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.
Ramakrishnan B  Boeggeman E  Qasba PK 《Biochemistry》2004,43(39):12513-12522
Beta-1,4-galactosyltransferase (beta4Gal-T1) in the presence of manganese ion transfers galactose from UDP-galactose (UDP-Gal) to N-acetylglucosamine (GlcNAc) that is either free or linked to an oligosaccharide. Crystallographic studies on bovine beta4Gal-T1 have shown that the primary metal binding site is located in the hinge region of a long flexible loop, which upon Mn(2+) and UDP-Gal binding changes from an open to a closed conformation. This conformational change creates an oligosaccharide binding site in the enzyme. Neither UDP nor UDP analogues efficiently induce these conformational changes in the wild-type enzyme, thereby restricting the structural analysis of the acceptor binding site. The binding of Mn(2+) involves an uncommon coordination to the Sdelta atom of Met344; when it is mutated to His, the mutant M344H, in the presence of Mn(2+) and UDP-hexanolamine, readily changes to a closed conformation, facilitating the structural analysis of the enzyme bound with an oligosaccharide acceptor. Although the mutant M344H loses 98% of its Mn(2+)-dependent activity, it exhibits 25% of its activity in the presence of Mg(2+). The crystal structures of M344H-Gal-T1 in complex with either UDP-Gal.Mn(2+) or UDP-Gal.Mg(2+), determined at 2.3 A resolution, show that the mutant enzyme in these complexes is in a closed conformation, and the coordination stereochemistry of Mg(2+) is quite similar to that of Mn(2+). Although either Mn(2+) or Mg(2+), together with UDP-Gal, binds and changes the conformation of the M344H mutant to the closed one, it is the Mg(2+) complex that engages efficiently in catalyses. Thus, this property enabled us to crystallize the M344H mutant for the first time with the acceptor substrate chitobiose in the presence of UDP-hexanolamine and Mn(2+). The crystal structure determined at 2.3 A resolution reveals that the GlcNAc residue at the nonreducing end of chitobiose makes extensive hydrophobic interactions with the highly conserved Tyr286 residue.  相似文献   

2.
Recent structural investigations on the beta-1,4-galactosyltransferase-1 (Gal-T1) and lactose synthase (LS) have revealed that they are akin to an exquisite mechanical device with two well-coordinated flexible loops that are contained within the Gal-T1 catalytic domain. The smaller one has a Trp residue (Trp314) flanked by glycine residues. The larger one comprises amino acid residues 345 to 365. Upon substrate binding, the Trp314 side chain moves to lock the sugar nucleotide in the binding site, while the large loop undergoes a conformational change, masking the sugar nucleotide binding site, and creates (i) the oligosaccharide binding cavity; (ii) a protein-protein interacting site for the enzyme's partner, alpha-lactalbumin (LA); and (iii) a metal ion binding site. Only in conformation II do Gal-T1 and LA form the LS complex, enabling Gal-T1 to choose the new substrate glucose. LA holds and puts Glc right in the acceptor binding site of Gal-T1, which then maximizes the interactions with Glc, thereby making it a preferred acceptor for the LS reaction. The interaction of LA with Gal-T1 in conformation II also stabilizes the sugar-nucleotide-enzyme complex, kinetically enhancing the sugar transfer, even from the less preferred sugar nucleotides. The conformational change that masks the sugar nucleotide binding site can also be induced by the acceptor alone, thus making it possible for the protein to act as a specific lectin.  相似文献   

