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.
Homodimeric hexokinase 2 from Saccharomyces cerevisiae is known to have two sites of phosphorylation: for serine-14 the modification in vivo increases with glucose exhaustion [Kriegel et al. (1994) Biochemistry 33, 148-152], while for serine-157 it occurs in vitro with ATP in the presence of nonphosphorylateable five-carbon analogues of glucose [Heidrich et al. (1997) Biochemistry 36, 1960-1964]. We show now by site-directed mutagenesis and sedimentation analysis that serine-14 phosphorylation affects the oligomeric state of hexokinase, its substitution by glutamate causing complete dissociation; glutamate exchange for serine-157 does not. Phosphorylation of wild-type hexokinase at serine-14 likewise causes dissociation in vitro. In view of the higher glucose affinity of monomeric hexokinase and the high hexokinase concentration in yeast [Womack, F., and Colowick, S. P. (1978) Arch. Biochem. Biophys. 191, 742-747; Mayes, E. L., Hoggett, J. G., and Kellett, G. L. (1983) Eur. J. Biochem. 133, 127-134], we speculate that the in vivo phosphorylation at serine-14 as transiently occurring in glucose derepression might provide a mechanism to improve glucose utilization from low level and/or that nuclear localization of the monomer might be involved in the signal transduction whereby glucose causes catabolite repression.  相似文献   

2.
3-Phosphoglycerate kinase is a typical two-domain "hinge-bending" enzyme, which is known to be regulated by multivalent anions. Here a relationship between this regulation and the hinge-bending domain closure is proposed on the basis of enzyme kinetic analysis and molecular modeling. Activation of the pig muscle enzyme at low concentrations and inhibition at high concentrations of various anionic analogues of the substrate 3-phosphoglycerate or of the nonsubstrate metal-free ATP are described by a two-site model assuming separate sites for activation and inhibition, respectively. Kinetic experiments with various pairs of analogues suggest the presence of a common site for activation by all effectors, separate from the catalytic site for 3-phosphoglycerate; and a common site for inhibition, except for metal-free ATP, identical with the catalytic site of 3-phosphoglycerate. An additional inhibiting site for all of the anions investigated, including metal-free ATP, is also proposed. A similar two-site model can describe activation of the enzyme by a large excess of each substrate; here the ligand binds to the catalytic site as a substrate and to the regulatory site as an activator. Activation is exerted not only by the physiological substrate, 3-phophoglycerate, but also by a synthetic weak substrate. The activity in the reaction with 3-phosphoglycerate and MgATP is greatly enhanced by the simultaneous presence of the weak substrate. This finding clearly proves the existence of a regulatory site, separate from the catalytic site. This regulatory site, however, may only exist in the catalytically competent closed conformation of the enzyme, as indicated by molecular modeling. Docking of the regulator anions into the known X-ray structures of the enzyme revealed the appearance of an anion binding site between the two domains, including the invariant residues of Lys-215 (C-domain) and of Arg-65 among other residues of the basic cluster (N-domain), as a consequence of the large-scale substrate-induced conformational change that leads to domain closure.  相似文献   

3.
The Type I isozyme of mammalian hexokinase has evolved by a gene duplication-fusion mechanism, with resulting internal duplication of sequence and ligand binding sites. However, 1:1 binding stoichiometry indicates that only one of these is available for binding the product inhibitor, Glc-6-P; the location of that site, in the N- or C-terminal half, remains under debate. Recent structural studies (Aleshin et al., Structure 6, 39-50, 1998; Mulichak et al., Nature Struct. Biol. 5, 555-560, 1998) implicated Asp 84 or its analog in the C-terminal half, Asp 532, in binding of Glc-6-P. Zeng et al. (Biochemistry 35, 13157-13164, 1996) demonstrated that mutation of Asp 532 to Lys or Glu did not affect inhibition by the Glc-6-P analog, 1,5-anhydroglucitol-6-P. These same mutations, as well as mutation to Ala, at the Asp 84 position are now shown to result in increased Ki for 1,5-anhydroglucitol-6-P. The ability of Pi to antagonize inhibition by the Glc-6-P analog is severely diminished or abolished by these mutations, suggesting that antagonism is dependent on precise positioning of the inhibitory hexose 6-phosphate. The structure of the enzyme complexed with Glc and Pi has been determined, and shows that Pi occupies the same site as the 6-phosphate group in the complex with Glc-6-P. Thus, antagonism between these ligands results from competition for a common anion binding site in the N-terminal half.  相似文献   

