Numerous probes have been used and provide information about the structure and thermodynamics of the binding site, but no single probe can provide the complete picture [see Iizuka, T. & Yonetani, T. (1970) Adv. Biophys. 1 , 157–182; Smith, D. W. & Williams, R. J. P. (1970) Struct. Bond. 7 , 1–45; and Spiro, T. G. (1975) Biochim. Biophys. Acta 416 , 169–189].  相似文献   

12.
Analogues of a synthetic heptapeptide substrate corresponding to the sequence around a phosphorylation site in histone H2B [Glass, D. B. & Krebs, E. G. (1982) J. Biol. Chem. 257, 1196-1200] were used to assess interactions between the peptide substrate and the ATP binding sites of cGMP-dependent protein kinase and the catalytic subunit of cAMP-dependent protein kinase. The affinity of each protein kinase for lin-benzo-ADP was determined in the absence and presence of substrate peptide by fluorescence anisotropy titrations [Bhatnagar, D., Roskoski, R., Jr., Rosendahl, M. S., & Leonard, N. J. (1983) Biochemistry 22, 6310-6317]. The Kd values of cGMP-dependent protein kinase for lin-benzo-ADP in the absence and presence of cGMP were 7.6 and 9.7 microM, respectively. Histone H2B(29-35) (Arg-Lys-Arg-Ser-Arg-Lys-Glu) had no effect on nucleotide affinity in either the absence or presence of cGMP. However, when lysine-34 located two residues after the phosphorylatable serine is replaced with an alanyl residue, the resulting [Ala34]histone H2B(29-35) and its analogue peptides interact with cGMP-dependent protein kinase and/or the nucleotide in a fashion that decreases nucleotide binding affinity approximately 3-fold. This amino acid replacement had previously been shown to cause an increase in Vmax and a decrease in the pH optimum for the phosphotransferase reaction. Replacement of positively charged residues at positions 30 and 31 of the peptide also decreased nucleotide affinity. Other analogues of histone H2B(29-35) failed to affect binding of lin-benzo-ADP to the active site of the cGMP-dependent enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Synthetic peptides corresponding to sequence segments of the nicotinic acetylcholine receptor (nAChR) alpha subunits have been used to identify regions that contribute to formation of the binding sites for cholinergic ligands. We have previously defined alpha-bungarotoxin (alpha-BTX) binding sequences between residues 180 and 199 of a putative rat neuronal nAChR alpha subunit, designated alpha 5 [McLane, K. E., Wu, X., & Conti-Tronconi, B. M. (1990) J. Biol. Chem. 265, 9816-9824], and between residues 181 and 200 of the chick neuronal alpha 7 and alpha 8 subunits [McLane, K. E., Wu, X., Schoepfer, R., Lindstrom, J., & Conti-Tronconi, B. M. (1991) J. Biol. Chem. (in press)]. These sequences are relatively divergent compared with the Torpedo and muscle nAChR alpha 1 alpha-BTX binding sites, which indicates a serious limitation of predicting functional domains of proteins based on homology in general. Given the highly divergent nature of the alpha 5 sequence, we were interested in determining the critical amino acid residues for alpha-BTX binding. In the present study, the effects of single amino acid substitutions of Gly or Ala for each residue of the rat alpha 5(180-199) sequence were tested, using a competition assay, in which peptides compete for 125I-alpha-BTX binding with native Torpedo nAChR.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Long chain curarimimetic toxins from snake venom bind with high affinities to both muscular type nicotinic acetylcholine receptors (AChRs) (K(d) in the pm range) and neuronal alpha 7-AChRs (K(d) in the nm range). To understand the molecular basis of this dual function, we submitted alpha-cobratoxin (alpha-Cbtx), a typical long chain curarimimetic toxin, to an extensive mutational analysis. By exploring 36 toxin mutants, we found that Trp-25, Asp-27, Phe-29, Arg-33, Arg-36, and Phe-65 are involved in binding to both neuronal and Torpedo (Antil, S., Servent, D., and Ménez, A. (1999) J. Biol. Chem. 274, 34851-34858) AChRs and that some of them (Trp-25, Asp-27, and Arg-33) have similar binding energy contributions for the two receptors. In contrast, Ala-28, Lys-35, and Cys-26-Cys-30 selectively bind to the alpha 7-AChR, whereas Lys-23 and Lys-49 bind solely to the Torpedo AChR. Therefore, alpha-Cbtx binds to two AChR subtypes using both common and specific residues. Double mutant cycle analyses suggested that Arg-33 in alpha-Cbtx is close to Tyr-187 and Pro-193 in the alpha 7 receptor. Since Arg-33 of another curarimimetic toxin is close to the homologous alpha Tyr-190 of the muscular receptor (Ackermann, E. J., Ang, E. T. H., Kanter, J. R., Tsigelny, I., and Taylor, P. (1998) J. Biol. Chem. 273, 10958-10964), toxin binding probably occurs in homologous regions of neuronal and muscular AChRs. However, no coupling was seen between alpha-Cbtx Arg-33 and alpha 7 receptor Trp-54, Leu-118, and Asp-163, in contrast to what was observed in a homologous situation involving another toxin and a muscular receptor (Osaka, H., Malany, S., Molles, B. E., Sine, S. M., and Taylor, P. (2000) J. Biol. Chem. 275, 5478-5484). Therefore, although occurring in homologous regions, the detailed modes of toxin binding to alpha 7 and muscular receptors are likely to be different. These data offer a molecular basis for the design of toxins with predetermined specificities for various members of the AChR family.  相似文献   