3.
Structural studies of the streptavidin binding loop.   总被引:7,自引:5,他引:2       下载免费PDF全文
The streptavidin-biotin complex provides the basis for many important biotechnological applications and is an interesting model system for studying high-affinity protein-ligand interactions. We report here crystallographic studies elucidating the conformation of the flexible binding loop of streptavidin (residues 45 to 52) in the unbound and bound forms. The crystal structures of unbound streptavidin have been determined in two monoclinic crystal forms. The binding loop generally adopts an open conformation in the unbound species. In one subunit of one crystal form, the flexible loop adopts the closed conformation and an analysis of packing interactions suggests that protein-protein contacts stabilize the closed loop conformation. In the other crystal form all loops adopt an open conformation. Co-crystallization of streptavidin and biotin resulted in two additional, different crystal forms, with ligand bound in all four binding sites of the first crystal form and biotin bound in only two subunits in a second. The major change associated with binding of biotin is the closure of the surface loop incorporating residues 45 to 52. Residues 49 to 52 display a 3(10) helical conformation in unbound subunits of our structures as opposed to the disordered loops observed in other structure determinations of streptavidin. In addition, the open conformation is stabilized by a beta-sheet hydrogen bond between residues 45 and 52, which cannot occur in the closed conformation. The 3(10) helix is observed in nearly all unbound subunits of both the co-crystallized and ligand-free structures. An analysis of the temperature factors of the binding loop regions suggests that the mobility of the closed loops in the complexed structures is lower than in the open loops of the ligand-free structures. The two biotin bound subunits in the tetramer found in the MONO-b1 crystal form are those that contribute Trp 120 across their respective binding pockets, suggesting a structural link between these binding sites in the tetramer. However, there are no obvious signatures of binding site communication observed upon ligand binding, such as quaternary structure changes or shifts in the region of Trp 120. These studies demonstrate that while crystallographic packing interactions can stabilize both the open and closed forms of the flexible loop, in their absence the loop is open in the unbound state and closed in the presence of biotin. If present in solution, the helical structure in the open loop conformation could moderate the entropic penalty associated with biotin binding by contributing an order-to-disorder component to the loop closure.  相似文献   

4.
Oligosaccharide chains of glycoproteins, glycolipids and glycosaminoglycans are synthesized by glycosyltransferases by the transfer of specific glycosyl moieties from activated sugar-nucleotide donors to specific acceptors. Structural studies on several of these enzymes have shown that one or two flexible loops at the substrate-binding site of the enzymes undergo a marked conformational change from an open to a closed conformation on binding the donor substrate. This conformational change, in which the loop acts as a lid covering the bound donor substrate, creates an acceptor-binding site. After the glycosyl unit is transferred from the donor to the acceptor, the saccharide product is ejected and the loop reverts to its native conformation, thereby releasing the remaining nucleotide moiety. The specificity of the sugar donor is determined by a few residues in the sugar-nucleotide-binding pocket of the enzyme that are conserved among the family members from different species.  相似文献   

5.
The structure determination of apotransketolase and the comparison of its three-dimensional structure with that of the holoenzyme has revealed that no large conformational changes are associated with cofactor binding. Two loops at the active site are flexible in the apoenzyme which enables ThDP to reach its binding site. Binding of the cofactor induces defined conformations for these two loops at the active site. One of these loops is directly involving in binding of the cofactors, Ca2+ and ThDP. This loop acts like a flap which closes off the diphosphate binding site. After binding of the cofactor, residues of this loop form interactions to residues of loop 383-398 from the second subunit. These interactions stabilize the conformation of the two loops from a flexible to a 'closed' conformation.  相似文献   

6.
Shi W  Tanaka KS  Crother TR  Taylor MW  Almo SC  Schramm VL 《Biochemistry》2001,40(36):10800-10809
Adenine phosphoribosyltransferase (APRTase) is a widely distributed enzyme, and its deficiency in humans causes the accumulation of 2,8-dihydroxyadenine. It is the sole catalyst for adenine recycling in most eukaryotes. The most commonly expressed APRTase has subunits of approximately 187 amino acids, but the only crystal structure is from Leishmania donovani, which expresses a long form of the enzyme with 237 residues. Saccharomyces cerevisiae APRTase was selected as a representative of the short APRTases, and the structure of the apo-enzyme and sulfate bound forms were solved to 1.5 and 1.75 A, respectively. Yeast APRTase is a dimeric molecule, and each subunit is composed of a central five-stranded beta-sheet surrounded by five alpha-helices, a structural theme found in all known purine phosphoribosyltransferases. The structures reveal several important features of APRTase function: (i) sulfate ions bound at the 5'-phosphate and pyrophosphate binding sites; (ii) a nonproline cis peptide bond (Glu67-Ser68) at the pyrophosphate binding site in both apo-enzyme and sulfate-bound forms; and (iii) a catalytic loop that is open and ordered in the apo-enzyme but open and disordered in the sulfate-bound form. Alignment of conserved amino acids in short-APRTases from 33 species reveals 13 invariant and 15 highly conserved residues present in hinges, catalytic site loops, and the catalytic pocket. Mutagenesis of conserved residues in the catalytic loop, subunit interface, and phosphoribosylpyrophosphate binding site indicates critical roles for the tip of the catalytic loop (Glu106) and a catalytic site residue Arg69, respectively. Mutation of one loop residue (Tyr103Phe) increases k(cat) by 4-fold, implicating altered dynamics for the catalytic site loop.  相似文献   

