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.
Chen S  Xu XL  Grant GA 《Biochemistry》2012,51(26):5320-5328
Bacterial l-serine dehydratases differ from mammalian l- and d-serine dehydratases and bacterial d-serine dehydratases by the presence of an iron-sulfur center rather than a pyridoxyl phosphate prosthetic group. They exist in two forms, types 1 and 2, distinguished by their sequence and oligomeric configuration. Both types contain an ASB domain, and the type 1 enzymes also contain an ACT domain in a tandem arrangement with the ASB domain like that in type 1 d-3-phosphoglycerate dehydrogenases (PGDHs). This investigation reveals striking kinetic differences between l-serine dehydratases from Bacillus subtilis (bsLSD, type 1) and Legionella pneumophila (lpLSD, type 2). lpLSD is activated by monovalent cations and inhibited by monovalent anions. bsLSD is strongly activated by cations, particularly potassium, and shows a mixed response to anions. Flouride is a competitive inhibitor for lpLSD but an apparent activator for bsLSD at low concentrations and an inhibitor at high concentrations. The reaction products, pyruvate and ammonia, also act as activators but to different extents for each type. Pyruvate activation is competitive with l-serine, but activation of the enzyme is not compatible with it simply competing for binding at the active site and suggests the presence of a second, allosteric site. Because activation can be eliminated by higher levels of l-serine, it may be that this second site is actually a second serine binding site. This is consistent with type 1 PGDH in which the ASB domain functions as a second site for substrate binding and activation.  相似文献   

2.
A structural database search has revealed that the same fold found in the allosteric substrate binding (ASB) domain of Mycobacterium tuberculosis D-3-phosphoglycerate dehydrogenase (PGDH) is found in l-serine dehydratase from Legionella pneumophila. The M. tuberculosis PGDH ASB domain functions in the control of catalytic activity. Bacterial l-serine dehydratases are 4Fe-4S proteins that convert l-serine to pyruvate and ammonia. Sequence homology reveals two types depending on whether their α and β domains are on the same (Type 2) or separate (Type 1) polypeptides. The α domains contain the catalytic iron-sulfur center while the β domains do not yet have a described function, but the structural homology with PGDH suggests a regulatory role. Type 1 β domains also contain additional sequence homologous to PGDH ACT domains. A continuous assay for l-serine dehydratase is used to demonstrate homotropic cooperativity, a broad pH range, and essential irreversibility. Product inhibition analysis reveals a Uni-Bi ordered mechanism with ammonia dissociating before pyruvate. l-Threonine is a poor substrate and l-cysteine and d-serine are competitive inhibitors with K(i) values that differ by almost 10-fold from those reported for Escherichia colil-serine dehydratase. Mutagenesis identifies the three cysteine residues at the active site that anchor the iron-sulfur complex.  相似文献   

3.
Escherichia coli d-3-phosphoglycerate dehydrogenase (PGDH) is a homotetrameric enzyme whose activity is allosterically regulated by l-serine, the end-product of its metabolic pathway. Previous studies have shown that PGDH displays two modes of cooperative interaction. One is between the l-serine binding sites and the other is between the l-serine binding sites and the active sites. Tryptophan 139 participates in an intersubunit contact near the active site catalytic residues. Site-specific mutagenesis of tryptophan 139 to glycine results in the dissociation of the tetramer to a pair of dimers and in the loss of cooperativity in serine binding and between serine binding and inhibition. The results suggest that the magnitude of inhibition of activity at a particular active site is primarily dependent on serine binding to that subunit but that activity can be modulated in a cooperative manner by interaction with adjacent subunits. The disruption of the nucleotide domain interface in PGDH by mutating Trp-139 suggests the potential for a critical role of this interface in the cooperative allosteric processes in the native tetrameric enzyme.  相似文献   

