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 421 毫秒
1.
Many proteins of the CLC gene family are Cl(-) channels, whereas others, like the bacterial ecClC-1 or mammalian ClC-4 and -5, mediate Cl(-)/H(+) exchange. Mutating a "gating glutamate" (Glu-224 in ClC-4 and Glu-211 in ClC-5) converted these exchangers into anion conductances, as did the neutralization of another, intracellular "proton glutamate" in ecClC-1. We show here that neutralizing the proton glutamate of ClC-4 (Glu-281) and ClC-5 (Glu-268), but not replacing it with aspartate, histidine, or tyrosine, rather abolished Cl(-) and H(+) transport. Surface expression was unchanged by these mutations. Uncoupled Cl(-) transport could be restored in the ClC-4(E281A) and ClC-5(E268A) proton glutamate mutations by additionally neutralizing the gating glutamates, suggesting that wild type proteins transport anions only when protons are supplied through a cytoplasmic H(+) donor. Each monomeric unit of the dimeric protein was found to be able to carry out Cl(-)/H(+) exchange independently from the transport activity of the neighboring subunit. NO(3)(-) or SCN(-) transport was partially uncoupled from H(+) countertransport but still depended on the proton glutamate. Inserting proton glutamates into CLC channels altered their gating but failed to convert them into Cl(-)/H(+) exchangers. Noise analysis indicated that ClC-5 switches between silent and transporting states with an apparent unitary conductance of 0.5 picosiemens. Our results are consistent with the idea that Cl(-)/H(+) exchange of the endosomal ClC-4 and -5 proteins relies on proton delivery from an intracellular titratable residue at position 268 (numbering of ClC-5) and that the strong rectification of currents arises from the voltage-dependent proton transfer from Glu-268 to Glu-211.  相似文献   

2.
Hiratsuka T 《Biochemistry》2006,45(4):1234-1241
The fluorescent probe 3-[4-(3-phenyl-2-pyrazolin-1-yl)benzene-1-sulfonyl amido]phenylboronic acid (PPBA) acts as a fluorescent inhibitor for the ATPases of skeletal [Hiratsuka (1994) J. Biol. Chem. 269, 27251-27257] and Dictyostelium discoideum [Bobkov et al. (1997) J. Muscle Res. Cell Motil. 18, 563-571] myosins. The former paper suggested that, upon addition of excess nucleotides to the binary complex of subfragment-1 from skeletal myosin (S1) with PPBA, a stable ternary complex of S1 with PPBA and nucleotide is formed. Useful fluorescence properties of PPBA enable us to distinguish the conformation of the myosin ATPase at the ATP state from that at the ADP state. In the present paper, to determine the PPBA-binding site in the complexes, enzymatic and fluorescence properties of the S1.PPBA.nucleotide complexes were investigated. Upon formation of the ternary complex with ATP, a new peak appeared at 398 nm in the PPBA fluorescence spectrum. Experiments using model compounds of aromatic amino acid suggested that this fluorescence peak at 398 nm is originated from PPBA interacting with Phe residue(s). Taking into account differences in fluorescence spectra between complexes of S1 and those of subfragment-1 from D. discoideum myosin (S1dC), in the ternary complex of S1 formed with ATP, PPBA was suggested to interact with Phe residue(s) that is absent in S1dC. Docking simulation of PPBA on the S1.nucleotide complex revealed that Phe472 interacts with PPBA. Binding sites of PPBA and blebbistatin, an inhibitor showing high affinity and selectivity toward myosin II [Kovács et al. (2004) J. Biol. Chem. 279, 35557-35563], seem to overlap at least partly.  相似文献   

3.
The polypeptides exposed to lipids in the membranous F0 sector of the mitochondrial and Escherichia coli ATP synthases were labelled with radioactive photoreactive lipids. Highly resolving gel electrophoretic conditions were used in order to separate all the eighteen components forming the bovine heart mitochondrial enzyme. The hydrophobic labelling was performed on fully active and inhibitor-sensitive ATP synthases. In the mitochondrial enzyme prepared according to Serrano et al. (1976) [J. Biol. Chem. 251, 2453-2461] seven polypeptides of Mr 30500; 11500; 10500; 10000; 9500; 8500 and 4500 were labelled. The major amount of radioactivity was associated with the 30500-Mr component, which is thought to be the adenine nucleotide carrier. In the preparation of Galante et al., (1979) which almost completely lacks this component [J. Biol. Chem. 254, 12372-12378] nine polypeptides of Mr 25000; 21000; 11500; 10500; 10000; 9500; 9200; 8500 and 4500 were labelled. In the ATPase synthase from E. coli the major amount of labelling was associated with subunit b and only a minor portion with subunit c.  相似文献   

