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 125 毫秒
1.
The structures of two catalytically defective semi-synthetic RNases obtained by replacing aspartic acid 121 with asparagine or alanine have been determined and refined at a resolution of 2.0 A (R = 0.186 and 0.172, respectively). When these structures are compared with the refined 1.8-A structure (R = 0.204) of the fully active aspartic acid-containing enzyme (Martin, P.D., Doscher, M.S., and Edwards, B. F. P. (1987) J. Biol. Chem. 262, 15930-15938), numerous and widespread changes, much greater in number and magnitude than the small structural variations noted previously between the semisynthetic complex and RNase A, are found to have occurred. These changes include the movement of the loop containing residues 65-72 away from the active site, a more or less generalized relocation of crystallographically bound water molecules, and a number of rearrangements in the hydrogen bonding network at the active site. Most changes are far removed from the immediate site of the modifications and are distributed essentially throughout the molecule. The details of many of these changes are unique to each analog. In the asparagine analog, a destabilization in the positioning of active site residue His-119 also appears to have occurred.  相似文献   

2.
The troponin I peptide N alpha-acetyl TnI (104-115) amide (TnIp) represents the minimum sequence necessary for inhibition of actomyosin ATPase activity of skeletal muscle (Talbot, J.A. & Hodges, R.S. 1981, J. Biol. Chem. 256, 2798-3802; Van Eyk, J.E. & Hodges, R.S., 1988, J. Biol. Chem. 263, 1726-1732; Van Eyk, J.E., Kay, C.M., & Hodges, R.S., 1991, Biochemistry 30, 9974-9981). In this study, we have used 1H NMR spectroscopy to compare the binding of this inhibitory TnI peptide to a synthetic peptide heterodimer representing site III and site IV of the C-terminal domain of troponin C (TnC) and to calcium-saturated skeletal TnC. The residues whose 1H NMR chemical shifts are perturbed upon TnIp binding are the same in both the site III/site IV heterodimer and TnC. These residues include F102, I104, F112, I113, I121, I149, D150, F151, and F154, which are all found in the C-terminal domain hydrophobic pocket and antiparallel beta-sheet region of the synthetic site III/site IV heterodimer and of TnC. Further, the affinity of TnIp binding to the heterodimer (Kd = 192 +/- 37 microM) was found to be similar to TnIp binding to TnC (48 +/- 18 microM [Campbell, A.P., Cachia, P.J., & Sykes, B.D., 1991, Biochem. Cell Biol. 69, 674-681]). The results indicate that binding of the inhibitory region of TnI is primarily to the C-terminal domain of TnC. The results also indicate how well the synthetic peptide heterodimer mimics the C-terminal domain of TnC in structure and functional interactions.  相似文献   

3.
The cholecystokinin (CCK) receptor-1 (CCK1R) is a G protein-coupled receptor, which mediates important central and peripheral cholecystokinin actions. Our aim was to progress in mapping of the CCK1R binding site by identifying residues that interact with the methionine and phenylalanine residues of the C-terminal moiety of CCK because these are crucial for its binding and biological activity, and to determine whether CCK and the selective non-peptide agonist, SR-146,131, share a common binding site. Identification of putative amino acids of the CCK1R binding site was achieved by dynamics-based docking of the ligand CCK in a refined three-dimensional model of the CCK1R using, as constraints, previous results that identified contact points between residues of CCK and CCK1R (Kennedy, K., Gigoux, V., Escrieut, C., Maigret, B., Martinez, J., Moroder, L., Frehel, D., Gully, D., Vaysse, N., and Fourmy, D. (1997) J. Biol. Chem. 272, 2920-2926 and Gigoux, V., Escrieut, C., Fehrentz, J. A., Poirot, S., Maigret, B., Moroder, L., Gully, D., Martinez, J., Vaysse, N., and Fourmy, D. (1999) J. Biol. Chem. 274, 20457-20464). By this approach, a series of residues forming connected hydrophobic clusters were identified. Pharmacological and functional analysis of mutated receptors indicated that a network of hydrophobic residues including Cys-94, Met-121, Val-125, Phe-218, Ile-329, Phe-330, Trp-326, Ile-352, Leu-356, and Tyr-360, is involved in the binding site for CCK and in the activation process of the CCK1R. Within this hydrophobic network, the physico-chemical nature of residue 121 seems to be essential for CCK1R functioning. Finally, the biological properties of mutants together with dynamic docking of SR-146,131 in the CCK1R binding site demonstrated that SR-146,131 occupies a region of CCK1R binding site which interacts with the C-terminal amidated tripeptide of CCK, i.e. Met-Asp-Phe-NH(2). These new and important insights will serve to better understand the activation process of CCK1R and to design or optimize ligands.  相似文献   