4.
D-3-Phosphoglycerate dehydrogenases (PGDH) exist with at least three different structural motifs and the enzymes from different species display distinctly different mechanisms. In many species, particularly bacteria, the catalytic activity is regulated allosterically through binding of l-serine to a distinct structural domain, termed the ACT domain. Some species, such as Mycobacterium tuberculosis, contain an additional domain, called the "allosteric substrate binding" or ASB domain, that functions as a co-domain in the regulation of catalytic activity. That is, both substrate and effector function synergistically in the regulation of activity to give the enzyme some interesting properties that may have physiological relevance for the persistent state of tuberculosis. Both enzymes function through a V-type regulatory mechanism and, in the Escherichia coli enzyme, it has been demonstrated that this results from a dead-end complex that decreases the concentration of active species rather than a decrease in the velocity of the active species. This review compares and contrasts what we know about these enzymes and provides additional insight into their mechanism of allosteric regulation.  相似文献   

5.
Pleckstrin homology domains are structurally conserved functional domains that can undergo both protein/protein and protein/lipid interactions. Pleckstrin homology domains can mediate inter- and intra-molecular binding events to regulate enzyme activity. They occur in numerous proteins including many that interact with Ras superfamily members, such as p120 GAP. The pleckstrin homology domain of p120 GAP is located in the NH(2)-terminal, noncatalytic region of p120 GAP. Overexpression of the noncatalytic domains of p120 GAP may modulate Ras signal transduction pathways. Here, we demonstrate that expression of the isolated pleckstrin homology domain of p120 GAP specifically inhibits Ras-mediated signaling and transformation but not normal cellular growth. Furthermore, we show that the pleckstrin homology domain binds the catalytic domain of p120 GAP and interferes with the Ras/GAP interaction. Thus, we suggest that the pleckstrin homology domain of p120 GAP may specifically regulate the interaction of Ras with p120 GAP via competitive intra-molecular binding.  相似文献   

6.
Swift RV  McCammon JA 《Biochemistry》2008,47(13):4102-4111
The addition of a N7-methyl guanosine cap to the 5' end of nascent mRNA is carried out by the mRNA-capping enzyme, a two-domain protein that is a member of the nucleotidyltransferase superfamily. The mRNA-capping enzyme is composed of a catalytic nucleotidyltransferase domain and a noncatalytic oligonucleotide/oligosaccharide binding (OB) domain. Large-scale domain motion triggered by substrate binding mediates catalytically requisite conformational rearrangement of the GTP substrate prior to the chemical step. In this study, we employ targeted molecular dynamics (TMD) on the PBCV-1 capping enzyme to probe the global domain dynamics and internal dynamics of conserved residues during the conformational transformation from the open to the closed state. Analysis of the resulting trajectories along with structural and sequence homology to other members of the superfamily allows us to suggest a conserved mechanism of conformational rearrangements spanning all mRNA-capping enzymes and all ATP-dependent DNA ligases. Our results suggest that the OB domain moves quasi-statically toward the nucleotidyltransferase domain, pivoting about a short linker region. The approach of the OB domain brings a conserved RxDK sequence, an element of conserved motif VI, within proximity of the triphosphate of GTP, destabilizing the unreactive conformation and thereby allowing thermal fluctuations to partition the substrate toward the catalytically competent state.  相似文献   

7.
An analysis of amino acid sequences and 3D structures of chloroplast, mitochondrial, and bacterial F1-ATPases revealed that in their α- and β-chains there are short highly conserved segments linking in pairs the catalytic and noncatalytic sites. The analysis was based on the reported effect of directed mutagenesis of amino acids forming these segments on catalytic properties of the F1-ATPases. It is proposed that one of these segments is responsible for transduction of a conformation signal from the noncatalytic to catalytic site upon ADP-for-ATP substitution at the noncatalytic site. At the catalytic site, this signal changes position of the terminal amino acid residue with respect to the adenine part of the molecule and results in a lower tightness of MgADP binding and its dissociation followed by enzyme activation. Mutagenesis of amino acids comprised by the two other segments was shown to produce an effect on the rate of cooperative catalysis, whereas the rate of single-site catalysis remained unaffected. This suggests that these segments are responsible for the cooperative mode of enzyme functioning.  相似文献   