15.
Y Snitko  S K Han  B I Lee  W Cho 《Biochemistry》1999,38(24):7803-7810
To identify the residues essential for interfacial binding and substrate binding of human pancreatic phospholipase A2 (hpPLA2), several ionic residues in the putative interfacial binding surface (R6E, K7E, K10E, and K116E) and substrate binding site (D53K and K56E) were mutated. Interfacial affinity of these mutants was measured using anionic polymerized liposomes, and their enzymatic activity was measured using various substrates including phospholipid monomers, zwitterionic and anionic micelles, and anionic polymerized mixed liposomes. Similar mutations (R6E, K10E, K56E, and K116E) were made to porcine pancreatic phospholipase A2 (ppPLA2), and the properties of mutants were measured by the same methods. Results indicate that hpPLA2 and ppPLA2 have similar interfacial binding mechanisms in which cationic residues in the amino terminus and Lys-116 in the carboxy terminus are involved in binding to anionic lipid surfaces. Small but definite differences between the two enzymes were observed in overall interfacial affinity and activity and the effects of the mutations on interfacial enzyme activity. The interfacial binding of hpPLA2 and ppPLA2 is distinct from that of bovine pancreatic phospholipase A2 in that Lys-56 is involved in the interfacial binding of the latter enzyme. The unique phospholipid headgroup specificity of hpPLA2 derives from the presence of Asp-53 in the substrate binding site. This residue appears to participate in stabilizing electrostatic interactions with the cationic ethanolamine headgroup, hence the phosphatidylethanolamine preference of hpPLA2. Taken together, these studies reveal the similarities and the differences in the mechanisms by which mammalian pancreatic phospholipases A2 interact with lipid aggregates and perform interfacial catalysis.  相似文献   

16.
The binding of Ca2+(4).calmodulin (CaM) to rabbit skeletal muscle myosin light chain kinase (MLCK) is required for expression of the enzyme's activity. While both MLCK and CaM were stable at 30 degrees C, their complex was not. The binding of CaM to MLCK resulted in a time- and temperature-dependent inactivation that reflected an intrinsic instability of the complex. Separation of the components of the inactive complex yielded functional CaM, but catalytically inert MLCK, indicating that the site of the inactivating event was confined to MLCK. The behavior of proteolytic fragments further localized this event to the C-terminal 60% of the 603-residue protein. Changes in the tryptophan fluorescence and proteolytic susceptibility of MLCK-CaM indicated that a conformational change accompanied, and thus may have caused, inactivation. Substrates protected against inactivation, as did millimolar concentrations of Mg2+, Mn2+, and Ca2+. These metals appeared to bind to a site on MLCK distinct from that which recognized Mg2+.ATP. A proteolytic fragment of MLCK lacking the ability to bind CaM, C beta 35 (residues 255-584; Edelman, A. M., Takio, K., Blumenthal, D. K., Hansen, R. S., Walsh, K. A., Titani, K., and Krebs, E. G. (1985) J. Biol. Chem. 260, 11275-11285), was unstable at 30 degrees C, whereas a similar fragment which does bind CaM, T beta 40 (residues 236-595; Edelman, A. M., Takio, K., Blumenthal, D. K., Hansen, R. S., Walsh, K. A., Titani, K., and Krebs, E. (1985) J. Biol. Chem. 260, 11275-11285), was unstable only when CaM was bound.  相似文献   