4.
The hydrolysis of fluorotyrosine phosphate by the calmodulin-activated phosphatase calcineurin has been monitored by 19F nuclear magnetic resonance spectroscopy. Previous work had established that the 19F nuclear magnetic resonance shift of the fluorine nucleus was altered after the phosphorylation of the phenolic hydroxyl group (B. Martin, C.J. Pallen, J.H. Wang, and D.J. Graves (1985) J. Biol. Chem. 260, 14592-14597). The disappearance of substrate and the appearance of product can be measured simultaneously with this approach. Application of the integrated form of the Michaelis-Menten equation yields estimates of the kinetic parameter, KM, close to the values obtained by initial rate kinetics. The velocity term, VM, was also evaluated to be approximately the same value. Calcineurin was determined not to be inactivated over the time period of the reaction. The results demonstrate that 19F nuclear magnetic resonance spectroscopy can be applied to the examination of enzyme-catalyzed reactions.  相似文献   

5.
Evidence that a 27-residue sequence is the actin-binding site of ABP-120   总被引:12,自引:0,他引:12  
Proteolysis experiments of ABP-120 from Dictyostelium discoideum have previously demonstrated that removal of residues 89-115 from a tryptic peptide which retains actin binding activity, abolishes actin binding (Bresnick, A. R., Warren, V., and Condeelis, J. (1990) J. Biol. Chem. 265, 9236-9240). Antibodies made against a synthetic peptide of this 27-amino acid sequence (27-mer) specifically immunoprecipitate native ABP-120 from Dictyostelium high speed supernatants, demonstrating that the 27-mer sequence is on the surface of the molecule as expected for an active site. ABP-120 is inhibited in its binding to F-actin by Fab' fragments of the anti-27-mer IgG. Half-maximal inhibition occurs at an approximate molar ratio of 7 Fab' fragments/ABP-120 monomer. Viscoelastic measurements indicate that ABP-120 forms fewer cross-links with F-actin in the presence of the 27-mer synthetic peptide than in its absence. In F-actin cosedimentation assays, the binding of ABP-120 to actin is inhibited by the 27-mer synthetic peptide. Furthermore, the 27-mer synthetic peptide cosediments with F-actin, whereas a control hydrophobic peptide and a synthetic peptide of residues 69-88 of ABP-120 do not cosediment with F-actin. These observations suggest a direct involvement of the 27-mer sequence in the actin binding activity of ABP-120.  相似文献   

6.
The present studies demonstrate that no single stretch of sequence in the third intracellular (3i) loop of the alpha(2A) adrenergic receptor (alpha(2A)-AR) can fully account for its previously described interactions with spinophilin (Richman, J. G., Brady, A. E., Wang, Q., Hensel, J. L., Colbran, R. J., and Limbird, L. E. (2001) J. Biol. Chem. 276, 15003-15008), 14-3-3zeta (Prezeau, L., Richman, J. G., Edwards, S. W., and Limbird, L. E. (1999) J. Biol. Chem. 274, 13462-13469), and arrestin 3 (Wu, G., Krupnick, J. G., Benovic, J. L., and Lanier, S. M. (1997) J. Biol. Chem. 272, 17836-17842), suggesting that a three-dimensional surface, rather than a linear sequence, provides the basis for these interactions as proposed for 3i loop tethering of the alpha(2A)-AR to the basolateral surface of Madin-Darby canine kidney cells (Edwards, S. W., and Limbird, L. E. (1999) J. Biol. Chem. 274, 16331-16336). Sequences at the extreme N-terminal and C-terminal ends of the 3i loop are critical for interaction with spinophilin but not for interaction with 14-3-3zeta or arrestin 3, for which the C-terminal half of the loop is more important. Competition binding for (35)S-labeled alpha(2A)-AR 3i loop binding to glutathione S-transferase (GST)-spinophilin amino acids 151-444 revealed a relative affinity of spinophilin congruent with arrestin > 14-3-3zeta for the unphosphorylated alpha(2A)-AR 3i loop. Agonist occupancy of the alpha(2A)-AR increases receptor association with spinophilin, and arrestin 3 appears to compete for this enrichment. However, when the G protein-coupled receptor kinase 2 substrate sequence was deleted from the 3i loop, arrestin 3 could not compete for the agonist-enriched binding of spinophilin to the mutant alpha(2A)-AR. These data are consistent with a model where sequential or competitive interactions among spinophilin, arrestin, and/or 14-3-3zeta play a role in alpha(2A)-AR functions.  相似文献   