8.
As shown by X-ray crystallography, horse liver alcohol dehydrogenase undergoes a global conformational change upon binding of NAD+ or NADH, involving a rotation of the catalytic domain relative to the coenzyme binding domain and the closing up of the active site to produce a catalytically efficient enzyme. The conformational change requires a complete coenzyme and is affected by various chemical or mutational substitutions that can increase the catalytic turnover by altering the kinetics of the isomerization and rate of dissociation of coenzymes. The binding of NAD+ is kinetically limited by a unimolecular isomerization (corresponding to the conformational change) that is controlled by deprotonation of the catalytic zinc-water to produce a negatively-charged zinc-hydroxide, which can attract the positively-charged nicotinamide ring. The deprotonation is facilitated by His-51 acting through a hydrogen-bonded network to relay the proton to solvent. Binding of NADH also involves a conformational change, but the rate is very fast. After the enzyme binds NAD+ and closes up, the substrate displaces the hydroxide bound to the catalytic zinc; this exchange may involve a double displacement reaction where the carboxylate group of a glutamate residue first displaces the hydroxide (inverting the tetrahedral coordination of the zinc), and then the exogenous ligand displaces the glutamate. The resulting enzyme-NAD+-alcoholate complex is poised for hydrogen transfer, and small conformational fluctuations may bring the reactants together so that the hydride ion is transferred by quantum mechanical tunneling. In the process, the nicotinamide ring may become puckered, as seen in structures of complexes of the enzyme with NADH. The conformational changes of alcohol dehydrogenase demonstrate the importance of protein dynamics in catalysis.  相似文献   

9.
Escherichia coli 3-phosphoglycerate dehydrogenase (PGDH) catalyzes the first step in serine biosynthesis, and is allosterically inhibited by serine. Structural studies revealed a homotetramer in which the quaternary arrangement of subunits formed an elongated ellipsoid. Each subunit consisted of three domains: nucleotide, substrate and regulatory. In PGDH, extensive interactions are formed between nucleotide binding domains. A second subunit-subunit interaction occurs between regulatory domains creating an extended beta sheet. The serine-binding sites overlap this interface. In these studies, the nucleotide and substrate domains (NSDs) were subcloned to identify changes in both catalytic and physical properties upon removal of a subunit-subunit interface. The NSDs did not vary significantly from PGDH with respect to kinetic parameters with the exception that serine no longer had an effect on catalysis. Temperature dependent dynamic light scattering (DLS) revealed the NSDs aggregated > 5 degrees C before PGDH, indicating decreased stability. DLS and gel filtration studies showed that the truncated enzyme formed a tetramer. This result negated the hypothesis that the removal of the regulatory domain would create an enzyme mimic of the unregulated, closely related dimeric enzymes. Expression of the regulatory domain, to study conformational changes induced by serine binding, yielded a product that by CD spectra contained stable secondary structure. DLS and pulsed field gradient NMR studies of the regulatory domain showed the presence of higher oligomers instead of the predicted dimer. We have concluded that the removal of the regulatory domain is sufficient to eliminate serine inhibition but does not have the expected effect on the quaternary structure.  相似文献   

10.
In crystal structures of bovine MF(1), the side chains of alpha F(357) and beta R(372) are near the adenines of nucleotides bound to noncatalytic sites. To determine if during catalysis these side chains must pass through the different arrangements in which they are present in crystal structures, the catalytic properties of the (alpha F(357)C)(3)(beta R(372)C)(3)gamma subcomplex of the TF(1)-ATPase were characterized before and after cross-linking the introduced cysteines with CuCl(2). The unmodified mutant enzyme hydrolyzes MgATP at 50% the rate exhibited by wild type. Detailed comparison of the catalytic properties of the double mutant enzyme before and after cross-linking with those of the wild-type subcomplex revealed the following. Before cross-linking, the (alpha F(357)C)(3)(beta R(372)C)(3)gamma subcomplex has less tendency than wild type to release inhibitory MgADP entrapped in a catalytic site during turnover when MgATP binds to noncatalytic sites. Following cross-linking, ATPase activity is reduced 5-fold, and inhibitory MgADP entrapped in a catalytic site during turnover does not release under conditions wherein binding of ATP to noncatalytic sites of the wild-type enzyme promotes release of MgADP from the affected catalytic site. When assayed in the presence of lauryldimethylamine oxide, which prevents turnover-dependent entrapment of inhibitory MgADP in a catalytic site, ATPase activity of the cross-linked form is 47% that of the unmodified mutant enzyme. These results suggest that, during catalysis, the side chains of alpha F(357) and beta R(372) do not pass through the extremely different relative positions in which they exist at the three noncatalytic site interfaces in crystal structures.  相似文献   