4.
Recently two reports [J. A. Robertson et al. (1986) J. Biol. Chem. 261, 15794-15799 and R. M. Bayney et al. (1987) J. Biol. Chem. 262, 572-575] have appeared concerning the nucleotide sequence of quinone reductase cDNA clones. Although the cDNA clones are virtually identical, they diverge in the 5' region that encodes the NH2 terminus of the protein. In order to clarify the sequence of this region, we have isolated quinone reductase clones from a rat genomic library using a cDNA clone, pDTD55, isolated and characterized by our laboratory. We have determined the sequence of exons 1 and 2 of the structural gene by double-stranded sequencing using oligonucleotide primers. The sequence of exons 1 and 2 of the quinone reductase structural gene along with our previous nucleotide sequence analysis of pDTD55 as well as conventional amino acid sequence analysis of the purified protein indicates that quinone reductase is composed of 274 amino acids with a molecular weight of 30,946. These data agree with the published sequence of lambda NMOR1 reported by Robertson et al.  相似文献   

5.
As in other adenine nucleotide binding cassette (ABC) proteins the nucleotide binding domains of the cystic fibrosis transmembrane conductance regulator (CFTR) bind and hydrolyze ATP and in some manner regulate CFTR ion channel gating. Unlike some other ABC proteins, however, there are preliminary indications that the two domains of CFTR are nonequivalent in their nucleotide interactions (Szabo, K., Szakacs, G., Hegeds, T., and Sarkadi, B. (1999) J. Biol. Chem. 274, 12209-12212; Aleksandrov, L., Mengos, A., Chang, X., Aleksandrov, A., and Riordan, J. R. (2001) J. Biol. Chem. 276, 12918-12923). We have now characterized the interactions of the 8-azido-photoactive analogues of ATP, ADP, and 5'-adenyl-beta,gamma-imidodiphosphate (AMP-PNP) with the two domains of functional membrane-bound CFTR. The results show that the two domains appear to act independently in the binding and hydrolysis of 8-azido-ATP. At NBD1 binding does not require a divalent cation. This binding is followed by minimal Mg(2+)-dependent hydrolysis and retention of the hydrolysis product, 8-azido-ADP, but not as a vanadate stabilized post-hydrolysis transition state complex. In contrast, at NBD2, MgN(3)ATP is hydrolyzed as rapidly as it is bound and the nucleoside diphosphate hydrolysis product dissociates immediately. Confirming this characterization of NBD1 as a site of more stable nucleotide interaction and NBD2 as a site of fast turnover, the non-hydrolyzable N(3)AMP-PNP bound preferentially to NBD1. This demonstration of NBD2 as the rapid nucleotide turnover site is consistent with the strong effect on channel gating kinetics of inactivation of this domain by mutagenesis.  相似文献   

6.
The ClC-2 epithelial cell chloride channel is a voltage-, tonicity- and pH-regulated member of the ClC super family. We have previously shown that rat lung ClC-2 (rClC-2) is down-regulated at birth, and molecular diversity is generated by alternative splicing [Murray et al. (1995) Am. J. Respir. Cell Mol. Biol. 12, 597-604; Murray et al. (1996) Am. J. Physiol. 271, L829-L837; Chu et al . (1996) Nucleic Acids Res. 24, 3453-3457]. To investigate other possible mRNA splice variations, we sequenced the entire rClC-2 gene and found that ClC-2Sa (formerly ClC-2S) results from the deletion of exon 20. The preceding intron 19 has an unusually high CT content and a rare AAG acceptor site. Because both features were also found in intron 13, we next tested the hypothesis that intron 13 would be involved in alternative splicing. As predicted, a second splice product, ClC-2Sb, was found by RT-PCR, but only in lung. When we compared the genomic maps of rClC-2 and human ClC-1 (hClC-1), striking similarities were found in each exon except for rClC-2 exon 20, which is absent in hClC-1. These observations suggest that ClC-1 and ClC-2 may have evolved by gene duplication, mutation and DNA rearrangement.  相似文献   

