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 15 毫秒
1.
We report the cloning and sequencing of a gene encoding the farnesyl pyrophosphate synthase of Trypanosoma cruzi. The protein (T. cruzi farnesyl pyrophosphate synthase, TcFPPS) is an attractive target for drug development, since the growth of T. cruzi is inhibited by carbocation transition state/reactive intermediate analogs of its substrates, the nitrogen-containing bisphosphonates currently in use in bone resorption therapy. The protein predicted from the nucleotide sequence of the gene has 362 amino acids and a molecular mass of 41.2 kDa. Several sequence motifs found in other FPPSs are present in TcFPPS. Heterologous expression of TcFPPS in Escherichia coli produced a functional enzyme that was inhibited by the nitrogen-containing bisphosphonates alendronate, pamidronate, homorisedronate, and risedronate but was less sensitive to the non-nitrogen-containing bisphosphonate etidronate, which, unlike the nitrogen-containing bisphosphonates, does not affect parasite growth. The protein contains a unique 11-mer insertion located near the active site, together with other sequence differences that may facilitate the development of novel anti-Chagasic agents.  相似文献   

2.
The inhibiting effects of several phenol compounds on electron transfer in the respiratory chain of submitochondrial particles were studied. It was shown that the terminal part of the succinate dehydrogenase complex contains a site, which specifically binds the negatively charged phenols (e. g. pentachlorophenol, 2,4-dibromophenol, 2-methoxy-4,6-dibromophenol). The efficiency of the inhibitor anion binding by this site is increased 12-fold after introduction of a methoxy-group into the o-position of 2,4-dibromophenol. Since this site binds both methoxy- and negatively charged phenol groups, it can also act as a possible site of ubisemiquinone interaction with succinate dehydrogenase. Based on the structural similiarity of the b-c1 complex inhibitors, e. g. antimycin, o-hydroxybenzoic acid amides, 2-hydroxy-3-alkyl-1,4-quinones, and ubisemiquinone, an assumption has been made on possible ubisemiquinone binding between cytochromes b and c1. A structural analysis of phenols inhibiting succinate dehydrogenase and the b-c1 complex revealed that in the region between cytochromes b and c1 there act only the inhimitors which contain: a) a negatively charged phenol group; b) a group comprising a heteroatom with an undepleted electron pair and which can act as a ligand, and c) a hydrophobic residue.  相似文献   

3.
It is shown here that Escherichia coli beta-galactosidase has a second Mg2+ binding site that is important for activity. Binding of Mg2+ to the second site caused the k(cat) (with oNPG as the substrate) to increase about 100 s(-1); the Km was not affected. The Kd for binding the second Mg2+ is about 10(-4)M. Since the concentration of free Mg2+ in E. coli is about 1-2 mM, the second site is physiologically significant. Non-polar substitutions (Ala or Leu) for Glu-797, a residue in an active site loop, eliminated the k(cat) increase. This indicates that the second Mg2+ site is near to Glu-797. The Ki values of transition state analogs were decreased by small but statistically significant amounts when the second Mg2+ site was occupied and Arrhenius plots showed that less entropic activation energy is required when the second site is occupied. These inhibitor and temperature results suggest that binding of the second Mg2+ helps to order the active site for stabilization of the transition state.  相似文献   

4.
Ala-Arg-Pro-Ala-Lys (ARPAK; also known as P6A) and 19 of its analogs were synthesized, and their thrombolytic activities were assessed in vitro and in vivo. The solution structures of 12 of the P6A analogs were determined using nuclear magnetic resonance (NMR) spectroscopy. The thrombolytic activity and conformational structure relationship was analyzed. We found that the Pro-Ala-Lys (PAK) sequence was essential for thrombolytic activity and was also responsible for the beta-turn structure found in the P6A analogs studied. The well defined beta turn may act as a binding head with the protruding lysine side-chain (positively charged) found at the target site for target recognition. Additionally, the N-terminal residue may be critical for thrombolytic activity, which for PAK-containing peptides, is likely achieved via a plasminogen-dependent pathway.  相似文献   