4.
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   

5.
The inhibitory effects of 4 kinds of 2'-deoxy-L-nucleoside 5'-triphosphates, which are enantiomers of natural dNTPs, on murine deoxycytidine kinase (dCK) were investigated. When ATP was used as the phosphate donor, L-dCTP showed significant inhibitory action noncompetitively and competitively with 2'-deoxycytidine (dCyd) and ATP, respectively. Thus L-dCTP, like dCTP, could serve as a feedback inhibitor for dCK. Recently, it has been demonstrated that human dCK can utilize L-dCyd as a substrate (Verri, A. et al. (1997) Mol. Pharmacol., 51, 132). The present results suggest that dCK is also unable to discriminate the chirality of nucleotides at the phosphate donor binding site of the enzyme.  相似文献   

6.
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   

7.
We have compared the active sites of Escherichia coli Fe-substituted (Mn)superoxide dismutase [Fe-sub-(Mn)SOD] and Fe-SOD to elucidate the basis for the inactivity of Fe-sub-(Mn)SOD, despite its apparent similarity to Fe-SOD. The active site of (reduced) Fe2+-sub-(Mn)SOD is qualitatively similar to that of native Fe2+-SOD, indicating similar active site structures and coordination environments for Fe2+. Its nativelike pK is indicative of nativelike local electrostatics, and consistent with Fe2+-sub-(Mn)SOD's retention of ability to reduce O2*- [Vance and Miller (1998) J. Am. Chem. Soc. 120(3), 461-467]. The active site of (oxidized) Fe3+-sub-(Mn)SOD differs from that of Fe3+-SOD with respect to the EPR signals produced at both neutral and high pH, indicating different coordination environments for Fe3+. Although Fe3+-sub-(Mn)SOD binds the small anions N3- and F-, the KD for N3- is tighter than that of Fe3+-SOD, suggesting that the (Mn)SOD protein favors anion binding more than does the (Fe)SOD protein. The EPR spectral consequences of binding F- are reminiscent of those observed upon binding the first F- to Fe3+-SOD, but the EPR spectrum obtained upon binding N3- is different, consistent with crystallographic observation of a different binding mode for N3- in Thermus thermophilus Mn-SOD than Fe-SOD [Lah, M., et al. (1995) Biochemistry 34, 1646-1660]. We find a pK of 8.5 to be associated with dramatic changes in the EPR spectrum. In addition, we confirm the pK between 6 and 7 that has previously been reported based on changes in the optical signal and N3- binding [Yamakura, F., et al. (1995) Eur. J. Biochem. 227, 700-706]. However, this latter pK appears to be associated with much subtler changes in the EPR spectrum. The non-native pKs observed in Fe3+-sub-(Mn)SOD and the differences in the Fe3+ coordination indicated by the EPR spectra are consistent with Fe3+-sub-(Mn)SOD's inability to oxidize O2*- and suggest that its low E degrees is due to perturbation of the oxidized state.  相似文献   

8.
The catalytic subunit of the cAMP-dependent protein kinase from Dictyostelium discoideum, PkaC, displays the same properties as its mammalian counterpart, except for being about twice as large in size. Sequence comparisons indicated the presence of a conserved alpha-helix (A-helix) within the N-terminal region of PkaC which could potentially establish close contacts with the catalytic core [Véron, M., et al. (1993) Proc. Natl. Acad. Sci. U.S.A. 90, 10618-10622]. We show in this report that a synthetic peptide with the A-helix sequence inhibits PKA activity, whereas unrelated peptides display no inhibitory activity. The inhibition seems competitive with respect to the kemptide substrate rather than due to binding to a secondary site. We further show by amino acid replacements that the last lysine of the A-helix sequence is involved in this specific inhibition. A model is proposed for the possible role of the A-helix.  相似文献   