7.
Proton translocating ATPase (F0F1) from bovine heart mitochondria was reconstituted into planar phospholipid bilayers, and its electrogenicity was directly demonstrated. The F0F1 ATPase was solubilized using 3-[(3-cholamidopropyl)-dimethylammonio]-1-propanesulfonic acid (CHAPS) as a detergent followed by sucrose density gradient centrifugation according to the method originally described by McEnery et al. for rat liver mitochondria (McEnery et al. (1986) J. Biol. Chem. 259, 4642-4651), with minor modifications. The purified ATPase was reconstituted into proteoliposomes and then reconstituted into planar phospholipid bilayers by the modified fusion method (Hirata et al. (1986) J. Biol. Chem. 261, 9839-9843). A short-circuit current of up to 0.4 pA was induced by adding ATP, and this current was suppressed by the F1 ATPase inhibitor NaN3 or by a specific mitochondrial F0 inhibitor, oligomycin. The direction of the current corresponded to the flow of positive charges from the F1 side to the F0 side. All these facts clearly demonstrate that the mitochondrial F0F1 ATPase was successfully reconstituted into planar phospholipid bilayers, and the current was generated by the ATPase.  相似文献   

8.
We have previously shown that inhibition of the ATPase activity of skeletal muscle myosin subfragment 1 (S1) by caldesmon is correlated with the inhibition of S1 binding in the presence of ATP or pyrophosphate (Chalovich, J., Cornelius, P., and Benson, C. (1987) J. Biol Chem. 262, 5711-5716). In contrast, Lash et al. (Lash, J., Sellers, J., and Hathaway, D. (1986) J. Biol. Chem. 261, 16155-16160) have shown that the inhibition of ATPase activity of smooth muscle heavy meromyosin (HMM) by caldesmon is correlated with an increase in the binding of HMM to actin in the presence of ATP. We now show, in agreement, that caldesmon does increase the binding of smooth muscle HMM to actin-tropomyosin while decreasing the ATPase activity. The effect of caldesmon on the binding of smooth HMM is reversed by Ca2+-calmodulin. Caldesmon strengthens the binding of smooth S1.ATP and skeletal HMM.ATP to actin-tropomyosin but to a lesser extent than smooth HMM.ATP. Furthermore, this increase in binding of smooth S1.ATP and skeletal HMM.ATP does not parallel the inhibition of ATPase activity. In contrast, in the absence of ATP, all smooth and skeletal myosin subfragments compete with caldesmon for binding to actin. Thus, the effect that caldesmon has on the binding of myosin subfragments to actin-tropomyosin depends on the source of myosin, the type of subfragment, and the nucleotide present. The inhibition of actin-activated ATP hydrolysis by caldesmon, however, is not greatly different for different smooth and skeletal myosin subfragments. Evidence is presented that caldesmon inhibits actin-activated ATP hydrolysis by attenuating the productive interaction between myosin and actin that normally accelerates ATP hydrolysis. The increased binding seen by some myosin subfragments, in the presence of ATP, may be due to binding of these subfragments to a nonproductive site on actin-caldesmon. The subfragments which show an increase in binding in the presence of ATP and caldesmon appear to bind directly to caldesmon as demonstrated by affinity chromatography.  相似文献   

9.
Lon is an ATP-dependent serine protease that degrades damaged and certain regulatory proteins in vivo. Lon exists as a homo-oligomer and represents one of the simplest ATP-dependent proteases because both the protease and ATPase domains are located within each monomeric subunit. Previous pre-steady-state kinetic studies revealed functional nonequivalency in the ATPase activity of the enzyme [Vineyard, D., et al. (2005) Biochemistry 44, 1671-1682]. Both a high- and low-affinity ATPase site has been previously reported for Lon [Menon, A. S., and Goldberg, A. L. (1987) J. Biol. Chem. 262, 14921-14928]. Because of the differing affinities for ATP, we were able to monitor the activities of the sites separately and determine that they were noninteracting. The high-affinity sites hydrolyze ATP very slowly (k(obs) = 0.019 +/- 0.002 s(-1)), while the low-affinity sites hydrolyze ATP quickly at a rate of 17.2 +/- 0.09 s(-1), which is comparable to the previously observed burst rate. Although the high-affinity sites hydrolyze ATP slowly, they support multiple rounds of peptide hydrolysis, indicating that ATP and peptide hydrolysis are not stoichiometrically linked. However, ATP binding and hydrolysis at both the high- and low-affinity sites are necessary for optimal peptide cleavage and the stabilization of the conformational change associated with nucleotide binding.  相似文献   