5.
Antibody 4C6 efficiently catalyzes a cationic cyclization reaction. Crystal structures of the antibody 4C6 Fab in complex with benzoic acid and in complex with its eliciting hapten were determined to 1.30A and 2.45A resolution, respectively. These crystal structures, together with computational analysis, have elucidated a possible mechanism for the monocyclization reaction. The hapten complex revealed a combining site pocket with high shape complementarity to the hapten. This active site cleft is dominated by aromatic residues that shield the highly reactive carbocation intermediates from solvent and stabilize the carbocation intermediates through cation-pi interactions. Modeling of an acyclic olefinic sulfonate ester substrate and the transition state (TS) structures shows that the chair-like transition state is favored, and trapping by water directly produces trans-2-(dimethylphenylsilyl)-cyclohexanol, whereas the less favored boat-like transition state leads to cyclohexene. The only significant change observed upon hapten binding is a side-chain rotation of Trp(L89), which reorients to form the base of the combining site. Intriguingly, a benzoic acid molecule was sequestered in the combining site of the unliganded antibody. The 4C6 active site was compared to that observed in a previously reported tandem cyclization antibody 19A4 hapten complex. These cationic cyclization antibodies exhibit convergent structural features with terpenoid cyclases that appear to be important for catalysis.  相似文献   

6.
We determined the structural features necessary for fatty acids to exert their action on K+ channels of gastric smooth muscle cells. Examination of the effects of a variety of synthetic and naturally occurring lipid compounds on K+ channel activity in cell-attached and excised membrane patches revealed that negatively charged analogs of medium to long chain fatty acids (but not short chain analogs) as well as certain other negatively charged lipids activate the channels. In contrast, positively charged, medium to long chain analogs suppress activity, and neutral analogs are without effect. The key requirements for effective compounds seem to be a sufficiently hydrophobic domain and the presence of a charged group. Furthermore, those negatively charged compounds unable to "flip" across the bilayer are effective only when applied at the cytosolic surface of the membrane, suggesting that the site of fatty acid action is also located there. Finally, because some of the effective compounds, for example, the fatty acids themselves, lysophosphatidate, acyl Coenzyme A, and sphingosine, are naturally occurring substances and can be liberated by agonist- activated or metabolic enzymes, they may act as second messengers targeting ion channels.  相似文献   

7.
Transition state analogs pepstatin methylester (PME) and L685458 have been shown to inhibit gamma-secretase non-competitively (Tian, G., Sobotka-Briner, C., Zysk, J., Liu, X., Birr, C., Sylvester, M. A., Edwards, P. D., Scott, C. W., and Greenberg, B. D. (2002) J. Biol. Chem. 277, 31499-31505). This unusual kinetics suggests physical separation of the sites for substrate binding and catalysis with binding of the transition state analogs to the catalytic site and not to the substrate binding site. Methods of inhibitor cross-competition kinetics and competition ligand binding were utilized to address whether non-transition state small molecule inhibitors, which also display non-competitive inhibition of gamma-secretase, inhibit the enzyme by binding to the catalytic site as well. Inhibitor cross-competition kinetics indicated competitive binding between the transition state analogs PME and L685458 and between small molecules arylsulfonamides and benzodiazepines, but non-competitive binding between the transition state analogs and the small molecule inhibitors. These results were indicative of two inhibitor binding sites, one for transition state analogs and the other for non-transition state small molecule inhibitors. The presence of two inhibitor binding sites for two different classes of inhibitors was corroborated by results from competition ligand binding using [3H]L685458 as the radioligand. Although L685458 and PME displaced the radioligand at the same concentrations as for enzyme inhibition, arylsulfonamides and benzodiazepines did not displace the radioligand at their Ki values, a result consistent with the presence of two inhibitor binding sites. These findings provide useful insights into the catalytic and regulatory mechanisms of gamma-secretase that may facilitate the design of novel gamma-secretase inhibitors.  相似文献   

8.
A series of transition state analogues of beta-secretases 1 and 2 (BACE1, 2) inhibitors containing fused-ring or biaryl moieties were designed computationally to probe the S2 pocket, synthesized, and tested for BACE1 and BACE2 inhibitory activity. It has been shown that unlike the biaryl analogs, the fused-ring moiety is successfully accommodated in the BACE1 binding site resulting in the ligands with excellent inhibitory activity. Ligand 5b reduced 65% of Aβ40 production in N2a cells stably transfected with Swedish human APP.  相似文献   