17.
The substrate specificity of bovine brain myelin basic protein (MBP)-specific protein methylase I (S-adenosyl-L-methionine:protein-L-arginine N-methyltransferase, EC 2.1.1.23), which methylates arginine residues of protein, has been studied using various MBPs, several synthetic peptides and heterogeneous nuclear ribonucleoprotein complex protein (hnRNP). (1) Among MBPs from different species of brain, the carp MBP was found to be the best substrate for MBP-specific protein methylase I. This high degree of methyl acceptability is most likely due to the fact that carp MBP is not in vivo methylated at the arginine residue (Deibler, G.E. and Martenson, R.E. (1973) J. Biol. Chem. 248, 2387-2391) and that the methylatable amino acid sequence is present in this protein. (2) In order to study the minimum chain length of MBP polypeptide which functions as the methyl acceptor, several synthetic polypeptides whose sequences are identical to the region surrounding the residue 107 of bovine MBP (the in vivo methylation site) were synthesized. It was found that the hexapeptide, Gly-Lys-Gly-Arg-Gly-Leu (corresponding to residues 104-109 of bovine MBP), was the shortest methyl accepting peptide, while the tetrapeptide, Gly-Arg-Gly-Leu (corresponding to residues 106-109) was inactive as a substrate. (3) hnRNP protein is known to contain methylarginine at residue 193 (Williams, K.R., Stone, K.L., LoPresti, M.B., Merrill, B.M. and Plank, S.R. (1985) Proc. Natl. Acad. Sci. USA 82, 5666-5670) which is post-translationally modified. Thus, the RNP protein overproduced in Escherichia coli and therefore did not contain methylarginine was examined for its methyl acceptability. It was found that neither MBP-specific nor histone-specific protein methylase I could methylate this methylarginine-less RNP protein. This suggests a possible existence of a distinct protein methylase I specific for this nuclear protein.  相似文献   

18.
Conventional kinesin is a highly processive, microtubule-based motor protein that drives the movement of membranous organelles in neurons. Using in vivo genetics in Drosophila melanogaster, Glu164 was identified as an amino acid critical for kinesin function [Brendza, K. M., Rose, D. J., Gilbert, S. P., and Saxton, W. M. (1999) J. Biol. Chem. 274, 31506-31514]. Glu164 is located at the beta-strand 5a/loop 8b junction of the catalytic core and projects toward the microtubule binding face in close proximity to key residues on beta-tubulin helix alpha12. Substitution of Glu(164) with alanine (E164A) results in a dimeric kinesin with a dramatic reduction in the microtubule-activated steady-state ATPase (5 s(-1) per site versus 22 s(-1) per site for wild-type). Our analysis shows that E164A binds ATP and microtubules with a higher affinity than wild-type kinesin. The rapid quench and stopped-flow results provide evidence that ATP hydrolysis is significantly faster and the precise coordination between the motor domains is disrupted. The data reveal an E164A intermediate that is stalled on the microtubule and cannot bind and hydrolyze ATP at the second head.  相似文献   

19.
Positively charged plastocyanin from Anabaena sp. PCC 7119 was investigated by site-directed mutagenesis. The reactivity of its mutants toward photosystem I was analyzed by laser flash spectroscopy. Replacement of arginine at position 88, which is adjacent to the copper ligand His-87, by glutamine and, in particular, by glutamate makes plastocyanin reduce its availability for transferring electrons to photosystem I. Such a residue in the copper protein thus appears to be isofunctional with Arg-64 (which is close to the heme group) in cytochrome c(6) from Anabaena (Molina-Heredia, F. P., Diaz-Quintana, A., Hervás, M., Navarro, J. A., and De la Rosa, M. A. (1999) J. Biol. Chem. 274, 33565-33570) and Synechocystis (De la Cerda, B., Diaz-Quintana, A., Navarro, J. A. , Hervás, M., and De la Rosa, M. A. (1999) J. Biol. Chem. 274, 13292-13297). Other mutations concern specific residues of plastocyanin either at its positively charged east face (D49K, H57A, H57E, K58A, K58E, Y83A, and Y83F) or at its north hydrophobic pole (L12A, K33A, and K33E). Mutations altering the surface electrostatic potential distribution allow the copper protein to modulate its kinetic efficiency: the more positively charged the interaction site, the higher the rate constant. Whereas replacement of Tyr-83 by either alanine or phenylalanine has no effect on the kinetics of photosystem I reduction, Leu-12 and Lys-33 are essential for the reactivity of plastocyanin.  相似文献   

20.
Huang S  Vandenberg RJ 《Biochemistry》2007,46(34):9685-9692
L-Glutamate is the predominant excitatory neurotransmitter in the brain, and its extracellular concentration is tightly controlled by the excitatory amino acid transporters (EAATs). The transport of 1 glutamate molecule is coupled to the cotransport of 3 Na+ and 1 H+ and the countertransport of 1 K+. In addition to substrate transport, the binding of glutamate and Na+ activates an anion current which is thermodynamically uncoupled from the transport process. We have identified three amino acid residues in EAAT1 (D272 in TM5, K384 and R385 in TM7) that influence the amplitude of the anion channel current relative to the transport current. Transporters containing the mutations R268A, D272A, D272K, K384A, K384D, R385A, and R385D were expressed in Xenopus laevis oocytes and their transport and anion channel functions measured using the two-electrode voltage clamp techniques. The D272, K384, and R385 mutant transporters showed no change in transport properties but have increased levels of anion channel activity compared to wild-type transporters. These results identify additional residues of the EAAT1 transporter that may contribute to the gating mechanism of the anion channel of glutamate transporters and also provide hints as to how substrate binding leads to channel activation.  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