10.
ClC-5, an endosomal Cl/H+ antiporter that is mutated in Dent disease, is essential for endosomal acidification and re-uptake of small molecular weight proteins in the renal proximal tubule. Eukaryotic chloride channels (CLCs) contain two cytoplasmic CBS domains, motifs present in different proteins, the function of which is still poorly understood. Structural studies have shown that ClC-5 can bind to ATP at the interface between the CBS domains, but so far the potential functional consequences of nucleotide binding to ClC-5 have not been investigated. Here, we show that the direct application of ATP, ADP and AMP in inside-out patch experiments potentiates the current mediated by ClC-5 with similar affinities. The nucleotides increase the probability of ClC-5 to be in an active, transporting state. The residues Tyr 617 and Asp 727, but not Ser 618, are crucial for the potentiation. These results provide a mechanistic and structural framework for the interpretation of nucleotide regulation of a CLC transporter.  相似文献   

11.
The reactivation of mitochondrial ATPase by acidic and isoelectric phospholipids was studied comparatively with two purified enzyme preparations exhibiting different gel electrophoretic patterns: the preparation of Serrano et al. (1976, J. Biol. Chem. 251, 2453-2461) and the complex V of Galante et al. (1979, J. Biol. Chem. 254, 12372-12379). Isoelectric phosphatidylcholine liposomes showed marked differences in affinity for the two ATPase complexes and produced different maximal reactivations, whereas no significant differences were found with negatively charged liposomes. Analysis of residual phospholipids associated with the two ATPase preparations revealed a greater relative cardiolipin content in complex V. It is proposed that the different patterns of reactivation of the two ATPase preparations by isoelectric phospholipids result from different contents in residual cardiolipin and adenine nucleotide carrier.  相似文献   

12.
Pancreatic duct cells secrete bicarbonate-rich fluids, which are important for maintaining the patency of pancreatic ductal trees as well as intestinal digestive function. The bulk of bicarbonate secretion in the luminal membrane of duct cells is mediated by a Cl(-)-dependent mechanism (Cl(-)/HCO(3)(-) exchange), and we previously reported that the mechanism is CFTR-dependent and cAMP-activated (Lee, M. G., Choi, J. Y., Luo, X., Strickland, E., Thomas, P. J., and Muallem, S. (1999) J. Biol. Chem. 274, 14670-14677). In the present study, we provide comprehensive evidence that calcium signaling also activates the same CFTR- and Cl(-)-dependent HCO(3)(-) transport. ATP and trypsin evoked intracellular calcium signaling in pancreatic duct-derived cells through the activation of purinergic and protease-activated receptors, respectively. Cl(-)/HCO(3)(-) exchange activity was measured by recording pH(i) in response to [Cl(-)](o) changes of the perfusate. In perfusate containing high concentrations of K(+), which blocks Cl(-) movement through electrogenic or K(+)-coupled pathways, ATP and trypsin highly stimulated luminal Cl(-)/HCO(3)(-) exchange activity in CAPAN-1 cells expressing wild-type CFTR, but not in CFPAC-1 cells that have defective (DeltaF508) CFTR. Notably, adenoviral transfection of wild-type CFTR in CFPAC-1 cells completely restored the stimulatory effect of ATP on luminal Cl(-)/HCO(3)(-) exchange. In addition, the chelation of intracellular calcium by 1,2-bis(2-aminophenoxy)ethane-N,N,N,N'-tetraacetic acid (BAPTA) treatment abolished the effect of calcium agonists on luminal Cl(-)/HCO(3)(-) exchange. These results provide a molecular basis for calcium-induced bicarbonate secretion in pancreatic duct cells and highlight the importance of CFTR in epithelial bicarbonate secretion induced by various stimuli.  相似文献   