9.
Mutations in the gene encoding the cystic fibrosis transmembrane conductance regulator (CFTR) cause cystic fibrosis (CF) (Collins, 1992). Over 500 naturally occurring mutations have been identified in CF gene which are located in all of the domains of the protein (Kerem et al., 1990; Mercier et al., 1993; Ghanem et al., 1994; Fanen et al., 1992; Ferec et al., 1992; Cutting et al., 1990). Early studies by several investigators characterized CFTR as a chloride channel (Anderson et al.; 1991b,c; Bear et al., 1991). The complex secondary structure of the protein suggested that CFTR might possess other functions in addition to being a chloride channel. Studies have established that the CFTR functions not only as a chloride channel but is indeed a regulator of sodium channels (Stutts et al., 1995), outwardly rectifying chloride channels (ORCC) (Gray et al., 1989; Garber et al., 1992; Egan et al., 1992; Hwang et al., 1989; Schwiebert et al., 1995) and also the transport of ATP (Schwiebert et al., 1995; Reisin et al., 1994). This mini-review deals with the studies which elucidate the functions of the various domains of CFTR, namely the transmembrane domains, TMD1 and TMD2, the two cytoplasmic nucleotide binding domains, NBD1 and NBD2, and the regulatory, R, domain.  相似文献   

10.
Upon binding of substrates the catalytic subunit (C) of cAMP-dependent protein kinase (cAPK) undergoes significant induced conformational changes that lead to catalysis. For the free apoenzyme equilibrium favors a more open and malleable conformation while the ternary complex of C, MgATP, and a 20-residue inhibitor peptide [PKI (5-24)] adopts a tight and closed conformation [Zheng, J., et al. (1993) Protein Sci. 2, 1559]. It is not clear that binding of either ligand alone is responsible for this conformational switch or whether both are required. In addition, the catalytic subunit binds MgATP and inhibitor peptide synergistically. The structural basis for this synergism is also not defined at present. Using an Fe-EDTA-mediated protein footprinting technique, the conformational changes associated with the binding of MgATP and the heat stable protein kinase inhibitor (PKI) were probed by mapping the solvent-accessible surface and structural dynamics of C. The conformation of the free enzyme was clearly distinguished from the ternary complex. Furthermore, binding of MgATP alone induced extensive conformational changes, both local and global, that include the glycine-rich loop, the linker connecting the small and large lobes, the catalytic loop, the Mg2+ positioning loop, the activation loop, and the F helix. These changes, similar to those seen in the ternary complex, are consistent with a transition from an open to a more closed conformation and likely reflect the motions that are associated with catalysis and product release. In contrast, the footprinting pattern of C.PKI resembled free C, indicating minimal conformational changes. Binding of MgATP, by shifting the equilibrium to a more closed conformation, "primes" the enzyme so that it is poised for the docking of PKI and provides an explanation for synergism between MgATP and PKI.  相似文献   

11.
The interaction of ATP with the active site of hexokinase is unknown since the crystal structure of the hexokinase-ATP complex is unavailable. It was found that the ATP binding site of brain hexokinase is homologous to that of actin, heat shock protein hsc70, and glycerol kinase. On the basis of these similarities, the ATP molecule was positioned in the catalytic domain of human brain hexokinase, which was modeled from the X-ray structure of yeast hexokinase. Site-directed mutagenesis was performed to test the function of residues presumably involved in interaction with the tripolyphosphoryl moiety of ATP. Asp532, which is though to be involved in binding the Mg2+ ion of the MgATP2- complex, was mutated to Lys and Glu. The kcat values decreased 1000- and 200-fold, respectively, for the two mutants. Another residue, Thr680 was proposed to interact with the gamma-phosphoryl group of ATP through hydrogen bonds and was mutated to Val and Ser. The kcat value of the Thr680Val mutant decreased 2000-fold, whereas the kcat value of the Thr680Ser decreased only 2.5-fold, implying the importance of the hydroxyl group. The Km and dissociation constant values for either ATP or glucose of all the above mutants showed little or no change relative to the wild-type enzyme. The Ki values for the glucose 6-phosphate analogue 1,5-anhydroglucitol 6-phosphate, were the same as that of the wild-type enzyme, and the inhibition was reversed by inorganic phosphate (Pi) for all four mutants. The circular dichroism spectra of the mutants were the same as that of the wild-type enzyme. The results from the site-directed mutagenesis demonstrate that the presumed interactions of investigated residues with ATP are important for the stabilization of the transition state.  相似文献   