11.
Ensemble kinetics and single-molecule fluorescence microscopy were used to study conformational transitions associated with enzyme catalysis by dihydrofolate reductase (DHFR). The active site loop of DHFR was labeled with a fluorescence quencher, QSY35, at amino acid position 17, and the fluorescent probe, Alexa555, at amino acid 37, by introducing cysteines at these sites with site-specific mutagenesis. The distance between the probes was such that approximately 50% fluorescence resonance energy transfer (FRET) occurred. The double-labeled enzyme retained essentially full catalytic activity, and stopped-flow studies of both the forward and reverse reactions revealed that the distance between probes increased prior to hydride transfer. A fluctuation in fluorescence intensity of single molecules of DHFR was observed in an equilibrium mixture of substrates but not in their absence. Ensemble rate constants were derived from the distributions of lifetimes observed and attributed to a reversible conformational change. Studies were carried out with both NADPH and NADPD as substrates, with no measurable isotope effect. Similar studies with a G121V mutant DHFR resulted in smaller rate constants. This mutant DHFR has reduced catalytic activity, so that the collective data for the conformational change suggest that the conformational change being observed is associated with catalysis and probably represents a conformational change prior to hydride transfer. If the change in fluorescence is attributed to a change in FRET, the distance change associated with the conformational change is approximately 1-2 A. These results are correlated with other measurements related to conformation coupled catalysis.  相似文献   

12.
D Sondhi  P A Cole 《Biochemistry》1999,38(34):11147-11155
Csk (C-terminal Src kinase) is a protein tyrosine kinase that phosphorylates Src family member C-terminal tails, resulting in downregulation of Src family members. It is composed of three principal domains: an SH3 (Src homology 3) domain, an SH2 (Src homology 2) domain, and a catalytic domain. The impact of the noncatalytic domains on kinase catalysis was investigated. The Csk catalytic domain was expressed in Escherichia coli as a recombinant glutathione S-transferase-fusion protein and demonstrated to have 100-fold reduced catalytic efficiency. Production of the catalytic domain by proteolysis of full-length Csk afforded a similar rate reduction. This suggested that the reduction in catalytic efficiency of the recombinant catalytic domain was intrinsic to the sequence and not an artifact related to faulty expression. This rate reduction was similar for peptide and protein substrates and was due almost entirely to a reduced k(cat) rather than to effects on substrate K(m)s. Viscosity experiments on the catalytic fragment kinase reaction demonstrated that the chemical (phosphoryl transfer) step had a reduced rate. While the Csk SH2 domain had no intermolecular effect on the kinase activity of the Csk catalytic domain, the SH3 domain and SH3-SH2 fragment led to a partial rescue (4-5-fold) of the lost kinase activity. This rescue was not achieved with two other SH3 domains (lymphoid cell kinase, Abelson kinase). The extrapolated K(d) of interaction for the Csk catalytic domain with the Csk SH3 domain was 2.2 microM and that of the Csk catalytic domain with the Csk SH3-SH2 fragment was 8.8 microM. Taken together, these findings suggest that there is likely an intramolecular interaction between the catalytic and SH3 domains in full-length Csk that is important for efficient catalysis. By employing a Csk SH3 specific type II polyproline helix peptide and carrying out site-directed mutagenesis, it was established that the SH3 surface that interacts with the catalytic domain was distinct from the surface that binds type II polyproline helix peptides. This finding suggests a novel mode of protein-protein interaction for an SH3 domain. The implications for Csk substrate selectivity, regulation, and function are discussed.  相似文献   