7.
6-Oxo camphor hydrolase (OCH) is an enzyme of the crotonase superfamily that catalyzes carbon-carbon bond cleavage in bicyclic beta-diketones via a retro-Claisen reaction (Grogan, G., Roberts, G. A., Bougioukou, D., Turner, N. J., and Flitsch, S. L. (2001) J. Biol. Chem. 276, 12565-12572). The native structure of OCH has been solved at 2.0-A resolution with selenomethionine multiple wave anomalous dispersion and refined to a final R(free) of 19.0. The structure of OCH consists of a dimer of trimers that resembles the "parent" enzyme of the superfamily, enoyl-CoA hydratase. In contrast to enoyl-CoA hydratase, however, two octahedrally coordinated sodium atoms are found at the 3-fold axis of the hexamer of OCH, and the C-terminal helix of OCH does not form a discrete domain. Models of the substrate, 6-oxo camphor, and a proposed enolate intermediate in the putative active site suggest possible mechanistic roles for Glu-244, Asp-154, His-122, His-45, and His-145.  相似文献   

8.
Arg-120 is located near the mouth of the hydrophobic channel that forms the cyclooxygenase active site of prostaglandin endoperoxide H synthases (PGHSs)-1 and -2. Replacement of Arg-120 of ovine PGHS-1 with a glutamine increases the apparent Km of PGHS-1 for arachidonate by 1,000-fold (Bhattacharyya, D. K., Lecomte, M., Rieke, C. J., Garavito, R. M., and Smith, W. L. (1996) J. Biol. Chem. 271, 2179-2184). This and other evidence indicate that the guanido group of Arg-120 forms an ionic bond with the carboxylate group of arachidonate and that this interaction is an important contributor to the overall strength of arachidonate binding to PGHS-1. In contrast, we report here that R120Q human PGHS-2 (hPGHS-2) and native hPGHS-2 have very similar kinetic properties, but R120L hPGHS-2 catalyzes the oxygenation of arachidonate inefficiently. Our data indicate that the guanido group of Arg-120 of hPGHS-2 interacts with arachidonate through a hydrogen bond rather than an ionic bond and that this interaction is much less important for arachidonate binding to PGHS-2 than to PGHS-1. The Km values of PGHS-1 and -2 for arachidonate are the same, and all but one of the core residues of the active sites of the two isozymes are identical. Thus, the results of our studies of Arg-120 of PGHS-1 and -2 imply that interactions involved in the binding of arachidonate to PGHS-1 and -2 are quite different and that residues within the hydrophobic cyclooxygenase channel must contribute more significantly to arachidonate binding to PGHS-2 than to PGHS-1. As observed previously with R120Q PGHS-1, flurbiprofen was an ineffective inhibitor of R120Q hPGHS-2. PGHS-2-specific inhibitors including NS398, DuP-697, and SC58125 had IC50 values for the R120Q mutant that were up to 1,000-fold less than those observed for native hPGHS-2; thus, the positively charged guanido group of Arg-120 interferes with the binding of these compounds. NS398 did not cause time-dependent inhibition of R120Q hPGHS-2, whereas DuP-697 and SC58125 were time-dependent inhibitors. Thus, Arg-120 is important for the time-dependent inhibition of hPGHS-2 by NS398 but not by DuP-697 or SC58125.  相似文献   

