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 62 毫秒
1.
用浸泡法得到了(E160A)天花粉蛋白(trichosanthin, TCS),(E160D)TCS与Ade 和(E160A)TCS与FMP复合物的晶体.在Mar Research 面探测器系统上分别收集了0.20 nm ,0.19nm 和0.205 nm 分辨率的X 射线衍射数据,数据处理用Mar Scale 程序系统完成.用同晶差值Fourier法解析了(E160A)TCS-Ade,(E160D)TCS-Ade 和(E160A)TCS-FMP的晶体结构,结构修正利用X-PLOR程序,修正结果,晶体学R因子分别为0.166,0.176,0.179.键长和键角的RMS偏差分别为0.0010 nm 和2.503°,0.0013 nm 和2.665°,0.0012 nm 和2.676°.在这三个结构中均未见到Glu189侧链方向的改变.Ade 或FMP仍结合在N-糖苷酶活性口袋之中,它夹在Tyr70和Tyr111两个侧链环之间,与Tyr70环近乎平行.这一结果表明:TCS中的Glu160分别突变成Ala 和Asp,仍能与AMP发生N-糖苷酶反应,但是活性降低了一些.可见Glu160对TCS与AMP的作用是重要的,但不是必要的.  相似文献   

2.
培养了(E160A,E189A)TCS(天花粉蛋白)的单晶。用浸泡法得到了(E160A,E189A)TCS与Ade复合物的晶体。在MarResearch面探测器系统上分别收集了均为0.20nm分辨率的X射线衍射数据,数据处理用MarScale程序系统完成。用同晶差值Fourier法解析了(E160A,E189A)TCS和(E160A,E189A)TCS-Ade的晶体结构,结构修正利用X-PLOR程序.修正结果,晶体学R因子分别为0.180、0.184,键长和键角的RMS偏差分别为0.0012nm和2.566°、0.0012nm和2.622°。在(E160A,E189A)TCS-Ade中,Ade仍结合在N-糖苷酶活性口袋之中,它夹在Tyr70和Tyr111两个侧链环之间,与Tyr70环近乎平行。这一结果表明:TCS中的Glu160和Glu189同时突变成Ala,仍能与AMP发生N-糖苷酶反应.前文已经证明在(E160A)TCS中Glu189没有援救作用。目前,没有发现Glu189对TCS与AMP的直接作用,但Glu189与其它残基的协同作用及其在TCS与rRNA作用中扮演什么角色,尚待进一步研究。  相似文献   

3.
The enzyme adenine phosphoribosyltransferase (APRT) functions to salvage adenine by converting it to adenosine-5-monophosphate (AMP). APRT deficiency in humans is a well characterized inborn error of metabolism, and APRT may contribute to the indispensable nutritional role of purine salvage in protozoan parasites, all of which lack de novo purine biosynthesis. We determined crystal structures for APRT from Leishmania donovani in complex with the substrate adenine, the product AMP, and sulfate and citrate ions that appear to mimic the binding of phosphate moieties. Overall, these structures are very similar to each other, although the adenine and AMP complexes show different patterns of hydrogen-bonding to the base, and the active site pocket opens slightly to accommodate the larger AMP ligand. Whereas AMP adopts a single conformation, adenine binds in two mutually exclusive orientations: one orientation providing adenine-specific hydrogen bonds and the other apparently positioning adenine for the enzymatic reaction. The core of APRT is similar to that of other phosphoribosyltransferases, although the adenine-binding domain is quite different. A C-terminal extension, unique to Leishmania APRTs, extends an extensive dimer interface by wrapping around the partner molecule. The active site involves residues from both subunits of the dimer, indicating that dimerization is essential for catalysis.  相似文献   