13.
Xu  Guoqiang  Jin  Xuexia  Guo  Wen  Dou  Wenfang  Zhang  Xiaomei  Xu  Zhenghong 《Annals of microbiology》2015,65(2):929-935
The direct fermentative production of l-serine from renewable biomass using Corynebacterium glutamicum is attracting increasing attention. In this study, wild-type C. glutamicum SYPS-062 produced up to 6.65 ± 0.23 g/L l-serine; to further improve l-serine production, the serA gene was cloned, and the C-terminal domain of 3-phosphoglycerate dehydrogenase (PGDH) from this strain was truncated. When expressed in Escherichia coli, the resultant mutein SerAΔ197 showed a specific PGDH activity of 1.092 ± 0.05 U/mg protein, representing a decrease of 25.87 % from that encoded by serA, and was no longer sensitive to high concentrations of l-serine. When serA Δ591 was overexpressed in C. glutamicum SYPS-062, the activity of PGDH in C. glutamicum pJC1-tac-serA Δ591 increased by 47.72 %, and the resultant strain C. glutamicum pJC1-tac-serA Δ591 could accumulate 7.69 ± 0.22 g/L l-serine. Furthermore, when serA Δ591 was overexpressed in C. glutamicum SYPS-062ΔsdaA, the resultant strain could accumulate 8.84 ± 0.23 g/L l-serine at 102 h, and the yield of l-serine on cells (Y p/x) improved by 60 % when compared with that noted in the control. These results demonstrate that l-serine production in C. glutamicum SYPS-062 could be improved by overexpressing a C-terminal truncation of PGDH in combination with other genetic modifications.  相似文献   

14.
The dual specificity mitogen-activated protein kinase phosphatase MKP3 has been shown to down-regulate mitogenic signaling through dephosphorylation of extracellular signal-regulated kinase (ERK). Camps et al. (Camps, M., Nichols, A., Gillieron, C., Antonsson, B., Muda, M., Chabert, C., Boschert, U., and Arkinstall, S. (1998) Science 280, 1262-1265) had demonstrated that ERK binding to the noncatalytic amino-terminal domain of MKP3 can dramatically activate the phosphatase catalytic domain. The physical basis for this activation has not been established. Here, we provide detailed biochemical evidence that ERK activates MKP3 through the stabilization of the active phosphatase conformation, inducing closure of the catalytic "general acid" loop. In the closed conformation, this loop structure can participate efficiently in general acid/base catalysis, substrate binding, and transition-state stabilization. The pH activity profiles of ERK-activated MKP3 clearly indicated the involvement of general acid catalysis, a hallmark of protein-tyrosine phosphatase catalysis. In contrast, unactivated MKP3 did not display this enzymatic group as critical for the low activity form of the enzyme. Using a combination of Br?nsted analyses, pre-steady-state and steady-state kinetics, we have isolated all catalytic steps in the reaction and have quantified the specific rate enhancement. Through protonation of the leaving group and transition-state stabilization, activated MKP3 catalyzes formation of the phosphoenzyme intermediate approximately 100-fold faster than unactivated enzyme. In addition, ERK-activated MKP3 catalyzes intermediate hydrolysis 5-6-fold more efficiently and binds ligands up to 19-fold more tightly. Consistent with ERK stabilizing the active conformation of MKP3, the chemical chaperone dimethyl sulfoxide was able to mimic this activation. A general protein-tyrosine phosphatase regulatory mechanism involving the flexible general acid loop is discussed.  相似文献   