9.
The interaction between modifying ATP analogs containing alkylating or phosphorylating groups in the polyphosphate moiety of the ATP molecule and leucyl-tRNA synthetases from cytoplasm and chloroplasts of Euglena gracilis (strain Z) was studied. It was shown that most of the ATP analogs irreversibly inhibit the cytoplasmic enzyme, having no inhibiting effect on the chloroplast synthetase. The kinetic constants K1 and k2 for the interaction between the most effective irreversible inhibitors and the cytoplasmic enzyme were determined. The data on the protection of the enzyme activity by substrates against irreversible inhibition suggest, that the effect of the adenosine 5'-(beta-chloroethyl phosphate) is directed to the ATP-binding site of the cytoplasmic enzyme, whereas the mixed anhydride of AMP and mesithylene carbonic acid acts predominantly on the binding site of 3'-terminal adenosine of the tRNALeu molecule. ATP analogs may be effectively used for affinity labelling of the cytoplasmic leucyl-tRNA synthetase.  相似文献   

10.
This paper presents a study of the role of positive charge in the P(i) binding site of Escherichia coli ATP synthase, the enzyme responsible for ATP-driven proton extrusion and ATP synthesis by oxidative phosphorylation. Arginine residues are known to occur with high propensity in P(i) binding sites of proteins generally and in the P(i) binding site of the betaE catalytic site of ATP synthase specifically. Removal of natural betaArg-246 (betaR246A mutant) abrogates P(i) binding; restoration of P(i) binding was achieved by mutagenesis of either residue betaAsn-243 or alphaPhe-291 to Arg. Both residues are located in the P(i) binding site close to betaArg-246 in x-ray structures. Insertion of one extra Arg at beta-243 or alpha-291 in presence of betaArg-246 retained P(i) binding, but insertion of two extra Arg, at both positions simultaneously, abrogated it. Transition state stabilization was measured using phosphate analogs fluoroaluminate and fluoroscandium. Removal of betaArg-246 in betaR246A caused almost complete loss of transition state stabilization, but partial rescue was achieved in betaN243R/betaR246A and alphaF291R/betaR246A. BetaArg-243 or alphaArg-291 in presence of betaArg-246 was less effective; the combination of alphaF291R/betaN243R with natural betaArg-246 was just as detrimental as betaR246A. The data demonstrate that electrostatic interaction is an important component of initial P(i) binding in catalytic site betaE and later at the transition state complex. However, since none of the mutants showed significant function in growth tests, ATP-driven proton pumping, or ATPase activity assays, it is apparent that specific stereochemical interactions of catalytic site Arg residues are paramount.  相似文献   

11.
Cao R  Chen CK  Guo RT  Wang AH  Oldfield E 《Proteins》2008,73(2):431-439
We report the X-ray crystallographic structures of the bisphosphonate N-[methyl(4-phenylbutyl)]-3-aminopropyl-1-hydroxy-1,1-bisphosphonate (BPH-210), a potent analog of pamidronate (Aredia), bound to farnesyl diphosphate synthase (FPPS) from Trypanosoma brucei as well as to geranylgeranyl diphosphate synthase from Saccharomyces cerevisiae. BPH-210 binds to FPPS, together with 3 Mg(2+), with its long, hydrophobic phenylbutyl side-chain being located in the same binding pocket that is occupied by allylic diphosphates and other bisphosphonates. Binding is overwhelmingly entropy driven, as determined by isothermal titration calorimetry. The structure is of interest since it explains the lack of potency of longer chain analogs against FPPS, since these would be expected to have a steric clash with an aromatic ring at the distal end of the binding site. Unlike shorter chain FPPS inhibitors, such as pamidronate, BPH-210 is also found to be a potent inhibitor of human geranylgeranyl diphosphate synthase. In this case, the bisphosphonate binds only to the GGPP product inhibitory site, with only 1 (chain A) or 0 (chain B) Mg(2+), and DeltaS is much smaller and DeltaH is approximately 6 k cal more negative than in the case of FPPS binding. Overall, these results are of general interest since they show that some bisphosphonates can bind to more than one trans-prenyl synthase enzyme which, in some cases, can be expected to enhance their overall activity in vitro and in vivo.  相似文献   