12.
While most of the sequence of myosin's motor domain is highly conserved among various organisms and tissue types, the junctions between the 25 and 50 kDa domains and the 50 and 20 kDa domains are strikingly divergent. The 50-20K loop is positioned to interact with actin, while the 25-50K loop is situated nearer the ATP binding site [Rayment, I., et al. (1993) Science 261, 50-58]. Chimeric studies of the 50-20K loop [Uyeda, T. Q.-P., et al. (1994) Nature 368, 567-569; Rovner, A. S., et al. (1995) J. Biol. Chem. 270 (51), 30260-30263] have shown that this loop affects actin activation of ATPase activity. Given the function of myosin as a molecular motor, it was proposed that the 25-50K loop might specifically alter ADP release [Spudich, J. A. (1994) Nature 374, 515-518]. Here we study the role of this loop by engineering chimeras containing the Dictyostelium myosin heavy chain with loops from two enzymatically diverse myosins, rabbit skeletal and Acanthamoeba. The chimeric myosins complement the myosin null phenotype in vivo, bind nucleotide normally, interact normally with actin, and display wild-type levels of actin-activated ATPase activity. However, the rate of ADP release from the myosins, normally the slowest step involved in motility, was changed in a manner that reflects the activity of the donor myosin. In summary, studies of Dictyostelium myosin heavy chain chimeras have shown that the 50-20K sequence specifically affects the actin-activated ATPase activity [Uyeda, T. Q.-P., et al. (1994)] while the 25-50K sequence helps determine the rate of ADP release.  相似文献   

13.
The wild-type binding pentamer of Shiga-like toxin IIe (SLT-IIe) binds both the globotriaosylceramide (Gb3) and globotetraosylceramide (Gb4) cell surface glycolipids, whereas the double mutant GT3 (Q65E/K67Q) exhibits a marked preference for Gb3 [Tyrrell, G. J., et al. (1992) Proc. Natl. Acad. Sci. U.S.A. 89, 524-528]. We modeled three unique sites (sites 1-3) for binding of the carbohydrate moiety of Gb3 to GT3 and SLT-IIe, on the basis of the three sites observed for the SLT-I pentamer [Ling, H., et al. (1998) Biochemistry 37, 1777-1788]. Examination of the three sites in light of various mutation and binding data strongly suggested that one of the binding sites plays a role in the change of specificity observed for the GT3 mutant. We applied several modeling techniques, and developed a model for binding of the carbohydrate moiety of Gb4 to this site of the SLT-IIe binding pentamer. This model is consistent with a wide variety of mutation and binding data and clearly shows the importance of the terminal GalNAc residue of Gb4, as well as that of the two mutated residues of GT3, to the intermolecular interaction.  相似文献   

14.
The solution structure of the peptide antigen from the receptor binding domain of Pseudomonas aeruginosa strain P1 has been determined using two-dimensional 1H NMR techniques. Ensembles of solution conformations for the trans form of this 23-residue disulfide bridged peptide have been generated using a simulated annealing procedure in conjunction with distance and torsion angle restraints derived from NMR data. Comparison of the NMR-derived solution structures of the P1 peptide with those previously determined for the 17-residue PAK, PAO and KB7 strain peptides [McInnes, C., et al. (1993) Biochemistry 32, 13432-13440; Campbell, A.P., et al. (1995) Biochemistry 34, 16255-16268] reveals the common structural motif of a beta-turn, which may be the necessary structural requirement for recognition of a common cell surface receptor and a common cross-reactive antibody to which all four strains bind. The importance of this conserved beta-turn in the PAK, PAO, KB7 and P1 peptides is discussed with regard to the design of a synthetic peptide vaccine effective against multiple strains of Pseudomonas aeruginosa infections.  相似文献   

15.
Previous studies have shown that Schwann cells (SCs) differentiate into myelin-forming or ensheathing cells only under conditions which allow the deposition of basal lamina and extracellular collagen [Bunge (1993) Peripheral Neuropathy, pp. 299-316]. SC adhesion to basal lamina is mediated by beta1 integrins and function blocking antibodies to beta1 integrins inhibit myelination [Fernandez-Valle et al. (1993) Development 119:867-880]. Recently, focal adhesion kinase (FAK), a cytoplasmic non-receptor tyrosine kinase, was found to mediate beta1 integrin-dependent signalling in a variety of cultured cell types adhering to ECM components such as fibronectin [reviewed in Schwartz et al. (1995) Ann. Rev. Cell Biol. 11:549-599; Ilic et al. (1997) J. Cell Sci. 110:401-407]. In the present study, we have determined more precisely the respective time courses of ECM deposition and myelination. In addition, we have studied by immunocytochemistry, immuno-gold labelling, and electron microscopy the expression and subcellular localization of FAK in nondifferentiating SCs and in SCs differentiating into myelinating cells. We show that the development of basal lamina and extracellular collagen fibrils precedes by 3 days the appearance of the first myelin sheaths. FAK was detected by immunocytochemistry or immuno-gold labelling only in SCs differentiating in the presence of ascorbic acid. Localization of FAK to the abaxonal plasma membrane was dependent upon ECM deposition. Cytochalasin D did not prevent or disrupt localization of FAK to the plasma membrane. These data support the possibility that FAK acts as an intermediate in the pathway by which basal lamina regulates SC differentiation.  相似文献   