13.
Synthetic high- and low-molecular-mass atrial peptides were phosphorylated in vitro by cyclic AMP-dependent protein kinase and [32P]ATP. From a series of atrial peptide analogs, it was deduced that the amino acid sequence, Arg101-Ser104 of atriopeptin was required for optimal phosphorylation. Phosphorylated AP(99-126) was less potent than the parent atriopeptin in vasorelaxant activity and receptor-binding properties. These results indicate that the presence of a phosphate group at the N-terminus of AP(99-126) decreases the interaction of the peptide with its receptor and, as a consequence, decreases bioactivity. These observations are in contrast to those of Rittenhouse et al. [(1986) J. Biol. Chem. 261, 7607-7610] who reported that phosphorylation of AP(101-126) enhanced the stimulation of Na/K/Cl cotransport in cultured vascular smooth muscle cells.  相似文献   

14.
The effect of intracellular adenosine triphosphate (ATP) on the “common gating” of the CLC-1 chloride channel has been studied by several laboratories with controversial results. Our previous study on the channel expressed in Xenopus oocytes using excised inside-out patch-clamp methods showed a robust effect of ATP in shifting the open probability curve of the common gate toward more depolarizing voltages (Tseng, P.Y., B. Bennetts, and T.Y. Chen. 2007. J. Gen. Physiol. 130:217–221). The results were consistent with those from studying the channel expressed in mammalian cells using whole cell recording methods (Bennetts, B., M.W. Parker, and B.A. Cromer. 2007. J. Biol. Chem. 282:32780–32791). However, a recent study using excised-patch recording methods for channels expressed in Xenopus oocytes reported that ATP had no direct effect on CLC-1 (Zifarelli, G., and M. Pusch. 2008. J. Gen. Physiol. 131:109–116). Here, we report that oxidation of CLC-1 may be the culprit underlying the controversy. When patches were excised from mammalian cells, the sensitivity to ATP was lost quickly—within 2–3 min. This loss of ATP sensitivity could be prevented or reversed by reducing agents. On the other hand, CLC-1 expressed in Xenopus oocytes lost the ATP sensitivity when patches were treated with oxidizing reagents. These results suggest a novel view in muscle physiology that the mechanisms controlling muscle fatigability may include the oxidation of CLC-1.  相似文献   

15.
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   

16.
Significance of phosphorylation of phosphofructokinase   总被引:1,自引:0,他引:1  
In order to understand the effect of phosphorylation on phosphofructokinase, the allosteric kinetic behavior, ligand binding at various pHs, and pH-dependent cold inactivation of phosphofructokinase phosphorylated to different extents were studied. A subtilisin-digested phosphofructokinase from which a COOH-terminal peptide containing a phosphorylation site has been cleaved (Riquelme, P. T., and Kemp, R. G. (1980) J. Biol. Chem. 255, 4367-4371) was also included in these studies in order to investigate the possible role of this region of the molecule. Allosteric kinetics and direct binding experiments have shown that increasing phosphorylation of phosphofructokinase results in increased sensitivity to ATP inhibition and stronger binding of ATP to the inhibitory site of the enzyme. Ths subtilisin-cleaved phosphofructokinase is the least sensitive to the inhibition and shows the weakest binding of ATP. The opposite effect is observed with the binding isotherms of fructose-6-P. There is no difference in the binding of fructose-2,6-P2 among these enzymes. Binding of ATP to the inhibitory site of these enzymes as determined by fluorescence quenching (Pettigrew, D. W., and Frieden, C. (1979) J. Biol. Chem. 254, 1887-1895) is affected by pH; the binding is greatly enhanced at lower pH. Moreover, there is little difference in the binding among the modified enzymes at pH 8, but at lower pHs the binding to the phosphorylated enzyme is much more enhanced than the dephosphoenzyme. A pH-dependent cold inactivation study has shown that the phosphorylation of the enzyme causes an increase in the pK value for the inactivation, and the extent of the pK shift depends upon the degree of phosphorylation. Based on these results, a model originally proposed by Frieden et al. (Frieden, C., Gilbert, H. R., and Bock, P. E. (1976) J. Biol. Chem. 251, 5644-5647) can be applied to explain a possible role for the phosphorylation and the peptide portion of phosphofructokinase in its complex allosteric kinetic behavior.  相似文献   