7.
N-Acetyllactosamine is the most prevalent disaccharide moiety in the glycans on the surface of mammalian cells and often found as repeat units in the linear and branched polylactosamines, known as i- and I-antigen, respectively. The β1-4-galactosyltransferase-I (β4Gal-T1) enzyme is responsible for the synthesis of the N-acetyllactosamine moiety. To understand its oligosaccharide acceptor specificity, we have previously investigated the binding of tri- and pentasaccharides of N-glycan with a GlcNAc at their nonreducing end and found that the extended sugar moiety in these acceptor substrates binds to the crevice present at the acceptor substrate binding site of the β4Gal-T1 molecule. Here we report seven crystal structures of β4Gal-T1 in complex with an oligosaccharide acceptor with a nonreducing end GlcNAc that has a β1-6-glycosidic link and that are analogous to either N-glycan or i/I-antigen. In the crystal structure of the complex of β4Gal-T1 with I-antigen analog pentasaccharide, the β1-6-branched GlcNAc moiety is bound to the sugar acceptor binding site of the β4Gal-T1 molecule in a way similar to the crystal structures described previously; however, the extended linear tetrasaccharide moiety does not interact with the previously found extended sugar binding site on the β4Gal-T1 molecule. Instead, it interacts with the different hydrophobic surface of the protein molecule formed by the residues Tyr-276, Trp-310, and Phe-356. Results from the present and previous studies suggest that β4Gal-T1 molecule has two different oligosaccharide binding regions for the binding of the extended oligosaccharide moiety of the acceptor substrate.  相似文献   

8.
The active site loop of triosephosphate isomerase (TIM) exhibits a hinged-lid motion, alternating between the two well defined "open" and "closed" conformations. Until now the closed conformation had only been observed in protein complexes with substrate analogues. Here, we present the first rabbit muscle apo TIM structure, refined to 1.5A resolution, in which the active site loop is either in the open or in the closed conformation in different subunits of the enzyme. In the closed conformation described here, the lid loop residues participate in stabilizing hydrogen bonds characteristic of holo TIM structures, whereas chemical interactions observed in the open loop conformation are similar to those found in the apo structures of TIM. In the closed conformation, a number of water molecules are observed at the projected ligand atom positions that are hydrogen bonded to the active site residues. Additives used during crystallization (DMSO and Tris molecules and magnesium atoms) were modeled in the electron density maps. However, no specific binding of these molecules is observed at, or close to, the active site and the lid loop. To further investigate this unusual closed conformation of the apo enzyme, two more rabbit muscle TIM structures, one in the same and another in a different crystal form, were determined. These structures present the open lid conformation only, indicating that the closed conformation cannot be explained by crystal contact effects. To rationalize why the active site loop is closed in the absence of ligand in one of the subunits, extensive comparison with previously solved TIM structures was carried out, supported by the bulk of available experimental information about enzyme kinetics and reaction mechanism of TIM. The observation of both open and closed lid conformations in TIM crystals might be related to a persistent conformational heterogeneity of this protein in solution.  相似文献   