15.
Studies of cGMP binding to both the native cyclic GMP-stimulated phosphodiesterase and to two unique isolated chymotryptic fragments lacking the catalytic domain suggest that the enzyme contains two noncatalytic cGMP-binding sites/homodimer. In the presence of high concentrations of ammonium sulfate, 2 mol of cGMP are bound/mol of cGMP-stimulated phosphodiesterase homodimer. Under these conditions, linear Scatchard plots of binding are obtained that give an apparent Kd of approximately 2 microM. The inclusion of 3-isobutyl-1-methylxanthine produces a curvilinear plot. In the absence of ammonium sulfate, the dissociation of cGMP from the holoenzyme is rapid, having a t1/2 of less than 10 s, and addition of ammonium sulfate to the incubation greatly decreases this rate of dissociation. The native enzyme is resistant to degradation by chymotrypsin in the absence of cGMP; however, in its presence, chymotrypsin treatment produces several discrete fragments. Similarly, in the presence but not in the absence of cGMP, dicyclohexylcarbodiimide causes an irreversible activation of the enzyme without cross-linking the nucleotide to the phosphodiesterase. Both observations provide evidence that a different conformation in the enzyme results from cGMP binding. Only the conformation formed upon cGMP binding is easily attacked by chymotrypsin or permanently activated by treatment with dicyclohexylcarbodiimide. One major chymotryptic cleavage site exposed by cGMP binding is at tyrosine 553, implying that this region takes part in the conformational change. Limited proteolysis experiments indicate that these noncatalytic binding sites are located within a region of internal sequence homology previously proposed to include the cGMP-binding site(s) and that they retain a high affinity and specificity for cGMP independent of the catalytic domain of the enzyme. The products formed by partial proteolysis can be separated into individual catalytically active and cGMP-binding fractions by anion exchange chromatography. Gel filtration and electrophoresis analysis of the isolated fractions suggest that the cGMP-binding peak has a dimeric structure. Moreover, it can be further resolved by polyethyleneimine high performance liquid chromatography into two peaks (Peaks IIIA and IIIB). Peak IIIA binds 2 mol of cGMP/mol of dimer with an apparent Kd of 0.2 microM. Peak IIIB, however, has greatly reduced cGMP binding. Further digestion of these fragments with cyanogen bromide show that the differences between Peaks IIIA and IIIB are due to one or more additional proteolytic nicks in IIIB that remove a few residues near its C terminus, most probably residues 523-550 or 534-550. This in turn suggests that this region is essential for cGMP-binding activity.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
R T Raines  J R Knowles 《Biochemistry》1987,26(22):7014-7020
Triosephosphate isomerase has been shown to exist in two unliganded forms, one of which binds and isomerizes (R)-glyceraldehyde 3-phosphate and the other of which binds and isomerizes dihydroxyacetone 3-phosphate. The tracer perturbation method of Britton demonstrates the kinetic significance of the interconversion of these two enzyme forms at high substrate concentrations and yields a rate constant of about 10(6) s-1 for the interconversion. Although the molecular nature of the two forms of unliganded enzyme is not defined by these experiments, a shuffling of protons among active site residues, or a protein conformational change, or both, may be involved. This study, coupled with the known rate constants for the substrate-handling steps of triosephosphate isomerase catalysis, completes the kinetic characterization of the catalytic cycle for this enzyme.  相似文献   

17.
When heat-activated F1-ATPase from chloroplasts was repeatedly exposed to Mg2+ and 2-azido-ATP, followed by separation from medium nucleotides and photolysis, a total of two sites per enzyme, both catalytic and noncatalytic, were labeled. In a coupled assay with pyruvate kinase about half the activity was lost when one site per enzyme was modified. However, increased modification resulted in no further loss of activity. In contrast, methanol-sulfite activation of the enzyme showed a loss of most of the catalytic capacity when one site per enzyme was modified. Predominant labeling of either one catalytic or one noncatalytic site caused a loss of most of the activity in either assay. An indication that the enzyme modified at one site retained some catalytic activity was verified by measurement of the [18O]Pi species formed when [gamma-18O]ATP was hydrolyzed by partially derivatized enzyme. With either catalytic or noncatalytic site modification, the distributions of [18O]Pi species formed showed that the modified enzyme had different catalytic characteristics. An interpretation is that with modification by azido nucleotides at either catalytic or noncatalytic sites, capacity for rapid catalysis is largely lost but the remaining sites retain weak modified catalytic properties.  相似文献   