4.
Trichosanthin (TCS) is a ribosome-inactivating protein (RIP) that possesses N-glycosidase activity. It inactivates ribosomes and arrests protein synthesis by removing a specific adenine from 28S rRNA. A molecular dynamics simulated annealing method was applied to study the binding modes of TCS with substrate analogs, three oligonucleotides GAG, GAGA, and CGAGAG, based on the crystal structures of the stable complexes of TCS with NADPH and with the reaction product adenine. A water molecule proposed to be responsible for hydrolyzing the N-glycosidic bond was included in the model. All the oligoribonucleotides can dock into the active cleft of TCS without unfavorable contacts. The interaction energies between TCS and the three oligonucleotides were calculated. The interactions of TCS with NADH were also studied by a molecular dynamics simulated annealing method. The interaction energy between NADH and TCS was compared with that between NADPH and TCS, showing that the lack of 2-phosphate group leads to an energy rise of 20 kcal/mol.  相似文献   

5.
To elucidate the structural basis for the alteration of coenzyme specificity from NADH toward NADPH in a malate dehydrogenase mutant EX7 from Thermus flavus, we determined the crystal structures at 2.0 A resolution of EX7 complexed with NADPH and NADH, respectively. In the EX7-NADPH complex, Ser42 and Ser45 form hydrogen bonds with the 2'-phosphate group of the adenine ribose of NADPH, although the adenine moiety is not seen in the electron density map. In contrast, although Ser42 and Ser45 occupy a similar position in the EX7-NADH complex structure, both the adenine and adenine ribose moieties of NADH are missing in the map. These results and kinetic analysis of site-directed mutant enzymes indicate (1) that the preference of EX7 for NADPH over NADH is ascribed to the recognition of the 2'-phosphate group by two Ser and Arg44, and (2) that the adenine moiety of NADPH is not recognized in this mutant.  相似文献   

6.
Although the activation of low-molecular weight protein tyrosine phosphatases by certain purines and purine derivatives was first described three decades ago, the mechanism of this rate enhancement was unknown. As an example, adenine activates the yeast low-molecular weight protein tyrosine phosphatase LTP1 more than 30-fold. To examine the structural and mechanistic basis of this phenomenon, we have determined the crystal structure of yeast LTP1 complexed with adenine. In the crystal structure, an adenine molecule is found bound in the active site cavity, sandwiched between the side chains of two large hydrophobic residues at the active site. Hydrogen bonding to the side chains of other active site residues, as well as some water-mediated hydrogen bonds, also helps to fix the position of the bound adenine molecule. An ordered water was found in proximity to the bound phosphate ion present in the active site, held by hydrogen bonding to N3 of adenine and Odelta1 of Asp-132. On the basis of the crystal structure, we propose that this water molecule is the nucleophile that participates in the dephosphorylation of the phosphoenzyme intermediate. Solvent isotope effect studies show that there is no rate-determining transfer of a solvent-derived proton in the transition state for the dephosphorylation of the phosphoenzyme intermediate. Such an absence of general base catalysis of water attack is consistent with the stability of the leaving group, namely, the thiolate anion of Cys-13. Consequently, adenine activates the enzyme by binding and orienting a water nucleophile in proximity to the phosphoryl group of the phosphoenzyme intermediate, thus increasing the rate of the dephosphorylation step, a step that is normally the rate-limiting step of this enzymatic reaction.  相似文献   

7.
A bacterial flavin-containing monooxygenase (bFMO) catalyses the oxygenation of indole to produce indigoid compounds. In the reductive half of the indole oxygenation reaction, NADPH acts as a reducing agent, and NADP(+) remains at the active site, protecting bFMO from reoxidation. Here, the crystal structures of bFMO and bFMO in complex with NADP(+), and a mutant bFMO(Y207S), which lacks indole oxygenation activity, with and without indole are reported. The crystal structures revealed overlapping binding sites for NADP(+) and indole, suggestive of a double-displacement reaction mechanism for bFMO. In biochemical assays, indole inhibited NADPH oxidase activity, and NADPH in turn inhibited the binding of indole and decreased indoxyl production. Comparison of the structures of bFMO with and without bound NADP(+) revealed that NADPH induces conformational changes in two active site motifs. One of the motifs contained Arg-229, which participates in interactions with the phosphate group of NADPH and appears be a determinant of the preferential binding of bFMO to NADPH rather than NADH. The second motif contained Tyr-207. The mutant bFMO(Y207S) exhibited very little indoxyl producing activity; however, the NADPH oxidase activity of the mutant was higher than the wild-type enzyme. It suggests a role for Y207, in the protection of hydroperoxyFAD. We describe an indole oxygenation reaction mechanism for bFMO that involves a ping-pong-like interaction of NADPH and indole.  相似文献   