9.
The cloning of the gene for staphylococcal nuclease A in the pIN-III-OmpA secretion vector results in a hybrid protein which is processed by signal peptidase I, yielding an active form of the nuclease that is secreted across the cytoplasmic membrane (Takahara, M., Hibler, D., Barr, P. J., Gerlt, J. A., and Inouye, M. (1985) J. Biol. Chem. 260, 2670-2674). Using oligonucleotide-directed site-specific mutagenesis, we have constructed a set of mutants at the cleavage site area of the precursor hybrid protein designed to alter progressively the predicted secondary structure of the cleavage site. Our results show that processing becomes increasingly defective as the turn probability decreases. These results are consistent with the structural requirement that we found for the processing of lipoprotein by signal peptidase II (Inouye, S., Duffaud, G., and Inouye, M. (1986) J. Biol. Chem. 261, 10970-10975). We conclude that secretory precursor proteins have a distinct secondary structural requirement at their cleavage site for processing by signal peptidase I, as well as by signal peptidase II.  相似文献   

10.
The three-dimensional structure of the subtilisin from Bacillus amyloliquefaciens (BAS) has been refined to 1.8 A using the amino acid sequence deduced from the DNA coding sequence. The structure is essentially the same as the previously reported structures of subtilisin BPN' (Wright, C.S., Alden, R.A., and Kraut, J. (1969) Nature 221, 235-242) and Novo (Drenth, J., Hol, W. G. J., Jansonius, J. N., and Koekoek, R. (1972) Eur. J. Biochem. 26, 177-181) determined in different crystal forms, at 2.5 and 2.8 A resolution, respectively. The largest differences in the three crystallographic models are seen in regions where the amino acid sequence used in the fit to the electron density maps of BPN' and Novo differs from the gene sequence of BAS (Wells, J. A., Ferrari, E., Henner, D. J., Estell, D. A., and Chen, E. Y. (1983) Nucleic Acids Res. 11, 7911-7925). The refined BAS model shows new features of cation binding, hydrogen bonding, and internal solvent structure. The refined BAS model has served as a basis for the analysis of stereochemical factors involved in the peroxide inactivation of the enzyme. Methionine 222, which is adjacent to the catalytic Ser221, is quantitatively oxidized to the sulfoxide by hydrogen peroxide as had been previously shown for the related Bacillus licheniformis enzyme (Stauffer, C. E., and Etson, D. (1969) J. Biol. Chem. 244, 5333-5338). In addition to this site of modification, we observe partial to full oxidation of two of the four remaining methionines. The oxidation of the methionines does not correlate well with their solvent accessibility calculated from the x-ray structure coordinates; in addition, only one of the two possible stereoisomers of methionine sulfoxide is formed. We also detect hydrogen peroxide-induced modification of the hydroxyl groups of two tyrosines. Modeling suggests that most of the observed effect of oxidation on the enzyme's catalytic efficiency can be attributed to unfavorable interactions at the oxyanion binding site between the sulfoxide group at 222 and the carbonyl oxygen of the scissile peptide bond of the bound substrate.  相似文献   

11.
Three unique bilin peptides, a beta subunit peptide bearing a doubly linked phycourobilin (PUB), and two gamma subunit peptides with singly linked PUB groups, were obtained by enzymatic degradation of Gastroclonium coulteri R-phycoerythrin. These peptides were shown to have the sequences (Klotz, A. V., and Glazer, A. N. (1985) J. Biol. Chem. 260, 4856-4863): (Formula: see text) The sequence of peptide beta-3T was identical to that previously established for a doubly linked phycoerythrobilin (PEB) peptide derived from a B-phycoerythrin (Lundell, D. J., Glazer, A. N., DeLange, R. J., and Brown, D. M. (1984) J. Biol. Chem. 259, 5472-5480). Secondary ion mass spectrometry of beta-3T yielded a protonated molecular ion of 1629 mass units, the same as that given by the doubly linked PEB peptide (Schoenleber, R. W., Lundell, D. J., Glazer, A. N., and Rapoport, H. (1984) J. Biol. Chem. 259, 5481-5484), indicating that the doubly linked PUB and PEB tetrapyrroles were isomeric structures. High resolution 1H NMR analyses of peptides beta-3T, gamma-BV8, and gamma-DP provided unambiguous structural assignments for the singly and doubly linked PUB chromophores and indicated that the peptides in gamma-BV8 and gamma-DP were linked to ring A. The determination of which peptide fragment is linked to ring A and which to ring D in peptide beta-3T was not achieved in this study. 1H NMR analyses of three PEB-peptides from G. coulteri R-phycoerythrin--alpha-1 Cys(PEB)-Tyr-Arg, alpha-2 Leu-Cys(PEB)-Val-Pro-Arg, and beta-1 Met-Ala-Ala-Cys(PEB)-Leu-Arg--showed that they were identical to previously described corresponding chromopeptides from Porphyridium cruentum B-phycoerythrin, with the peptide linked to ring A of PEB in each instance (Schoenleber, R. W., Lundell, D. J., Glazer, A. N., and Rapoport, H. (1984) J. Biol. Chem. 259, 5485-5489). This is the first documented report on the structure of singly or doubly linked phycourobilins.  相似文献   

