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.
2.
Most investigations of the allosteric properties of the regulatory enzyme aspartate transcarbamoylase (ATCase) from Escherichia coli are based on the sigmoidal dependence of enzyme activity on substrate concentration and the effects of the inhibitor, CTP, and the activator, ATP, on the saturation curves. Interpretations of these effects in terms of molecular models are complicated by the inability to distinguish between changes in substrate binding and catalytic turnover accompanying the allosteric transition. In an effort to eliminate this ambiguity, the binding of the 3H-labeled bisubstrate analog N-(phosphonacetyl)-L-aspartate (PALA) to aspartate transcarbamoylase in the absence and presence of the allosteric effectors ATP and CTP has been measured directly by equilibrium dialysis at pH 7 in phosphate buffer. PALA binds with marked cooperativity to the holoenzyme with an average dissociation constant of 110 nM. ATP and CTP alter both the average affinity of ATCase for PALA and the degree of cooperativity in the binding process in a manner analogous to their effects on the kinetic properties of the enzyme; the average dissociation constant of PALA decreases to 65 nM in the presence of ATP and increases to 266 nM in the presence of CTP while the Hill coefficient, which is 1.95 in the absence of effectors, becomes 1.35 and 2.27 in the presence of ATP and CTP, respectively. The isolated catalytic subunit of ATCase, which lacks the cooperative kinetic properties of the holoenzyme, exhibits only a very slight degree of cooperativity in binding PALA. The dissociation constant of PALA from the catalytic subunit is 95 nM. Interpretation of these results in terms of a thermodynamic scheme linking PALA binding to the assembly of ATCase from catalytic and regulatory subunits demonstrates that saturation of the enzyme with PALA shifts the equilibrium between holoenzyme and subunits slightly toward dissociation. Ligation of the regulatory subunits by either of the allosteric effectors leads to a change in the effect of PALA on the association-dissociation equilibrium.  相似文献   

3.
4.
In an X-ray diffraction study using the method of multiple isomorphous replacement, the structure of aspartate carbamoyltransferase (EC 2.1.3.2) complexed with the bisubstrate analog N-(phosphonacetyl)-L-aspartate (PALA) has been solved to 2.5 A. Ten rounds of model building and 123 cycles of restrained reciprocal space refinement have resulted in a model containing 94.4% of the theoretical atoms of the protein-inhibitor complex with an R-factor of 0.231. The fit of the model to the density is excellent, except for occasional side-chains and two sections of the regulatory chains that may be disordered. The electron density for the PALA molecule is readily identifiable for both catalytic (c) chains of the asymmetric unit and bonding interactions with several important residues including Ser52, Arg54, Thr55, Ser80, Lys84, Arg105, His134, Arg165, Arg229 and Gln231 are apparent. The carboxylate groups of the PALA molecule are in a nearly cis conformation. Gross quaternary changes between the T and R forms are noted and in agreement with earlier work from this laboratory. Namely, in the new structure the catalytic trimers move apart by 12 A along the 3-fold axis of the enzyme and relocate by 10 degrees relative to each other, adopting a more eclipsed position. The regulatory (r) chains in the new structure reorient about their 2-fold axis by 15 degrees. Large tertiary changes that include domain migration and rearrangement are also present between these two forms. In the R form both domains of the catalytic chain relocate closer to each other in order to bind to the inhibitor. The polar domain seems to bind primarily to the carbamoyl phosphate moiety of PALA, and the equatorial domain binds primarily to the L-aspartate moiety. Other changes in tertiary structure bring the 80s loop (from an adjacent catalytic chain) and the 240s loop into a position to interact with the PALA molecule. Changes have been searched for in all interface regions of the enzyme. While the C1-C4 and C1-R4 regions have been completely altered, most of the other interchain interfaces are similar in the T and R forms. The intrachain interfaces, between domains of the same catalytic chains, have undergone some reorganization as these domains move closer to each other when the inhibitor is bound. This new structure allows a reinterpretation of genetic and chemical modification studies done to date.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