17.
Some CLC proteins function as passive Cl(-) ion channels whereas others are secondary active chloride/proton antiporters. Voltage-dependent gating of the model Torpedo channel ClC-0 is modulated by intracellular and extracellular pH, possibly reflecting a mechanistic relationship with the chloride/proton coupling of CLC antiporters. We used inside-out patch clamp measurements and mutagenesis to explore the dependence of the fast gating mechanism of ClC-0 on intracellular pH and to identify the putative intracellular proton acceptor(s). Among the tested residues (S123, K129, R133, K149, E166, F214L, S224, E226, V227, C229, R305, R312, C415, H472, F418, V419, P420, and Y512) only mutants of E166, F214, and F418 qualitatively changed the pH(int) dependence. No tested amino acid emerged as a valid candidate for being a pH sensor. A detailed kinetic analysis of the dependence of fast gate relaxations on pH(int) and [Cl(-)](int) provided quantitative constraints on possible mechanistic models of gating. In one particular model, a proton is generated by the dissociation of a water molecule in an intrapore chloride ion binding site. The proton is delivered to the side chain of E166 leading to the opening of the channel, while the hydroxyl ion is stabilized in the internal/central anion binding site. Deuterium isotope effects confirm that proton transfer is rate limiting for fast gate opening and that channel closure depends mostly on the concentration of OH(-) ions. The gating model is in natural agreement with the finding that only the closing rate constant, but not the opening rate constant, depends on pH(int) and [Cl(-)](int).  相似文献   

18.
Caveolins are scaffolding proteins able to collect on caveolae a large number of signalling proteins bearing a caveolin-binding motif. The proteins of the striatin family, striatin, SG2NA, and zinedin, are composed of several conserved, collinearly aligned, protein-protein association domains, among which a putative caveolin-binding domain [Castets et al. (2000) J. Biol. Chem. 275, 19970-19977]. They are associated in part with membranes. These proteins are mainly expressed within neurons and thought to act both as scaffolds and as Ca(2+)-dependent signalling proteins [Bartoli et al. (1999) J. Neurobiol. 40, 234-243]. Here, we show that (1) rat brain striatin, SG2NA and zinedin co-immunoprecipitate with caveolin-1; (2) all are pulled down by glutathione-S-transferase (GST)-caveolin-1; (3) a fragment of recombinant striatin containing the putative caveolin-binding domain binds GST-caveolin-1. Hence, it is likely that the proteins of the striatin family are addressed to membrane microdomains by their binding to caveolin, in accordance with their putative role in membrane trafficking [Baillat et al. (2001) Mol. Biol. Cell 12, 663-673].  相似文献   

19.
E Muneyuki  H Hirata 《FEBS letters》1988,234(2):455-458
Kinetic analysis of both proton translocating and steady-state ATP hydrolytic activities catalyzed by F0F1 ATPase in submitochondrial particles were carried out over an ATP concentration range of 1-2000 microM. The results were examined in relation to the prediction based on the alternate binding change model proposed by Gresser et al. [(1982) J. Biol. Chem. 257, 12030-12038] in which energy transduction occurs only at the tri-site catalytic cycle. The present results essentially contrast with the model and rather indicate that if the alternate binding mechanism holds for the ATP hydrolytic reaction, the proton translocation should be coupled to at least both bi-site and tri-site cycles.  相似文献   

20.
Under conditions of molar excess of enzyme, isolated F1-ATPase from beef heart mitochondria catalyses ATP hydrolysis biphasically. The rate constants for product release are approximately 10(-1) and 10(-4)-10(-3) s-1, respectively. The slow phase of ATP hydrolysis is insensitive to EDTA. [gamma-32P]ATP splitting in the slow phase cannot be chased from the enzyme during several catalytic turnovers. It follows from these results that the slow single-site hydrolysis of ATP (kcat approximately 10(-4) s-1), initially observed by Grubmeyer et al. [(1982) J. Biol. Chem. 257, 12092-12100], is not carried out by the normal catalytic site. In contrast, the phase of rapid ATP hydrolysis (kcat approximately 10(-1) s-1) is completely prevented by EDTA and is believed to be the normal function of the normal catalytic site of F1-ATPase.  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