12.
The DNA duplex corresponding to the entire length (126 nucleotides) of the precursor for an Escherichia coli tyrosine tRNA has been synthesized. Duplex [I] (Sekiya, T., Besmer, P., Takeya, T., and Khorana, H. G.(1976) J. Biol. Chem. 251, 634-641), corresponding to the nucleotide sequence 1-26, containing single-stranded ends and carrying one appropriately labeled 5'-phosphate group, was joined to duplex [II] (Loewen, P. C., Miller, R. C., Panet, A., Sekiya, T., and Khorana, H. G. (1976) J. Biol. Chem. 251, 642-650) (nucleotide sequence 23-66 or 23-60) was phosphorylated with [gamma-33P]ATP at the 5'-OH ends. Duplex [III] (Panet, A., Kleppe, R., Kleppe, K., and Khorana, H. G. (1976) J. Biol. Chem. 251, 651-657) (nucleotide sequence 57-94 (Fig. 2)) was also phosphorylated at 5'-ends with [gamma-33P]ATP and was joined to duplex [IV] (Caruthers, M. H., Kleppe, R., Kleppe, K., and Khorana, H. G. (1976) J. Biol. Chem. 251, 658-666) (nucleotide sequence 90-126) which carried a 33P-labeled phosphate group on nucleotide 90. The joined product, duplex [III + IV] (nucleotide sequence 57-126) was characterized. The latter duplex was joined to the duplex [I + II] to give the total duplex. The latter contains singlestranded ends (nucleotides 1 to 6 and 121 to 126) which can either be "filled in" to produce the completely base-paired duplex or may be used to add the promoter and terminator regions at the appropriate ends.  相似文献   

13.
Previous crystallographic studies of the antibacterial trimethoprim in complexes with bacterial and avian dihydrofolate reductases have shown substantial differences in the mode of binding, providing plausible explanations for the origin of the remarkable species selectivity of this inhibitor (Matthews, D. A., Bolin, J. T., Burridge, J. M., Filman, D. J., Volz, K. W., Kaufman, B. T., Beddell, C. R., Champness, J. N., Stammers, D. K., and Kraut, J. (1985) J. Biol. Chem. 260, 381-391; Matthews, D. A., Bolin, J. T., Burridge, J. M., Filman, D. J., Volz, K. W., and Kraut, J. (1985) J. Biol. Chem. 260, 392-399). A major species difference between the active sites is that the only carboxylate present is always Glu in vertebrates and Asp in bacteria. Crystallographic studies of the wild-type and E30D mutant of the enzyme from mouse now reveal that in both cases trimethoprim is bound in an identical fashion to that observed with the bacterial enzyme, and there is no obvious single explanation for the origin of the 10(5)-fold selectivity of trimethoprim binding. In an earlier study of a mouse wild-type enzyme using more limited data it was proposed that trimethoprim bound in the avian mode (Stammers, D. K., Champness, J. N., Beddell, C. R., Dann, J. G., Eliopoulos, E. E., Geddes, A. J., Ogg, D., and North, A. C. T. (1987) FEBS Lett. 218, 178-184), but a re-examination indicates that the occupancy of the active site by trimethoprim is less than had been thought, and we are currently unable to make an unambiguous interpretation of the electron density maps and cannot confirm the avian mode of binding in those crystals.  相似文献   