5.
6.
7.
Interaction of a 70-amino acid zinc-binding polypeptide from the regulatory chain of aspartate transcarbamoylase (ATCase) with the catalytic (C) subunit leads to dramatic changes in enzyme activity and affinity for ligand binding at the active sites. The complex between the polypeptide (zinc domain) and wild-type C trimer exhibits hyperbolic kinetics in contrast to the sigmoidal kinetics observed with the intact holoenzyme. Moreover, the Scatchard plot for binding N-(phosphonacetyl)-L-aspartate (PALA) to the complex is linear with a Kd corresponding to that evaluated for the holoenzyme converted to the relaxed (R) state. Additional evidence that the binding of the zinc domain to the C trimer converts it to the R state was attained with a mutant form of ATCase in which Lys 164 in the catalytic chain is replaced by Glu. As shown previously (Newell, J.O. & Schachman, H.K., 1990, Biophys. Chem. 37, 183-196), this mutant holoenzyme, which exists in the R conformation even in the absence of active site ligands, has a 50-fold greater affinity for PALA than the free C subunit. Adding the zinc domain to the C trimer containing the Lys 164-->Glu substitution leads to a 50-fold enhancement in the affinity for the bisubstrate analog yielding a value of Kd equal to that for the holoenzyme. A different mutant ATCase containing the Gln 231 to Ile replacement was shown (Peterson, C.B., Burman, D.L., & Schachman, H.K., 1992, Biochemistry 31, 8508-8515) to be much less active as a holoenzyme than as the free C trimer. For this mutant holoenzyme, the addition of substrates does not cause its conversion to the R state. However, the addition of the zinc domain to the Gln 231-->Ile C trimer leads to a marked increase in enzyme activity, and PALA binding data indicate that the complex resembles the R state of the holoenzyme. This interaction leading to a more active conformation serves as a model of intergenic complementation in which peptide binding to a protein causes a conformational correction at a site remote from the interacting surfaces resulting in activation of the protein. This linkage was also demonstrated by difference spectroscopy using a chromophore covalently bound at the active site, which served as a spectral probe for a local conformational change. The binding of ligands at the active sites was shown also to lead to a strengthening of the interaction between the zinc domain and the C trimer.  相似文献   

8.
Eleven independent lines of Syrian hamster cells were selected by using very low levels of N-(phosphonacetyl)-L-aspartate (PALA), an inhibitor of aspartate transcarbamylase. The protocol employed insured that each resistant cell arose during one of the last divisions before selection was applied. Cells of each mutant line contained an amplification of the structural gene for CAD, a trifunctional protein which includes aspartate transcarbamylase and two other enzymes of UMP biosynthesis. Strikingly, despite the minimal selection employed, the degree of amplification of the CAD gene was 6 to 10 times the normal diploid number in all 11 cases. In situ hybridization indicated that the amplified CAD genes were almost always present at a single chromosomal site in each line. Therefore, one of the two alleles was amplified 11- to 19-fold. The rates at which cells became resistant to PALA, determined by fluctuation analysis, were 100 times less dependent on drug concentration than were the frequencies of resistant cells in steady-state populations. The relatively shallow dependence of this rate upon PALA concentration is consistent with our independent observation that most events gave rise to a similar degree of amplification. In six of six cell lines examined, the levels of CAD mRNA and aspartate transcarbamylase activity were elevated two- to fourfold. These lines were resistant to PALA concentrations 20- to 80-fold higher than the ones used for selection. The organization of amplified DNA was examined by hybridizing Southern blots with cloned DNA fragments containing amplified sequences, previously isolated from two cell lines resistant to high levels of PALA. A contiguous region of DNA approximately 44 kilobases long which included the CAD gene was amplified in five of five single-step mutants examined. Outside this region, these mutants shared amplified sequences with only one of the two highly resistant lines.  相似文献   