8.
Bovine CD38/NAD(+)glycohydrolase (bCD38) catalyses the hydrolysis of NAD(+) into nicotinamide and ADP-ribose and the formation of cyclic ADP-ribose (cADPR). We solved the crystal structures of the mono N-glycosylated forms of the ecto-domain of bCD38 or the catalytic residue mutant Glu218Gln in their apo state or bound to aFNAD or rFNAD, two 2'-fluorinated analogs of NAD(+). Both compounds behave as mechanism-based inhibitors, allowing the trapping of a reaction intermediate covalently linked to Glu218. Compared to the non-covalent (Michaelis) complex, the ligands adopt a more folded conformation in the covalent complexes. Altogether these crystallographic snapshots along the reaction pathway reveal the drastic conformational rearrangements undergone by the ligand during catalysis with the repositioning of its adenine ring from a solvent-exposed position stacked against Trp168 to a more buried position stacked against Trp181. This adenine flipping between conserved tryptophans is a prerequisite for the proper positioning of the N1 of the adenine ring to perform the nucleophilic attack on the C1' of the ribofuranoside ring ultimately yielding cADPR. In all structures, however, the adenine ring adopts the most thermodynamically favorable anti conformation, explaining why cyclization, which requires a syn conformation, remains a rare alternate event in the reactions catalyzed by bCD38 (cADPR represents only 1% of the reaction products). In the Michaelis complex, the substrate is bound in a constrained conformation; the enzyme uses this ground-state destabilization, in addition to a hydrophobic environment and desolvation of the nicotinamide-ribosyl bond, to destabilize the scissile bond leading to the formation of a ribooxocarbenium ion intermediate. The Glu218 side chain stabilizes this reaction intermediate and plays another important role during catalysis by polarizing the 2'-OH of the substrate NAD(+). Based on our structural analysis and data on active site mutants, we propose a detailed analysis of the catalytic mechanism.  相似文献   

9.
To study the flexibility of the substrate-binding site and in particular of Gln262, we have performed adiabatic conformational search and molecular dynamics simulations on the crystal structure of the catalytic domain of wild-type protein-tyrosine phosphatase (PTP) 1B, a mutant PTP1B(R47V,D48N,M258C,G259Q), and a model of the catalytically active form of PTPalpha. For each molecule two cases were modeled: the Michaelis-Menten complex with the substrate analogue p-nitrophenyl phosphate (p-PNPP) bound to the active site and the cysteine-phosphor complex, each corresponding to the first and second step of the phosphate hydrolysis. Analyses of the trajectories revealed that in the cysteine-phosphor complex of PTP1B, Gln262 oscillates freely between the bound phosphate group and Gly259 frequently forming, as observed in the crystal structure, a hydrogen bond with the backbone oxygen of Gly259. In contrast, the movement of Gln262 is restricted in PTPalpha and the mutant due to interactions with Gln259 reducing the frequency of the oscillation of Gln262 and thereby delaying the positioning of this residue for the second step in the catalysis, as reflected experimentally by a reduction in k(cat). Additionally, in the simulation with the Michaelis-Menten complexes, we found that a glutamine in position 259 induces steric hindrance by pushing the Gln262 side chain further toward the substrate and thereby negatively affecting K(m) as indicated by kinetic studies. Detailed analysis of the water structure around Gln262 and the active site Cys215 reveals that the probability of finding a water molecule correctly positioned for catalysis is much larger in PTP1B than in PTP1B(R47V,D48N,M258C,G259Q) and PTPalpha, in accordance with experiments.  相似文献   