16.
Phosphoglycerate kinase (PGK), a key enzyme in glycolysis, catalyses the transfer of a phosphoryl-group from 1,3-bisphosphoglycerate to ADP to form 3-phosphoglycerate and ATP. Despite extensive kinetic and structural investigations over more than two decades, the conformation assumed by this enzyme during catalysis remained unknown. Here we present the 2.8 A crystal structure of a ternary complex of PGK from Trypanosoma brucei, the causative agent of sleeping sickness. This structure determination relied on a procedure in which fragments containing less than 10% of the scattering mass were successively positioned in the unit cell to obtain phases. The PGK ternary complex exhibits a dramatic closing of the large cleft between the two domains seen in all previous studies, thereby bringing the two ligands, 3-phosphoglycerate and ADP into close proximity. Our results demonstrate that PGK is a hinge-bending enzyme, reveal a novel mechanism in which substrate-induced effects combine synergistically to induce major conformational changes and, to our knowledge, afford the first observation of the PGK active site in a catalytic conformation.  相似文献   

17.
gamma-Aminobutyric acidA (GABAA) receptors are linked to ion channels which mediate many aspects of neural inhibition. Although the effects of phosphorylation on GABAA receptor function have been widely studied, the actual role of phosphorylation in the regulation of these receptors still remains controversial. In recent reports, we have described the effects of phosphorylating/dephosphorylating enzymes on the regulation of GABAA receptors in a rat cortical slice preparation (Shaw et al., Mol. Neuropharmacol., 2 (1992) 297-302; Shaw and Lanius, Dev. Brain Res., 70 (1992) 153-161; Pasqualotto et al., Neuroreport, 4 (1993) 447-450) and predicted that ionic co-factors are involved in mediating the regulation of GABAA receptors by kinases and phosphatases. In the present report, the effects of chloride, sodium, potassium, and calcium were examined alone and in the presence of cAMP-dependent protein kinase (protein kinase A) or alkaline phosphatase. The results showed a decrease in [3H]SR 95531 (GABAA receptor antagonist) binding after incubation with chloride alone; this decrease was further enhanced in the presence of protein kinase A. Both effects could be blocked by a protein kinase A inhibitor. Conversely, an increase in [3H]SR 95531 binding was observed after incubation with sodium alone; this increase was further enhanced in the presence of alkaline phosphatase. In both cases these increases in binding could be blocked by sodium orthovanadate, a phosphatase inhibitor. Potassium was ineffective under all conditions; calcium showed enzyme-independent effects at low concentrations only. These results suggest the existence of a novel chloride-dependent protein kinase which may have significant sequence homology to protein kinase A, and a novel sodium-dependent phosphatase.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Diabetes mellitus is associated with typical patterns of long term vascular complications which vary with the organ involved. The microvascular kidney disease (Olgemoller and Schleicher, 1993) is characterized by thickening of the capillary basement membranes and increased deposition of extracellular matrix components (ECM), while loss of microvessels with subsequent neovascularisation is predominant in the eye and peripheral nerves. On the other hand macrovascular disease is characterized by accelerated atherosclerosis. These complications are dependent on long term hyperglycemia. Specific biochemical pathways linking hyperglycaemia to microvascular changes were proposed: the polyol pathway (Greene et al., 1987), non-enzymatic glycation of proteins (Brownlee et al., 1988), glucose autooxidation and oxidative stress (Hunt et al., 1990), hyperglycemic pseudohypoxia (Williamson et al., 1993) enhanced activation of protein kinase C by de novo-synthesis of diacyl glycerol (Lee et al., 1989; DeRubertis and Craven 1994) and others. These pathways are not mutually exclusive (Larkins and Dunlop, 1992; Pfeiffer and Schatz, 1992). They may be linked to alterations in the synthesis of growth factors particularly since atherosclerosis and angioneogenesis are associated with increased proliferation of endothelial and smooth muscle cells. Increased synthesis of ECM components is stimulated by growth factors like transforming growth factor beta (TGF beta) (Derynck et al., 1984) and insulin-like growth factor I (IGF-I) (Moran et al., 1991). This review will summarize some of the recent evidence for an involvement of growth factors in diabetic vascular complications and will attempt to assign their emergence in the sequence of events leading to vascular complications.  相似文献   