9.
Nuclear magnetic resonance has been used to study the binding of [13C]carbamyl-P (90% enriched) to the catalytic subunit of Escherichia coli aspartate transcarbamylase. Upon forming a binary complex, there is a small change in the chemical shift of the carbonyl carbon resonance, 2 Hz upfield at pH 7.0, indicating that the environments of the carbonyl group in the active site and in water are similar. When succinate, an analog of L-aspartate, is added to form a ternary complex, there is a large downfield change in the chemical shift for carbamyl-P, consistent with interaction between the carbonyl group and a proton donor of the enzyme. The change might also be caused by a ring current froma nearby aromatic amino acid residue. From the pH dependence of this downfield change and from the effects of L-aspartate analogs other than succinate, the form of the enzyme involved is proposed to be an isomerized ternary complex, previously observed in temperature jump and proton NMR studies. The downfield change to chemical shift for carbamyl-P bound to the isomerized complex is 17.7 +/- 1.0 Hz. Using this value, the relative ability of other four-carbon dicarboxylic acids to form isomerized ternary complexes with the enzyme and carbamyl-P has been evaluated quantitatively. The 13C peak for the transition state analog N-(phosphonacetyl)-L-aspartate (PALA), 90% enriched specifically at the amide carbonyl group, is shifted 20 Hz downfield of the peak for free PALA upon binding to the catalytic subunit at pH 7.0. In contrast, the peak for [1-13C] phosphonaceatmide shifts upfield by about 6 Hz upon binding. Since PALA induces isomerization of the enzyme and phosphonacetamide does not, these data provide further evidence consistent with protonation of the carbonyl group only upon isomerization. The degrees of protonation is strong acids of the carbonyl groups of PALA, phosphonacetamide and urethan (a model for the labile carbamyl-P) have been determined, as have the chemical shifts for these compounds upon full protonation. From these data it is calculated that the amide carbonyl groups of carbamyl-P and PALA might be protonated to a maximum of about 20% in the isomerized complexes at pH 7.0. The change in conformation of the enzyme-carbamyl-P complex upon binding L-aspartate, previously proposed to aid catalysis by compressing the two substrates together in the active site, may be accompanied by polarization of the C=O bond, making this ordinarily unreactive group a much better electrophile. A keto analog of PALA, 4,5-dicarboxy-2-ketopentyl phosphonate, also binds tightly to the catalytic subunit and induces a very similar conformational change, whereas an alcohol analog, 4,5-dicarboxy-2-hydroxypentyl phosphonate, does not bind tightly, indicating the critical importance of an unhindered carbonyl group with trigonal geometry.  相似文献   

10.
Syrian hamster cell lines selected in multiple steps for resistance to high levels of N-(phosphonacetyl)-L-aspartate (PALA) contain many copies of the gene coding for the pyrimidine pathway enzyme CAD. Approximately 500 kilobases of additional DNA was coamplified with each copy of the CAD gene in several cell lines. To investigate its structure and organization, we cloned ca. 162 kilobases of coamplified DNA from cell line 165-28 and ca. 68 kilobases from cell line B5-4, using a screening method based solely on the greater abundance of amplified sequences in the resistant cells. Individual cloned fragments were then used to probe Southern transfers of genomic DNA from 12 different PALA-resistant mutants and the wild-type parents. A contiguous region of DNA ca. 44 kilobases long which included the CAD gene was amplified in all 12 mutants. However, the fragments cloned from 165-28 which were external to this region were not amplified in any other mutant, and the external fragments cloned from B5-4 were not amplified in two of the mutants. These results suggest that movement or major rearrangement of DNA may have accompanied some of the amplification events. We also found that different fragments were amplified to different degrees within a single mutant cell line. We conclude that the amplified DNA was not comprised of identical, tandemly arranged units. Its structure was much more complex and was different in different mutants. Several restriction fragments containing amplified sequences were found only in the DNA of the mutant cell line from which they were isolated and were not detected in DNA from wild-type cells or from any other mutant cells. These fragments contained novel joints created by rearrangement of the DNA during amplification. The cloned novel fragments hybridized only to normal fragments in every cell line examined, except for the line from which each novel fragment was isolated or the parental population for that line. This result argues that "hot spots" for forming novel joints are rare or nonexistent.  相似文献   