10.
Escherichia coli dihydrofolate reductase (DHFR) has several flexible loops surrounding the active site that play a functional role in substrate and cofactor binding and in catalysis. We have used heteronuclear NMR methods to probe the loop conformations in solution in complexes of DHFR formed during the catalytic cycle. To facilitate the NMR analysis, the enzyme was labeled selectively with [(15)N]alanine. The 13 alanine resonances provide a fingerprint of the protein structure and report on the active site loop conformations and binding of substrate, product, and cofactor. Spectra were recorded for binary and ternary complexes of wild-type DHFR bound to the substrate dihydrofolate (DHF), the product tetrahydrofolate (THF), the pseudosubstrate folate, reduced and oxidized NADPH cofactor, and the inactive cofactor analogue 5,6-dihydroNADPH. The data show that DHFR exists in solution in two dominant conformational states, with the active site loops adopting conformations that closely approximate the occluded or closed conformations identified in earlier X-ray crystallographic analyses. A minor population of a third conformer of unknown structure was observed for the apoenzyme and for the disordered binary complex with 5,6-dihydroNADPH. The reactive Michaelis complex, with both DHF and NADPH bound to the enzyme, could not be studied directly but was modeled by the ternary folate:NADP(+) and dihydrofolate:NADP(+) complexes. From the NMR data, we are able to characterize the active site loop conformation and the occupancy of the substrate and cofactor binding sites in all intermediates formed in the extended catalytic cycle. In the dominant kinetic pathway under steady-state conditions, only the holoenzyme (the binary NADPH complex) and the Michaelis complex adopt the closed loop conformation, and all product complexes are occluded. The catalytic cycle thus involves obligatory conformational transitions between the closed and occluded states. Parallel studies on the catalytically impaired G121V mutant DHFR show that formation of the closed state, in which the nicotinamide ring of the cofactor is inserted into the active site, is energetically disfavored. The G121V mutation, at a position distant from the active site, interferes with coupled loop movements and appears to impair catalysis by destabilizing the closed Michaelis complex and introducing an extra step into the kinetic pathway.  相似文献   

11.
Gu YJ  Xia ZX 《Proteins》2000,39(1):37-46
Four substrate analogs-nicotinamide adenine dinucleotide, adenylyl (3', 5') guanosine, guanylyl (3',5') adenosine, and adenosine 2', 5'-diphosphate-have been used to prepare the complexes with trichosanthin (TCS), a type I ribosome-inactivating protein that possesses the activity of N-glycosidase. The crystal structures of the complexes have been determined and refined at high resolution. The refined structures show that the N-glycosidic bonds of all the four substrate analogues are hydrolyzed and a common structure is shared by the four complexes, in which only adenine, the product of the enzymatic reaction, is bound in the active center. The structure is compared with those of native trichosanthin and a previously reported trichosanthin-NADPH complex in which the N-glycosidic bond is uncleaved. The structural comparison shows that the conformation of Tyr70 obviously differs from those in the latter two structures, i.e., the side chain of Tyr70 is rotated along its Cbeta-Cgamma bond by approximately 70 degrees. The water molecule found to be preassociated with the N-glycosidic bond in the TCS-NADPH complex structure and proposed to be the water candidate responsible for hydrolyzing the N-glycosidic bond disappears in the trichosanthin-product complex structure. Based on the comparison of the three structures representing the different stages of the enzymatic reaction, the catalytic mechanism of RNA N-glycosidase has been further elucidated. Proteins 2000;39:37-46.  相似文献   