9.
The 1.9 A X-ray structure of a membrane-associated glycosyltransferase involved in peptidoglycan biosynthesis is reported. This enzyme, MurG, contains two alpha/beta open sheet domains separated by a deep cleft. Structural analysis suggests that the C-terminal domain contains the UDP-GlcNAc binding site while the N-terminal domain contains the acceptor binding site and likely membrane association site. Combined with sequence data from other MurG homologs, this structure provides insight into the residues that are important in substrate binding and catalysis. We have also noted that a conserved region found in many UDP-sugar transferases maps to a beta/alpha/beta/alpha supersecondary structural motif in the donor binding region of MurG, an observation that may be helpful in glycosyltransferase structure prediction. The identification of a conserved structural motif involved in donor binding in different UDP-sugar transferases also suggests that it may be possible to identify--and perhaps alter--the residues that help determine donor specificity.  相似文献   

10.
Most lipases contain a lid domain to shield the hydrophobic binding site from the water environment. The lid, mostly in helical form, can undergo a conformational change to expose the active cleft during the interfacial activation. Here we report the crystal structures of Malassezia globosa LIP1 (SMG1) at 1.45 and 2.60 ? resolution in two crystal forms. The structures present SMG1 in its closed form, with a novel lid in loop conformation. SMG1 is one of the few members in the fungal lipase family that has been found to be strictly specific for mono- and diacylglycerol. To date, the mechanism for this substrate specificity remains largely unknown. To investigate the substrate binding properties, we built a model of SMG1 in open conformation. Based on this model, we found that the two bulky hydrophobic residues adjacent to the catalytic site and the N-terminal hinge region of the lid both may act as steric hindrances for triacylglycerols binding. These unique structural features of SMG1 will provide a better understanding on the substrate specificity of mono- and diacylglycerol lipases and a platform for further functional study of this enzyme.  相似文献   

11.
X B Tang  J R Casey 《Biochemistry》1999,38(44):14565-14572
AE1, the chloride/bicarbonate anion exchanger of the erythrocyte plasma membrane, is highly sensitive to inhibition by stilbene disulfonate compounds such as DIDS (4,4'-diisothiocyanostilbene-2, 2'-disulfonate) and DNDS (4,4'-dinitrostilbene-2,2'-disulfonate). Stilbene disulfonates recruit the anion binding site to an outward-facing conformation. We sought to identify the regions of AE1 that undergo conformational changes upon noncovalent binding of DNDS. Since conformational changes induced by stilbene disulfonate binding cause anion transport inhibition, identification of the DNDS binding regions may localize the substrate binding region of the protein. Cysteine residues were introduced into 27 sites in the extracellular loop regions of an otherwise cysteineless form of AE1, called AE1C(-). The ability to label these residues with biotin maleimide [3-(N-maleimidylpropionyl)biocytin] was then measured in the absence and presence of DNDS. DNDS reduced the ability to label residues in the regions around G565, S643-M663, and S731-S742. We interpret these regions either as (i) part of the DNDS binding site or (ii) distal to the binding site but undergoing a conformational change that sequesters the region from accessibility to biotin maleimide. DNDS alters the conformation of residues outside the plane of the bilayer since the S643-M663 region was previously shown to be extramembranous. Upon binding DNDS, AE1 undergoes conformational changes that can be detected in extracellular loops at least 20 residues away from the hydrophobic core of the lipid bilayer. We conclude that the TM7-10 region of AE1 is central to the stilbene disulfonate and substrate binding region of AE1.  相似文献   

12.
The degradation of transitory starch in the chloroplast to provide fuel for the plant during the night requires a suite of enzymes that generate a series of short chain linear glucans. However, glucans of less than four glucose units are no longer substrates for these enzymes, whereas export from the plastid is only possible in the form of either maltose or glucose. In order to make use of maltotriose, which would otherwise accumulate, disproportionating enzyme 1 (DPE1; a 4-α-glucanotransferase) converts two molecules of maltotriose to a molecule of maltopentaose, which can now be acted on by the degradative enzymes, and one molecule of glucose that can be exported. We have determined the structure of the Arabidopsis plastidial DPE1 (AtDPE1), and, through ligand soaking experiments, we have trapped the enzyme in a variety of conformational states. AtDPE1 forms a homodimer with a deep, long, and open-ended active site canyon contained within each subunit. The canyon is divided into donor and acceptor sites with the catalytic residues at their junction; a number of loops around the active site adopt different conformations dependent on the occupancy of these sites. The “gate” is the most dynamic loop and appears to play a role in substrate capture, in particular in the binding of the acceptor molecule. Subtle changes in the configuration of the active site residues may prevent undesirable reactions or abortive hydrolysis of the covalently bound enzyme-substrate intermediate. Together, these observations allow us to delineate the complete AtDPE1 disproportionation cycle in structural terms.  相似文献   