19.
Conformations of cation-nucleotide complexes bound to rabbit muscle creatine kinase were investigated by measuring paramagnetic effects on 13C spin relaxation in E.Mn[2-13C]ATP and E.Mn[2-13C]ADP at three different frequencies, viz., 50, 75, and 125 MHz, and as a function of temperature in the range of 7-35 degrees C (at 75 MHz). Arrhenius plots of the temperature dependencies of relaxation rates show a positive slope with low activation energies of 1.3 +/- 0.2 kcal/mol and 2.0 +/- 0.2 kcal/mol for E.Mn ATP and E.MnADP, respectively. The relaxation rates of both complexes show strong frequency dependence, indicating that these rates are not exchange limited. Analysis of the data yields Mn(II)-2C distances of 10.0 +/- 0.5 A for E.MnATP and 8.6 +/- 0.5 A for E.MnADP. These data were interpreted, along with previously published information, on the location of the cation with respect to the phosphate chain [Jarori, G. K., Ray, B.D., & Nageswara Rao, B. D. (1985) Biochemistry 24, 3487-3494], and on the adenosine conformation [Murali, N., Jarori, G. K., & Nageswara Rao, B. D. (1993) Biochemistry 32, 12941-12948] in these complexes. The Mn(II)-2C distances depend on the orientation of the phosphate chain relative to the adenosine moiety. Conformational searches were performed by varying the two torsion angles, phi 1 (C4'-C5'-O5'-P alpha), and phi 2 (C5'-O5'-P alpha-O alpha beta), along with CHARMm energy computations, in order to determine acceptable conformations compatible with the distances determined. The significant difference in the Mn(II)-2C distances in E.MnATP and E.MnADP is indicative of the structural alterations occurring at the active site as the enzyme turns over.  相似文献   

20.
The compound U0126 (1,4-diamino-2,3-dicyano-1, 4-bis[2-aminophenylthio]butadiene) was identified as an inhibitor of AP-1 transactivation in a cell-based reporter assay. U0126 was also shown to inhibit endogenous promoters containing AP-1 response elements but did not affect genes lacking an AP-1 response element in their promoters. These effects of U0126 result from direct inhibition of the mitogen-activated protein kinase kinase family members, MEK-1 and MEK-2. Inhibition is selective for MEK-1 and -2, as U0126 shows little, if any, effect on the kinase activities of protein kinase C, Abl, Raf, MEKK, ERK, JNK, MKK-3, MKK-4/SEK, MKK-6, Cdk2, or Cdk4. Comparative kinetic analysis of U0126 and the MEK inhibitor PD098059 (Dudley, D. T., Pang, L., Decker, S. J., Bridges, A. J., and Saltiel, A. R. (1995) Proc. Natl. Acad. Sci U. S. A. 92, 7686-7689) demonstrates that U0126 and PD098059 are noncompetitive inhibitors with respect to both MEK substrates, ATP and ERK. We further demonstrate that the two compounds bind to deltaN3-S218E/S222D MEK in a mutually exclusive fashion, suggesting that they may share a common or overlapping binding site(s). Quantitative evaluation of the steady state kinetics of MEK inhibition by these compounds reveals that U0126 has approximately 100-fold higher affinity for deltaN3-S218E/S222D MEK than does PD098059. We further tested the effects of these compounds on the activity of wild type MEK isolated after activation from stimulated cells. Surprisingly, we observe a significant diminution in affinity of both compounds for wild type MEK as compared with the deltaN3-S218E/S222D mutant enzyme. These results suggest that the affinity of both compounds is mediated by subtle conformational differences between the two activated MEK forms. The MEK affinity of U0126, its selectivity for MEK over other kinases, and its cellular efficacy suggest that this compound will serve as a powerful tool for in vitro and cellular investigations of mitogen-activated protein kinase-mediated signal transduction.  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