12.
Dihydrofolate reductase (DHFR) catalyzes the nicotinamide adenine dinucleotide phosphate (NADPH)-dependent reduction of 7,8-dihydrofolate (H2F) to 5,6,7,8-tetrahydrofolate (H4F). Because of the absence of any ionizable group in the vicinity of N5 of dihydrofolate it has been proposed that N5 could be protonated directly by a water molecule at the active site in the ternary complex of the Escherichia coli enzyme with cofactor and substrate. However, in the X-ray structures representing the Michaelis complex of the E. coli enzyme, a water molecule has never been observed in a position that could allow protonation of N5. In fact, the side chain of Met 20 blocks access to N5. Energy minimization reported here revealed that water could be placed in hydrogen bonding distance of N5 with only minor conformational changes. The r.m.s. deviation between the conformation of the M20 loop observed in the crystal structures of the ternary complexes and the conformation adopted after energy minimization was only 0.79 A. We performed molecular dynamics simulations to determine the accessibility by water of the active site of the Michaelis complex of DHFR. Water could access N5 relatively freely after an equilibration time of approximately 300 psec during which the side chain of Met 20 blocked water access. Protonation of N5 did not increase the accessibility by water. Surprisingly the number of near-attack conformations, in which the distance between the pro-R hydrogen of NADPH and C6 of dihydrofolate was less than 3.5 A and the angle between C4 and the pro-R hydrogen of NADPH and C6 of dihydrofolate was greater than 120 degrees, did not increase after protonation. However, when the hydride was transferred from NADPH to C6 of dihydrofolate before protonation, the side chain of Met 20 moved away from N5 after approximately 100 psec thereby providing water access. The average time during which water was found in hydrogen bonding distance to N5 was significantly increased. These results suggest that hydride transfer might occur early to midway through the reaction followed by protonation. Such a mechanism is supported by the very close contact between C4 of NADP+ and C6 of folate observed in the crystal structures of the ternary enzyme complexes, when the M20 loop is in its closed conformation.  相似文献   

13.
Selenophosphate synthetase (SPS) catalyzes the synthesis of selenophosphate, the selenium donor for the biosynthesis of selenocysteine and 2-selenouridine residues in seleno-tRNA. Selenocysteine, known as the 21st amino acid, is then incorporated into proteins during translation to form selenoproteins which serve a variety of cellular processes. SPS activity is dependent on both Mg(2+) and K(+) and uses ATP, selenide, and water to catalyze the formation of AMP, orthophosphate, and selenophosphate. In this reaction, the gamma phosphate of ATP is transferred to the selenide to form selenophosphate, while ADP is hydrolyzed to form orthophosphate and AMP. Most of what is known about the function of SPS has derived from studies investigating Escherichia coli SPS (EcSPS) as a model system. Here we report the crystal structure of the C17S mutant of SPS from E. coli (EcSPS(C17S)) in apo form (without ATP bound). EcSPS(C17S) crystallizes as a homodimer, which was further characterized by analytical ultracentrifugation experiments. The glycine-rich N-terminal region (residues 1 through 47) was found in the open conformation and was mostly ordered in both structures, with a magnesium cofactor bound at the active site of each monomer involving conserved aspartate residues. Mutating these conserved residues (D51, D68, D91, and D227) along with N87, also found at the active site, to alanine completely abolished AMP production in our activity assays, highlighting their essential role for catalysis in EcSPS. Based on the structural and biochemical analysis of EcSPS reported here and using information obtained from similar studies done with SPS orthologs from Aquifex aeolicus and humans, we propose a catalytic mechanism for EcSPS-mediated selenophosphate synthesis.  相似文献   

14.
Biliverdin reductase (BVR) catalyzes the last step in heme degradation by reducing the gamma-methene bridge of the open tetrapyrrole, biliverdin IXalpha, to bilirubin with the concomitant oxidation of a beta-nicotinamide adenine dinucleotide (NADH) or beta-nicotinamide adenine dinucleotide phosphate (NADPH) cofactor. Bilirubin is the major bile pigment in mammals and has antioxidant and anticompliment activity. We have determined X-ray crystal structures of apo rat BVR and its complex with NADH at 1.2 A and 1.5 A resolution, respectively. In agreement with an independent structure determination of the apo-enzyme, BVR consists of an N-terminal dinucleotide-binding domain (Rossmann-fold) and a C-terminal domain that contains a six-stranded beta-sheet that is flanked on one face by several alpha-helices. The C-terminal and N-terminal domains interact extensively, forming the active site cleft at their interface. The cofactor complex structure reported here reveals that the cofactor nicotinamide ring extends into the active site cleft, where it is adjacent to conserved amino acid residues and, consistent with the known stereochemistry of the reaction catalyzed by BVR, the si face of the ring is accessible for hydride transfer. The only titratable side-chain that appears to be suitably positioned to function as a general acid in catalysis is Tyr97. This residue, however, is not essential for catalysis, since the Tyr97Phe mutant protein retains 50% activity. This finding suggests that the dominant role in catalysis may be performed by hydride transfer from the cofactor, a process that may be promoted by proximity of the invariant residues Glu96, Glu123, and Glu126, to the nicotinamide ring.  相似文献   