13.
Artocarpin, a tetrameric lectin of molecular mass 65 kDa, is one of the two lectins extracted from the seeds of jackfruit. The structures of the complexes of artocarpin with mannotriose and mannopentose reported here, together with the structures of artocarpin and its complex with Me-alpha-mannose reported earlier, show that the lectin possesses a deep-seated binding site formed by three loops. The binding site can be considered as composed of two subsites; the primary site and the secondary site. Interactions at the primary site composed of two of the loops involve mainly hydrogen bonds, while those at the secondary site comprising the third loop are primarily van der Waals in nature. Mannotriose in its complex with the lectin interacts through all the three mannopyranosyl residues; mannopentose interacts with the protein using at least three of the five mannose residues. The complexes provide a structural explanation for the carbohydrate specificities of artocarpin. A detailed comparison with the sugar complexes of heltuba, the only other mannose-specific jacalin-like lectin with known three-dimensional structure in sugar-bound form, establishes the role of the sugar-binding loop constituting the secondary site, in conferring different specificities at the oligosaccharide level. This loop is four residues longer in artocarpin than in heltuba, providing an instance where variation in loop length is used as a strategy for generating carbohydrate specificity.  相似文献   

14.
Three-dimensional structures are not available for polysaccharide synthases and only minimal information on the molecular basis for catalysis is known. The Pasteurella multocida hyaluronan synthase (PmHAS) catalyzes the polymerization of the alternating beta1,3-N-acetylglucosamine-beta1,4-glucuronic acid sugar chain by the sequential addition of single monosaccharides to the non-reducing terminus. Therefore, PmHAS possesses both GlcNAc-transferase and glucuronic acid (GlcUA)-transferase activities. The recombinant Escherichia coli-derived PmHAS enzyme will elongate exogenously supplied hyaluronan chains in vitro with either a single monosaccharide or a long chain depending on the UDP-sugar availability. Competition studies using pairs of acceptors with distinct termini (where one oligosaccharide is a substrate that may be elongated, whereas the other cannot) were performed here; the lack of competition suggests that PmHAS contains at least two distinct acceptor sites. We hypothesize that the size of the acceptor binding pockets of the enzyme corresponds to the size of the smallest high efficiency substrates; thus we tested the relative activity of a series of authentic hyaluronan oligosaccharides and related structural analogs. The GlcUA-transferase site readily elongates (GlcNAc-GlcUA)(2), whereas the GlcNAc-transferase elongates GlcUA-Glc-NAc-GlcUA. The minimally sized oligosaccharides, elongated with high efficiency, both contain a trisaccharide with two glucuronic acid residues that enabled the identification of a synthetic, artificial acceptor for the synthase. PmHAS behaves as a fusion of two complete glycosyltransferases, each containing a donor site and an acceptor site, in one polypeptide. Overall, this information advances the knowledge of glycosaminoglycan biosynthesis as well as assists the creation of various therapeutic sugars for medical applications in the future.  相似文献   