14.
Two monoclonal antibodies against different epitopes in Escherichia coli ribosomal protein L7/L12 were prepared and characterized as reported previously (Sommer, A., Etchison, J.R., Gavino, G., Zecherle, N., Casiano, C., and Traud, R.R. (1985) J. Biol. Chem. 260, 6522-6527). Both antibodies strongly inhibited polyuridylic acid-directed polyphenylalanine synthesis, ribosome-dependent GTPase activity, and the binding of elongation factor G to the ribosome at mole ratios over ribosomes of 4:1 or less. One epitope was shown to be within residues 1-73 (Ab 1-73) and the other within 74-120 (Ab 74-120). Incubation of 50 S ribosomal subunits or 70 S ribosomes with Ab 1-73, but not with Ab 74-120, leads to a partial loss of L7/L12 from the particle with no loss of any other protein. The experiment was repeated with ribosomes reconstituted with pure radioactive L7/L12 of determined specific activity in order to quantify the L7/L12 in the antibody-treated particle. The protein-deficient core particles isolated by sucrose gradient centrifugation after incubation with Ab 1-73 were found to contain, on average, two copies of L7/L12 and one Ab 1-73. The constancy of this stoichiometry in many experiments and the demonstration of Ab 1-73 on all particles indicate the presence of a homogeneous population of ribosomes, each with only one of the two L7/L12 dimers originally present. The results show a difference in the interactions of the two dimers with the ribosome and present a means of preparing ribosomes with one dimer in a specific binding site. The accompanying paper (Olson, H.M., Sommer, A., Tewari, D. S., Traut, R.R., and Glitz, D.G. (1986) J. Biol. Chem. 261, 6924-6932) shows by immune electron microscopy the location of the two antibody-binding sites and the effect of Ab 1-73 on structure.  相似文献   

15.
16.
17.
We have previously shown that Phe(120), Glu(216), and Asp(301) in the active site of cytochrome P450 2D6 (CYP2D6) play a key role in substrate recognition by this important drug-metabolizing enzyme (Paine, M. J., McLaughlin, L. A., Flanagan, J. U., Kemp, C. A., Sutcliffe, M. J., Roberts, G. C., and Wolf, C. R. (2003) J. Biol. Chem. 278, 4021-4027 and Flanagan, J. U., Maréchal, J.-D., Ward, R., Kemp, C. A., McLaughlin, L. A., Sutcliffe, M. J., Roberts, G. C., Paine, M. J., and Wolf, C. R. (2004) Biochem. J. 380, 353-360). We have now examined the effect of mutations of these residues on interactions of the enzyme with the prototypical CYP2D6 inhibitor, quinidine. Abolition of the negative charge at either or both residues 216 and 301 decreased quinidine inhibition of bufuralol 1'-hydroxylation and dextromethorphan O-demethylation by at least 100-fold. The apparent dissociation constants (K(d)) for quinidine binding to the wild-type enzyme and the E216D and D301E mutants were 0.25-0.50 microm. The amide substitution of Glu(216) or Asp(301) resulted in 30-64-fold increases in the K(d) for quinidine. The double mutant E216Q/D301Q showed the largest decrease in quinidine affinity, with a K(d) of 65 microm. Alanine substitution of Phe(120), Phe(481),or Phe(483) had only a minor effect on the inhibition of bufuralol 1'-hydroxylation and dextromethorphan O-demethylation and on binding. In contrast to the wild-type enzyme, a number of the mutants studied were found to be able to metabolize quinidine. E216F produced O-demethylated quinidine, and F120A and E216Q/D301Q produced both O-demethylated quinidine and 3-hydroxyquinidine metabolites. Homology modeling and molecular docking were used to predict the modes of quinidine binding to the wild-type and mutant enzymes; these were able to rationalize the experimental observations.  相似文献   