15.
A gene encoding an sn‐glycerol‐1‐phosphate dehydrogenase (G1PDH) was identified in the hyperthermophilic archaeon Pyrobaculum calidifontis. The gene was overexpressed in Escherichia coli, and its product was purified and characterized. In contrast to conventional G1PDHs, the expressed enzyme showed strong preference for NADH: the reaction rate (Vmax) with NADPH was only 2.4% of that with NADH. The crystal structure of the enzyme was determined at a resolution of 2.45 Å. The asymmetric unit consisted of one homohexamer. Refinement of the structure and HPLC analysis showed the presence of the bound cofactor NADPH in subunits D, E, and F, even though it was not added in the crystallization procedure. The phosphate group at C2’ of the adenine ribose of NADPH is tightly held through the five biased hydrogen bonds with Ser40 and Thr42. In comparison with the known G1PDH structure, the NADPH molecule was observed to be pushed away from the normal coenzyme binding site. Interestingly, the S40A/T42A double mutant enzyme acquired much higher reactivity than the wild‐type enzyme with NADPH, which suggests that the biased interactions around the C2’‐phosphate group make NADPH binding insufficient for catalysis. Our results provide a unique structural basis for coenzyme preference in NAD(P)‐dependent dehydrogenases. Proteins 2016; 84:1786–1796. © 2016 Wiley Periodicals, Inc.  相似文献   

16.
Glutamate dehydrogenase is found in all organisms and catalyses the oxidative deamination of l-glutamate to 2-oxoglutarate. However, only animal GDH utilizes both NAD(H) or NADP(H) with comparable efficacy and exhibits a complex pattern of allosteric inhibition by a wide variety of small molecules. The major allosteric inhibitors are GTP and NADH and the two main allosteric activators are ADP and NAD(+). The structures presented here have refined and modified the previous structural model of allosteric regulation inferred from the original boGDH.NADH.GLU.GTP complex. The boGDH.NAD(+).alpha-KG complex structure clearly demonstrates that the second coenzyme-binding site lies directly under the "pivot helix" of the NAD(+) binding domain. In this complex, phosphates are observed to occupy the inhibitory GTP site and may be responsible for the previously observed structural stabilization by polyanions. The boGDH.NADPH.GLU.GTP complex shows the location of the additional phosphate on the active site coenzyme molecule and the GTP molecule bound to the GTP inhibitory site. As expected, since NADPH does not bind well to the second coenzyme site, no evidence of a bound molecule is observed at the second coenzyme site under the pivot helix. Therefore, these results suggest that the inhibitory GTP site is as previously identified. However, ADP, NAD(+), and NADH all bind under the pivot helix, but a second GTP molecule does not. Kinetic analysis of a hyperinsulinism/hyperammonemia mutant strongly suggests that ATP can inhibit the reaction by binding to the GTP site. Finally, the fact that NADH, NAD(+), and ADP all bind to the same site requires a re-analysis of the previous models for NADH inhibition.  相似文献   