12.
Complexes formed from A13+ or Be2+ and fluoride inhibit the single-stranded DNA-dependent ATPase activity of RecA protein. In contrast, poly(dT)-RecA-ADP complexes, which are inactive for cleavage of LexA protein, become fully active in the presence of AlF4- or BeF3- ions. These data suggest that fluoride complexes of aluminum and beryllium (called herein X) convert RecA-ADP complexes, which bind weakly to single-stranded DNA, into RecA-ADP-X complexes, which bind tightly to single-stranded DNA, the ADP-X moiety behaving as a nonhydrolyzable analogue of ATP. We propose that AlF4- and BeF3- ions act as analogues of inorganic phosphate by binding to the site of the gamma-phosphate of ATP on RecA-ADP complexes, hence mimicking the single-stranded DNA-RecA-ADP-Pi transition state. We conclude that the elementary reaction that switches RecA protein from a high affinity single-stranded DNA binding state to a low affinity single-stranded DNA binding state is not ATP hydrolysis per se but Pi release.  相似文献   

13.
The activity of nitrate reductase (+Mg(2+), NR(act)) in illuminated leaves from spinach, barley and pea was 50-80% of the maximum activity (+EDTA, NR(max)). However, NR from leaves of Ricinus communis L. had a 10-fold lower NR(act), while NR(max) was similar to that in spinach leaves. The low NR(act) of Ricinus was independent of day-time and nitrate nutrition, and varied only slightly with leaf age. Possible factors in Ricinus extracts inhibiting NR were not found. NR(act) from Ricinus, unlike the spinach enzyme, was very low at pH 7.6, but much higher at more acidic pH with a distinct maximum at pH 6.5. NR(max) had a broad pH response profile that was similar for the spinach and the Ricinus enzyme. Accordingly, the Mg(2+)-sensitivity of NR from Ricinus was strongly pH-dependent (increasing sensitivity with increasing pH), and as a result, the apparent activation state of NR from a Ricinus extract varied dramatically with pH and Mg(2+)concentration. Following a light-dark transition, NR(act) from Ricinus decreased within 1 h by 40%, but this decrease was paralleled by NR(max). In contrast to the spinach enzyme, Ricinus-NR was hardly inactivated by incubating leaf extracts with ATP plus okadaic acid. A competition analysis with antibodies against the potential 14-3-3 binding site around ser 543 of the spinach enzyme revealed that Ricinus-NR contains the same site. Removal of 14-3-3 proteins from Ricinus-NR by anion exchange chromatography, activated spinach-NR but caused little if any activation of Ricinus-NR. It is suggested that Mg(2+)-inhibition of Ricinus-NR does not require 14-3-3 proteins. The rather slow changes in Ricinus-NR activity upon a light/dark transient may be mainly due to NR synthesis or degradation.  相似文献   

14.
The effects of several phosphorylating and alkylating analogs of the substrate on the ATPase activity of myosin and heavy meromyosin were compared. The data obtained confirmed the previously made assumption on the existence of two types of substrate-like inhibitor binding sites in the enzyme molecule. In one of the sites, presumably in the active one, there occurs a reversible competitive inhibition characterized by a high affinity for the inhibitors, which are mixed anhydrides of various mononucleotides and mesitylcarboxylic acid or its derivatives. An enhancement of hydrophobicity of these compounds causes an increase in their affinity for this site. At much higher concentrations of the inhibitors an irreversible inhibition takes place, the rate of inhibition being decreased with an increase in the phosphorylating capacity of the compound. This site possesses a far lower affinity for the inhibitors and reveals a certain specificity with respect to the analog mononucleotide moiety structure, i.e. a replacement of the 6-NH2-group by the 6-OH-group or an increase in the number of the phosphate residues result in a decrease of the efficiency of inhibition. No correlation between the analog capacity to cause irreversible inhibition and to act as an effective competitive inhibitor of reversible type has been shown to exist, thus allowing to use inhibitors of preferable action in one of the two types of the binding sites. No irreversible inhibition site was revealed when the ATPase activity of myosin subfragment I with and without the DTNB chains was investigated. Actin protects myosin against the inhibiting action of the analogs tested.  相似文献   