18.
《Journal of molecular biology》2019,431(19):3677-3689
Photoreceptor phosphodiesterase (PDE6) is the central effector enzyme in the visual excitation pathway in rod and cone photoreceptors. Its tight regulation is essential for the speed, sensitivity, recovery, and adaptation of visual signaling. The rod PDE6 holoenzyme (Pαβγ2) is composed of a catalytic heterodimer (Pαβ) that binds two inhibitory γ subunits. Each of the two catalytic subunits (Pα and Pβ) contains a catalytic domain responsible for cGMP hydrolysis and two tandem GAF domains, one of which binds cGMP noncatalytically. Unlike related GAF-containing PDEs where cGMP binding allosterically activates catalysis, the physiological significance of cGMP binding to the GAF domains of PDE6 is unknown. To elucidate the structural determinants of PDE6 allosteric regulators, we biochemically characterized PDE6 complexes in various allosteric states (Pαβ, Pαβ–cGMP, Pαβγ2, and Pαβγ2–cGMP) with a quantitative cross-linking/mass spectrometry approach. We employed a normalization strategy to dissect the cross-linking reactivity of individual residues in order to assess the spatial cross-linking propensity of detected pairs. In addition to identifying cross-linked pairs that undergo conformational changes upon ligand binding, we observed an asymmetric binding of the inhibitory γ-subunit and the noncatalytic cGMP to the GAFa domains of rod PDE6, as well as a stable open conformation of Pαβ catalytic dimer in different allosteric states. These results advance our understanding of the exquisite regulatory control of the lifetime of rod PDE6 activation/deactivation during visual signaling, as well as providing a structural basis for interpreting how mutations in rod PDE6 subunits can lead to retinal diseases.  相似文献   

19.
The serine-histidine-aspartate triad is well known for its covalent, nucleophilic catalysis in a diverse array of enzymatic transformations. Here we show that its nucleophilicity is shielded and its catalytic role is limited to being a specific general base by an open-closed conformational change in the catalysis of (1R,6R)-2-succinyl-6-hydroxy-2,4-cyclohexadiene-1-carboxylate synthase (or MenH), a typical α/β-hydrolase fold enzyme in the vitamin K biosynthetic pathway. This enzyme is found to adopt an open conformation without a functional triad in its ligand-free form and a closed conformation with a fully functional catalytic triad in the presence of its reaction product. The open-to-closed conformational transition involves movement of half of the α-helical cap domain, which causes extensive structural changes in the α/β-domain and forces the side chain of the triad histidine to adopt an energetically disfavored gauche conformation to form the functional triad. NMR analysis shows that the inactive open conformation without a triad prevails in ligand-free solution and is converted to the closed conformation with a properly formed triad by the reaction product. Mutation of the residues crucial to this open-closed transition either greatly decreases or completely eliminates the enzyme activity, supporting an important catalytic role for the structural change. These findings suggest that the open-closed conformational change tightly couples formation of the catalytic triad to substrate binding to enhance the substrate specificities and simultaneously shield the nucleophilicity of the triad, thus allowing it to expand its catalytic power beyond the nucleophilic catalysis.  相似文献   

20.
6-Phosphogluconate dehydrogenase (6PGDH), the third enzyme of the pentose phosphate pathway, catalyzes the oxidative decarboxylation of 6-phosphogluconate, making ribulose 5-phosphate, along with the reduction of NADP+ to NADPH and the release of CO2. Here, we report the first apo-form crystal structure of the pathogenic Klebsiella pneumoniae 6PGDH (Kp6PGDH) and the structures of the highly homologous Escherichia coli K12 6PGDH (Ec6PGDH) complexed with substrate, substrate/NADPH and glucose at high resolution. The binding of NADPH to one subunit of the homodimeric structure triggered a 10° rotation and resulting in a 7 Å movement of the coenzyme-binding domain. The coenzyme was thus trapped in a closed enzyme conformation, in contrast to the open conformation of the neighboring subunit. Comparison of our Ec/Kp6PGDH structures with those of other species illustrated how the domain conformation can be affected upon binding of the coenzyme, which in turn gives rise to concomitant movements of two important NADP+-interacting amino acids, M14 and N102. We propose that the catalysis follows an ordered binding mechanism with alternating conformational changes in the corresponding subunits, involving several related amino acid residues.  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