11.
12.
1. The charge state of two derivatives of Sepharose prepared by the CNBr activation method were studied by acid-base titration and by ion-exchange chromatography. Dodecyl-Sepharose exhibited cationic groups (21mumol/ml of settled gel; pKa=9.6) that were tentatively assigned to the coupling isourea group. 2. CPAD-Sepharose [N-(3-carboxypropionyl)aminodecyl-Sepharose] has anionic (carboxyl) groups (pKa=4.5) and cationic groups (pKa=9.6) in roughly equal concentrations (e coupling group. CPAD-Sepharose is slightly negatively charged at pH 7.0 and substantially negatively charged at pH 8.5. 3. The pKa values of dodecyl-Sepharose and CPAD-Sepharose are unaffected by a 100-fold increase in the concentration of KCl. 4. CPAD-Sepharose has considerable affinity for wheat-germ aspartate transcarbamoylase at pH 8.5 when the adsorbent and enzyme are both negatively charged. The interaction involves the C10 chain but is relatively moderate compared with C10 chains associated only with positive charge. 5. Desorption of the enzyme adsorbed to CPAD-Sepharose can be achieved by raising the pH to increase the electrostatic repulsion, or by introducing the detergent sodium deoxycholate. Acetone and butan-1-ol also weaken the adsorption at pH 8.5. 6. High concentrations of sodium acetate or sodium phosphate induced the enzyme to bind more tightly to CPAD-Sepharose. 7. These results are discussed in terms of a 'repulsion-controlled' model or hydrophobic chromatography.  相似文献   

13.
The regulatory enzyme aspartate transcarbamoylase (ATCase), comprising 2 catalytic (C) trimers and 3 regulatory (R) dimers, owes its stability to the manifold interchain interactions among the 12 polypeptide chains. With the availability of a recombinant 70-amino acid zinc-containing polypeptide fragment of the regulatory chain of ATCase, it has become possible to analyze directly the interaction between catalytic and regulatory chains in a complex of simpler structure independent of other interactions such as those between the 2 C trimers, which also contribute to the stability of the holoenzyme. Also, the effect of the interaction between the polypeptide, termed the zinc domain, and the C trimer on the thermal stability and other properties can be measured directly. Differential scanning microcalorimetry experiments demonstrated that the binding of the zinc domain to the C trimer leads to a complex of markedly increased thermal stability. This was shown with a series of mutant forms of the C trimer, which themselves varied greatly in their temperature of denaturation due to single amino acid replacements. With some C trimers, for which tm varied over a range of 30 degrees C due to diverse amino acid substitutions, the elevation of tm resulting from the interaction with the zinc domain was as large as 18 degrees C. The values of tm for a variety of complexes of mutant C trimers and the wild-type zinc domain were similar to those observed when the holoenzymes containing the mutant C trimers were subjected to heat denaturation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Conformational changes of the cAMP-dependent protein kinase (PKA) catalytic (C) subunit are critical for the catalysis of gamma-phosphate transfer from adenosine 5'-triphosphate (ATP) to target proteins. Time-resolved fluorescence anisotropy (TRFA) was used to investigate the respective roles of Mg(2+), ATP, MgATP, and the inhibitor peptide (IP20) in the conformational changes of a 5,6-carboxyfluorescein succinimidyl ester (CF) labeled C subunit ((CF)C). TRFA decays were fit to a biexponential equation incorporating the fast and slow rotational correlation times phi(F) and phi(S). The (CF)C apoenzyme exhibited the rotational correlation times phi(F)=1.8+/-0.3 ns and phi(S)=20.1+/-0.6 ns which were reduced to phi(F)=1.1+/-0.2 ns and phi(S)=13.3+/-0.9 ns in the presence of MgATP. The reduction in rotational correlation times indicated that the (CF)C subunit adopted a more compact shape upon formation of a (CF)C.MgATP binary complex. Neither Mg(2+) (1-3 mM) nor ATP (0.4 mM) alone induced changes in the (CF)C subunit conformation equivalent to those induced by MgATP. The effect of MgATP was removed in the presence of ethylenediaminetetraacetic acid (EDTA). The addition of IP20 and MgATP to form the (CF)C x MgATP x IP20 ternary complex produced rotational correlation times similar to those of the (CF)C x MgATP binary complex. However, IP20 alone did not elicit an equivalent reduction in rotational correlation times. The results indicate that binding of MgATP to the C subunit may induce conformation changes in the C subunit necessary for the proper stereochemical alignment of substrates in the subsequent phosphorylation.  相似文献   