15.
Stable chemical analogues of enzymatic transition states are imperfect mimics since they lack the partial bond character of the transition state. We synthesized structural variants of the Immucillins as transition state analogues for purine nucleoside phosphorylase and characterized them with the enzyme from Mycobacterium tuberculosis (MtPNP). PNPs form transition states with ribooxacarbenium ion character and catalyze nucleophilic displacement reactions by migration of the cationic ribooxacarbenium carbon between the enzymatically immobilized purine and phosphate nucleophiles. As bond-breaking progresses, carbocation character builds on the ribosyl group, the distance between the purine and the carbocation increases, and the distance between carbocation and phosphate anion decreases. Transition state analogues were produced with carbocation character and increased distance between the ribooxacarbenium ion and the purine mimics by incorporating a methylene bridge between these groups. Immucillin-H (ImmH), DADMe-ImmH, and DADMe-ImmG mimic the transition state of MtPNP and are slow-onset, tight-binding inhibitors of MtPNP with equilibrium dissociation constants of 650, 42, and 24 pM. Crystal structures of MtPNP complexes with ImmH and DADMe-ImmH reveal an ion-pair between the inhibitor cation and the nucleophilic phosphoryl anion. The stronger ion-pair (2.7 A) is found with DADMe-ImmH. The position of bound ImmH resembles the substrate side of the transition state barrier, and DADMe-ImmH more closely resembles the product side of the barrier. The ability to probe both substrate and product sides of the transition state barrier provides expanded opportunities to explore transition state analogue design in N-ribosyltransferases. This approach has resulted in the highest affinity transition state analogues known for MtPNP.  相似文献   

16.
Some information about the lactate dehydrogenase NAD binding site has been obtained by working with coenzymes analogs of incomplete molecules. 5'AMP, 5'-ADP, ATP, 5'-c-AMP and 3'(2)-AMP inhibit chicken liver LDH activity competitively with NADH. 5"-AMP and 5'-ADP show a stronger inhibition power than ATP, suggesting that the presence of one or two phosphate groups at the 5' position of adenosine, is essential for the binding of the coenzyme analogs at the enzyme binding site. Ribose and ribose-5'-P do not appear to inhibit the LDH activity, proving that purine base lacking mononucleotides do not bind to the enzyme. 5"-ADPG inhibits LDH activity in the exactly as 5'-ADP, showing that ribose moiety may be replaced by glucose, without considerable effects on the coenzyme analog binding. 2'-desoxidenosin-5'-phosphate proves to be a poorer inhibitor of the LDH activity than 5'-AMP, indicating that an interaction between the--OH groups and the amino-acids of the LDH active center takes place. Nicotinamide does not produce any inhibition effect, while NMN and CMP induce a much weaker inhibition than the adenine analogues, thus indicating a lesser binding capacity to the enzyme. Therefore, the LDH binding site seems to show some definite specificity towards the adenina groups of the coenzyme.  相似文献   

17.
Jiang YL  Ichikawa Y  Song F  Stivers JT 《Biochemistry》2003,42(7):1922-1929
The reaction catalyzed by the DNA repair enzyme uracil DNA glycosylase (UDG) proceeds through an unprecedented stepwise mechanism involving a positively charged oxacarbenium ion sugar and uracil anion leaving group. Here we use a novel approach to evaluate the catalytic contribution of electrostatic interactions between four essential phosphodiester groups of the DNA substrate and the cationic transition state. Our strategy was to substitute each of these phosphate groups with an uncharged (R)- or (S)-methylphosphonate linkage (MeP). We then compared the damaging effects of these methylphosphonate substitutions on catalysis with their damaging effects on binding of a cationic 1-azadeoxyribose (1-aza-dR(+)) oxacarbenium ion analogue to the UDG-uracil anion binary complex. A plot of log k(cat)/K(m) for the series of MeP-substituted substrates against log K(D) for binding of the 1-aza-dR(+) inhibitors gives a linear correlation of unit slope, confirming that the electronic features of the transition state resemble that of the 1-aza-dR(+), and that the anionic backbone of DNA is used in transition state stabilization. We estimate that all of the combined phosphodiester interactions with the substrate contribute 6-8 kcal/mol toward lowering the activation barrier, a stabilization that is significant compared to the 16 kcal/mol catalytic power of UDG. However, unlike groups of the enzyme that selectively stabilize the charged transition state by an estimated 7 kcal/mol, these phosphodiester groups also interact strongly in the ground state. To our knowledge, these results provide the first experimental evidence for electrostatic stabilization of a charged enzymatic transition state and intermediate using the anionic backbone of DNA.  相似文献   