17.
Creatininase is a binuclear zinc enzyme and catalyzes the reversible conversion of creatinine to creatine. It exhibits an open-closed conformational change upon substrate binding, and the differences in the conformations of Tyr121, Trp154, and the loop region containing Trp174 were evident in the enzyme-creatine complex when compared to those in the ligand-free enzyme. We have determined the crystal structure of the enzyme complexed with a 1-methylguanidine. All subunits in the complex existed as the closed form, and the binding mode of creatinine was estimated. Site-directed mutagenesis revealed that the hydrophobic residues that show conformational change upon substrate binding are important for the enzyme activity.We propose a catalytic mechanism of creatininase in which two water molecules have significant roles. The first molecule is a hydroxide ion (Wat1) that is bound as a bridge between the two metal ions and attacks the carbonyl carbon of the substrate. The second molecule is a water molecule (Wat2) that is bound to the carboxyl group of Glu122 and functions as a proton donor in catalysis. The activity of the E122Q mutant was very low and it was only partially restored by the addition of ZnCl2 or MnCl2. In the E122Q mutant, kcat is drastically decreased, indicating that Glu122 is important for catalysis. X-ray crystallographic study and the atomic absorption spectrometry analysis of the E122Q mutant-substrate complex revealed that the drastic decrease of the activity of the E122Q was caused by not only the loss of one Zn ion at the Metal1 site but also a critical function of Glu122, which most likely exists for a proton transfer step through Wat2.  相似文献   

18.
The active site of pig kidney fructose-1,6-bisphosphatase (EC 3.1.3.11) is shared between subunits, Arg-243 of one chain interacting with fructose-1,6-bisphosphate or fructose-2,6-bisphosphate in the active site of an adjacent chain. In this study, we present the X-ray structures of the mutant version of the enzyme with Arg-243 replaced by alanine, crystallized in both T and R allosteric states. Kinetic characteristics of the altered enzyme showed the magnesium binding and inhibition by AMP differed slightly; affinity for the substrate fructose-1,6-bisphosphate was reduced 10-fold and affinity for the inhibitor fructose-2,6-bisphosphate was reduced 1,000-fold (Giroux E, Williams MK, Kantrowitz ER, 1994, J Biol Chem 269:31404-31409). The X-ray structures show no major changes in the organization of the active site compared with wild-type enzyme, and the structures confirm predictions of molecular dynamics simulations involving Lys-269 and Lys-274. Comparison of two independent models of the T form structures have revealed small but significant changes in the conformation of the bound AMP molecules and small reorganization of the active site correlated with the presence of the inhibitor. The differences in kinetic properties of the mutant enzyme indicate the key importance of Arg-243 in the function of fructose-1,6-bisphosphatase. Calculations using the X-ray structures of the Arg-243-->Ala enzyme suggest that the role of Arg-243 in the wild-type enzyme is predominantly electrostatic in nature.  相似文献   

19.
天花粉蛋白与FMP复合物的晶体结构   总被引:6,自引:1,他引:5  
用浸泡法得到了天花粉蛋白(TCS)与FMP复合物的晶体,在SIMENNSX-200B面探测器系统上收集了一套2.0分辨率的X射线衍射数据。用同晶差值傅立叶法解析了复合物的结构,经X—PLOR程序修正得到了TCS—FMP复合物的分子结构并找出了197个水分子,最后的R因子为0.172,键长和键角的RMS偏差分别为0.015和2.922度。TCS—FMP复合物中,FMP与天花粉蛋白分子有较好的结合,其结合位置正处于根据三维结构和突变体信息推测的N一糖苷酶活性口袋之中。它的类嘌呤环夹在Y70和Y111两个侧链环之间,与Y70环近乎平行,其N7和N6分别与TCS分子的G1094羰基氧和I71的N成氢键,N3靠近R163的侧链,其磷酸根则伸向活性口袋的底部,与E189、E160和R163等残基作用。  相似文献   

20.
Complex of a mutant ribonuclease T1 (Y4SW) with a non-cognizable ribonucleotide, 2′AMP, has been determined and refined by X-ray diffraction at 1.7 Å resolution. The 2′AMP molecule locates at a new base-binding site which is remote from the guanine-recognition site, where 2′GMP was found to be bound. The nucleotide adopts the anti conformation of the glycosidic bond and C3′-exo sugar pucker. There exists a single hydrogen bond between the adenine base and the enzyme, and, therefore, the site found is apparently a non-specific binding site. The results indicate that the binding of 2′AMP to the guanine-recognition site is weaker than that to the new binding site.  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