15.
16.
13C isotope effects have been measured for the aspartate transcarbamylase holoenzyme (ATCase) and catalytic subunit catalyzed reactions in the presence of the bisubstrate analog N-(phosphonoacetyl)-L-aspartate (PALA). For holoenzyme-catalyzed reactions in the physiological direction with very low levels of L-aspartate as substrate, or with L-cysteine sulfinate as substrate, or in the reverse direction with carbamyl-L-aspartate and phosphate as substrates, the isotope effect data show a slight dependence on PALA concentration. Under these conditions, PALA first stimulates the rate and then inhibits it at higher concentrations. The observed isotope effect at maximum stimulation by PALA is slightly smaller than in the absence of the analog, but as the PALA concentration is increased to reduce the rate to its original value, the observed isotope effect also increases and approaches the value of the isotope effect determined in the absence of PALA. These data suggest that the kinetic properties of the active enzyme are affected by the number of active sites occupied by PALA, indicating communication between subunits, and a mathematical model is proposed which explains our experimental observations. In contrast to these results with the holoenzyme, isotope effects measured for the reaction catalyzed by the isolated catalytic subunits are not altered in the presence of PALA. Taken together, these data are consistent with the two-state model for the homotropic regulation of ATCase.  相似文献   

17.
The chemical and kinetic mechanisms of the reaction catalyzed by the catalytic trimer of aspartate transcarbamoylase have been examined. The variation of the kinetic parameters with pH indicated that at least four ionizing amino acid residues are involved in substrate binding and catalysis. The pH dependence of K(ia) for carbamoyl phosphate and the K(i) for N-(phosphonoacetyl)-L- aspartate revealed that a protonated residue with a pK value of 9.0 is required for the binding of carbamoyl phosphate. However, the variation with pH of K(i) for succinate, a competitive inhibitor of aspartate, and for cysteine sulfinate, a slow substrate, showed that a single residue with a pK value of 7.3 must be protonated for binding these analogues and, by inference, aspartate. The profile of log V against pH displayed a decrease in reaction rate at low and high pH, suggesting that two groups associated with the Michaelis complex, a deprotonated residue with a pK value of 7.2 and a protonated group with a pK value of 9.5, are involved in catalysis. By contrast, the catalytically productive form of the enzyme-carbamoyl phosphate complex, as illustrated in the bell-shaped pH dependence of log (V/K)(asp), is one in which a residue with a pK value of 7.0 must be protonated while a group with a pK value of 9.1 is deprotonated. This interpretation is supported by the results from the temperature dependence of the V and V/K profiles and from the pH dependence of pK(i) for the aspartate analogues.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Snapshots of the catalytic cycle of the allosteric enzyme aspartate transcarbamoylase have been obtained via X-ray crystallography. The enzyme in the high-activity high-affinity R state contains two catalytic chains in the asymmetric unit that are different. The active site in one chain is empty, while the active site in the other chain contains an analog of the first substrate to bind in the ordered mechanism of the reaction. Small angle X-ray scattering shows that once the enzyme is converted to the R state, by substrate binding, the enzyme remains in the R state until substrates are exhausted. Thus, this structure represents the active form of the enzyme trapped at two different stages in the catalytic cycle, before the substrates bind (or after the products are released), and after the first substrate binds. Opening and closing of the catalytic chain domains explains how the catalytic cycle occurs while the enzyme remains globally in the R-quaternary structure.  相似文献   

19.
1. The steady-state kinetics of the bisubstrate reaction catalysed by aspartate transcarbamoylase purified from wheat (Triticum vulgare)-germ have been studied at 25 degrees C, pH 8.5 AND I 0.10-0.12. Initial-velocity and product-inhibition results are consistent with an ordered sequential mechanism in which carbamoyl phosphate is the first substrate to bind, followed by L-aspartate, and carbamoyl aspartate is the first product to leave, followed by Pi. The order of substrate addition is supported by dead-end inhibition studies using pyrophosphate and maleate as inhibitory analogues of the substrates. Product inhibition permitted a minimum value for the dissociation constant of L-aspartate from the ternary complex to be estimated. This minimum is of the same order as the dissociation constant (Ki) of succinate. 2. A range of dicarboxy analogues of L-aspartate were tested as possible inhibitors of the enzyme. These studies suggested that L-aspartate is bound with its carboxy groups in the eclipsed configuration, and that the stereochemical constraints around the binding site are very similar to those reported for the catalytic subunit of the enzyme from Escherichia coli [Davies, Vanaman & Stark (1970) J. Biol. Chem. 245, 1175-1179].  相似文献   

20.
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