15.
To elucidate the influence of local motion of the polypeptide chain on the catalytic mechanism of an enzyme, we have measured (15)N relaxation data for Escherichia coli dihydrofolate reductase in three different complexes, representing different stages in the catalytic cycle of the enzyme. NMR relaxation data were analyzed by the model-free approach, corrected for rotational anisotropy, to provide insights into the backbone dynamics. There are significant differences in the backbone dynamics in the different complexes. Complexes in which the cofactor binding site is occluded by the Met20 loop display large amplitude motions on the picosecond/nanosecond time scale for residues in the Met20 loop, the adjacent betaF-betaG loop and for residues 67-69 in the adenosine binding loop. Formation of the closed Met20 loop conformation in the ternary complex with folate and NADP(+), results in attenuation of the motions in the Met20 loop and the betaF-betaG loop but leads to increased flexibility in the adenosine binding loop. New fluctuations on a microsecond/millisecond time scale are observed in the closed E:folate:NADP(+) complex in regions that form hydrogen bonds between the Met20 and the betaF-betaG loops. The data provide insights into the changes in backbone dynamics during the catalytic cycle and point to an important role of the Met20 and betaF-betaG loops in controlling access to the active site. The high flexibility of these loops in the occluded conformation is expected to promote tetrahydrofolate-assisted product release and facilitate binding of the nicotinamide ring to form the Michaelis complex. The backbone fluctuations in the Met20 loop become attenuated once it closes over the active site, thereby stabilizing the nicotinamide ring in a geometry conducive to hydride transfer. Finally, the relaxation data provide evidence for long-range motional coupling between the adenosine binding loop and distant regions of the protein.  相似文献   

16.
Twelve plant lectins from the Papilionoideae subfamily were selected to represent a range of carbohydrate specificities, and their sequences were aligned. Two variability indices were applied to the aligned sequences and the results were analysed using the three-dimensional structures of concanavalin A and the pea lectin. The areas of greatest variability were located in the carbohydrate-binding site region, forming a perimeter around a well-conserved core. These residues are inferred to be specificity determining, in the manner of antibodies, and the most variable position corresponded to Tyr100 in concanavalin A, a known ligand contact residue. In addition to the five peptide loops known to form the binding site from crystallographic studies, a sixth segment with variable residues was located in the binding-site region, and this may contribute to oligosaccharide specificity. In their overall composition, the lectin sites resemble those of the sugar-transport proteins rather than antibodies. The prospects for modelling lectin binding sites by the methods used for antibodies were also assessed.  相似文献   

17.
The crystal structure of the catalytic domain of bovine beta1,4-galactosyltransferase (Gal-T1) co-crystallized with UDP-Gal and MnCl(2) has been solved at 2.8 A resolution. The structure not only identifies galactose, the donor sugar binding site in Gal-T1, but also reveals an oligosaccharide acceptor binding site. The galactose moiety of UDP-Gal is found deep inside the catalytic pocket, interacting with Asp252, Gly292, Gly315, Glu317 and Asp318 residues. Compared to the native crystal structure reported earlier, the present UDP-Gal bound structure exhibits a large conformational change in residues 345-365 and a change in the side-chain orientation of Trp314. Thus, the binding of UDP-Gal induces a conformational change in Gal-T1, which not only creates the acceptor binding pocket for N-acetylglucosamine (GlcNAc) but also establishes the binding site for an extended sugar acceptor. The presence of a binding site that accommodates an extended sugar offers an explanation for the observation that an oligosaccharide with GlcNAc at the non-reducing end serves as a better acceptor than the monosaccharide, GlcNAc. Modeling studies using oligosaccharide acceptors indicate that a pentasaccharide, such as N-glycans with GlcNAc at their non-reducing ends, fits the site best. A sequence comparison of the human Gal-T family members indicates that although the binding site for the GlcNAc residue is highly conserved, the site that binds the extended sugar exhibits large variations. This is an indication that different Gal-T family members prefer different types of glycan acceptors with GlcNAc at their non-reducing ends.  相似文献   