18.
Two epidermal growth factor-stimulated protein kinases that correspond to ERK1 and ERK2 have been purified from human epidermoid carcinoma cells (Northwood, I. C., Gonzalez, F. A., Wartmann, M., Raden, D. L., and Davis, R. J. (1991) J. Biol. Chem. 266, 15266-15276). A consensus primary sequence for substrates of ERK1 has been identified as -Pro-Leu-Ser/Thr-Pro- (Alvarez, E., Northwood, I. C., Gonzalez, F. A., Latour, D. A., Seth, A., Abate, C., Curran, T., and Davis, R. J. (1991) J. Biol. Chem. 266, 15277-15285). However, the structural determinants for substrate recognition are not understood. We performed a systematic analysis of the effect of point mutations in the primary sequence of peptide substrates on the rate of phosphorylation by ERK1 and ERK2. The results of this investigation demonstrate that the substrate specificities of the ERK1 and ERK2 protein kinases are very similar. We propose that the primary sequence of substrates for ERK1 and ERK2 protein kinases can be generalized as -Pro-Xaan-Ser/Thr-Pro- (where Xaa is a neutral or basic amino acid and n = 1 or 2).  相似文献   

19.
Using a combination of theoretical sequence structure recognition predictions and experimental disulfide bond assignments, a three-dimensional (3D) model of human interleukin-7 (hIL-7) was constructed that predicts atypical surface chemistry in helix D that is important for receptor activation. A 3D model of hIL-7 was built using the X-ray crystal structure of interleukin-4 (IL-4) as a template (Walter MR et al., 1992, J Mol Biol. 224:1075-1085; Walter MR et al., 1992, J Biol Chem 267:20371-20376). Core secondary structures were constructed from sequences of hIL-7 predicted to form helices. The model was constructed by superimposing IL-7 helices onto the IL-4 template and connecting them together in an up-up down-down topology. The model was finished by incorporating the disulfide bond assignments (Cys3, Cys142), (Cys35, Cys130), and (Cys48, Cys93), which were determined by MALDI mass spectroscopy and site-directed mutagenesis (Cosenza L, Sweeney E, Murphy JR, 1997, J Biol Chem 272:32995-33000). Quality analysis of the hIL-7 model identified poor structural features in the carboxyl terminus that, when further studied using hydrophobic moment analysis, detected an atypical structural property in helix D, which contains Cys 130 and Cys142. This analysis demonstrated that helix D had a hydrophobic surface exposed to bulk solvent that accounted for the poor quality of the model, but was suggestive of a region in IL-7 that maybe important for protein interactions. Alanine (Ala) substitution scanning mutagenesis was performed to test if the predicted atypical surface chemistry of helix D in the hIL-7 model is important for receptor activation. This analysis resulted in the construction, purification, and characterization of four hIL-7 variants, hIL-7(K121A), hIL-7(L136A), hIL-7(K140A), and hIL-7(W143A), that displayed reduced or abrogated ability to stimulate a murine IL-7 dependent pre-B cell proliferation. The mutant hIL-7(W143A), which is biologically inactive and displaces [125I]-hIL-7, is the first reported IL-7R system antagonist.  相似文献   

20.
Passive uptake of potassium acetate into the mitochondrial matrix can be induced by nigericin, a K+/H+ antiporter, or by A23187, a Mg2+/2H+ antiporter. The latter process is thought to reflect operation of the Mg2+-dependent, endogenous K+/H+ antiporter, but there is ambiguity with respect to the mechanism of K+ transport in this assay (Nakashima, R.A., and Garlid, K.D. (1982) J. Biol. Chem. 257, 9252-9254). Kinetic analysis of potassium acetate transport provides verification that Mg2+ depletion 1) unmasks the K+/H+ antiporter, 2) opens up an intrinsic anion uniporter, 3) has no effect on acetic acid transport, and 4) does not induce high K+ uniport conductance. Mg2+-dependent uptake of potassium acetate is thereby shown to be mediated specifically by operation of the endogenous K+/H+ antiporter, as previously proposed. An extension of this analysis confirms that N,N'-dicyclohexylcarbodiimide and quinine block potassium acetate uptake via specific action on the K+/H+ antiporter. These findings support those of a previous study (Martin, W.H., Beavis, A.D., and Garlid, K.D. (1984) J. Biol. Chem. 259, 2062-2065) in which binding of [14C]N,N'-dicyclohexylcarbodiimide to membrane proteins under selective conditions was used to identify an 82,000-dalton band as the protein responsible for K+/H+ antiport in mitochondria.  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