18.
The influence of the binding of poly(l-lysine) (PLL) to negatively charged membranes containing phosphatidylglycerols (PG) was studied by DSC and FT-IR spectroscopy. We found a general increase in the main transition temperature as well as increase in hydrophobic order of the membrane upon PLL binding. Furthermore we observed stronger binding of hydration water to the lipid head groups after PLL binding. The secondary structure of the PLL after binding was studied by FT-IR spectroscopy. We found that PLL binds in an α-helical conformation to negatively charged DPPG membranes or membranes with DPPG-rich domains. Moreover we proved that PLL binding induces domain formation in the gel state of mixed DPPC/DPPG or DMPC/DPPG membranes as well as lipid remixing in the liquid–crystalline state. We studied these effects as a function of PLL chain length and found a significant dependence of the secondary structure, phase transition temperature and domain formation capacity on PLL chain length and also a correlation between the peptide secondary structure and the phase transition temperature of the membrane. We present a system in which the membrane phase transition triggers a highly cooperative secondary structure transition of the membrane-bound peptide from α-helix to random coil. Dedicated to Prof. K. Arnold on the occasion of his 65th birthday.  相似文献   

19.
We used a series of adamantane derivatives to probe the structure of the phencyclidine locus in either the resting or desensitized state of the nicotinic acetylcholine receptor (AChR). Competitive radioligand binding and photolabeling experiments using well-characterized noncompetitive antagonists such as the phencyclidine analogue [piperidyl-3,4-(3)H(N)]-N-[1-(2-thienyl)cyclohexyl]-3,4-piperidine ([(3)H]TCP), [(3)H]ethidium, [(3)H]tetracaine, [(14)C]amobarbital, and 3-(trifluoromethyl)-3-(m-[(125)I]iodophenyl)diazirine ([(125)I]TID) were performed. Thermodynamic and structure-function relationship analyses yielded the following results. (1) There is a good structure-function relationship for adamantane amino derivatives inhibiting [(3)H]TCP or [(3)H]tetracaine binding to the resting AChR. (2) Since the same derivatives inhibit neither [(14)C]amobarbital binding nor [(125)I]TID photoincorporation, we conclude that these positively charged molecules preferably bind to the TCP locus, perhaps interacting with alphaGlu(262) residues at position M2-20. (3) The opposite is true for the neutral molecule adamantane, which prefers the TID (or barbiturate) locus instead of the TCP site. (4) The TID site is smaller and more hydrophobic (it accommodates neutral molecules with a maximal volume of 333 +/- 45 A(3)) than the TCP locus, which has room for positively charged molecules with volumes as large as 461 A(3) (e.g., crystal violet). This supports the concept that the resting ion channel is tapering from the extracellular mouth to the middle portion. (5) Finally, although both the hydrophobic environment and the size of the TCP site are practically the same in both states, there is a more obvious cutoff in the desensitized state than in the resting state, suggesting that the desensitization process constrains the TCP locus. A plausible location of neutral and charged adamantane derivatives is shown in a model of the resting ion channel.  相似文献   

20.
Crystal structures have shown that the conserved TGES loop of the Ca2+-ATPase is isolated in the Ca2E1 state but becomes inserted in the catalytic site in E2 states. Here, we have examined the kinetics of the partial reaction steps of the transport cycle and the binding of the phosphoryl analogs BeF, AlF, MgF, and vanadate in mutants with alterations to the TGES residues. The mutations encompassed variation of size, polarity, and charge of the side chains. Differential effects on the Ca2E1P --> E2P, E2P --> E2, and E2 --> Ca2E1 reactions and the binding of the phosphoryl analogs were observed. In the E183D mutant, the E2P --> E2 dephosphorylation reaction proceeded at a rate as high as one-third that of the wild type, whereas it was very slow in the other Glu183 mutants, including E183Q, thus demonstrating the need for a negatively charged carboxylate group to catalyze dephosphorylation. By contrast, the Ca2E1P --> E2P transition was accomplished at a reasonable rate with glutamine in place of Glu183, but not with aspartate, indicating that the length of the Glu183 side chain, in addition to its hydrogen bonding potential, is critical for Ca2E1P --> E2P. This transition was also slowed in mutants with alteration to other TGES residues. The data provide functional evidence in support of the proposed role of Glu183 in activating the water molecule involved in the E2P --> E2 dephosphorylation and suggest a direct participation of the side chains of the TGES loop in the control and facilitation of the insertion of the loop in the catalytic site. The interactions of the TGES loop furthermore seem to facilitate its disengagement from the catalytic site during the E2 --> Ca2E1 transition.  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