18.
The luminescence of Tb(III) was used to explore the topography of the metal ion sites of Escherichia coli glutamine synthetase and the relationship between these sites and tryptophan residues of the enzyme. By irradiation of tryptophan residues at 295 nm and measurement of the resulting Tb(III) luminescence at 544 nm, a biphasic curve was obtained upon titrating apoenzyme with Tb(III) indicating sequential binding of Tb(III) ions to the two binding sites of glutamine synthetase. The luminescence intensity was greater in the second region of the titration curve which is mostly due to energy transfer from Trp-158 to the second Tb(III) binding site of the enzyme. By use of the F?rster equation for energy transfer from donor Trp to acceptor Tb(III), distances from Trp-57 to Tb(III) at the n1 and n2 sites were calculated, by using a mutant enzyme in which Trp-158 was replaced by Ser, to be 16.4 and 15.7 A, respectively; distances from Trp-158 to Tb(III) at the n1 and n2 sites were calculated, by using a mutant enzyme in which Trp-57 was replaced by Leu, to be 16.8 and 9.5 A, respectively. All the distances are in reasonably good agreement with the crystal structure distances from Salmonella typhimurium glutamine synthetase except the distance from Trp-158 to the second Tb(III) binding site. The discrepancies may result from a slightly different conformation of glutamine synthetase in solution and in the crystal and/or a slightly different conformation for trivalent Ln(III) binding compared to divalent Mn(II) binding.  相似文献   

19.
In order to better understand ligand-induced closure in domain enzymes, open unliganded X-ray structures and closed liganded X-ray structures have been studied in five enzymes: adenylate kinase, aspartate aminotransferase, citrate synthase, liver alcohol dehydrogenase, and the catalytic subunit of cAMP-dependent protein kinase. A sequential model of ligand binding and domain closure was used to test the hypothesis that the ligand actively drives closure from an open conformation. The analysis supports the assumption that each enzyme has a dedicated binding domain to which the ligand binds first and a closing domain. In every case, a small number of residues are identified to interact with the ligand to initiate and drive domain closure. In all cases except adenylate kinase, the backbone of residues located in an interdomain-bending region (hinge site) is identified to interact with the ligand to aid in driving closure. In adenylate kinase, the side-chain of a residue located directly adjacent to a bending region drives closure. It is thought that by binding near a hinge site the ligand is able to get within interaction range of residues when the enzyme is in the open conformation. Interdomain bending regions not involved in inducing closure are involved in control, helping to determine the location of the hinge axis. Similarities have been discovered between aspartate aminotransferase and citrate synthase that only come to light in the context of their dynamical behaviour in response to binding their substrate. Similarity also exists between liver alcohol dehydrogenase and cAMP-dependent protein kinase whereby groups on NAD and ATP, respectively, mimic the backbone of a single amino acid residue in a process where a three residue segment located at the terminus of a beta-sheet, moves to form hydrogen bonds with the mimic that resemble those found in a parallel beta-sheet. This interaction helps to drive domain closure in a process that has analogy to protein folding.  相似文献   

20.
The crystal structure of pyruvate decarboxylase from Kluyveromyces lactis has been determined to 2.26 A resolution. Like other yeast enzymes, Kluyveromyces lactis pyruvate decarboxylase is subject to allosteric substrate activation. Binding of substrate at a regulatory site induces catalytic activity. This process is accompanied by conformational changes and subunit rearrangements. In the nonactivated form of the corresponding enzyme from Saccharomyces cerevisiae, all active sites are solvent accessible due to the high flexibility of loop regions 106-113 and 292-301. The binding of the activator pyruvamide arrests these loops. Consequently, two of four active sites become closed. In Kluyveromyces lactis pyruvate decarboxylase, this half-side closed tetramer is present even without any activator. However, one of the loops (residues 105-113), which are flexible in nonactivated Saccharomyces cerevisiae pyruvate decarboxylase, remains flexible. Even though the tetramer assemblies of both enzyme species are different in the absence of activating agents, their substrate activation kinetics are similar. This implies an equilibrium between the open and the half-side closed state of yeast pyruvate decarboxylase tetramers. The completely open enzyme state is favoured for Saccharomyces cerevisiae pyruvate decarboxylase, whereas the half-side closed form is predominant for Kluyveromyces lactis pyruvate decarboxylase. Consequently, the structuring of the flexible loop region 105-113 seems to be the crucial step during the substrate activation process of Kluyveromyces lactis pyruvate decarboxylase.  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