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.
Saxena A  Hur RS  Luo C  Doctor BP 《Biochemistry》2003,42(51):15292-15299
Acetylcholinesterase isolated from fetal bovine serum (FBS AChE) was previously characterized as a globular tetrameric form. Analysis of purified preparations of FBS AChE by gel permeation chromatography revealed the presence of a stable, catalytically active, monomeric form of this enzyme. The two forms could be distinguished from each other based on their molecular weight, hydrodynamic properties, kinetic properties, thermal stability, and the type of glycans they carry. No differences between the two forms were observed for the binding of classical inhibitors such as edrophonium and propidium or inhibitors that are current or potential drugs for the treatment of Alzheimer's disease such as (-) huperzine A and E2020; tacrine inhibited the monomeric form 2-3-fold more potently than the tetrameric form. Sequencing of peptides obtained from an in-gel tryptic digest of the monomer and tetramer by tandem mass spectrometry indicated that the tetramer consists of 583 amino acid residues corresponding to the mature form of the enzyme, whereas the monomer consists of 543-547 amino acid residues. The subunit molecular weight of the protein component of the monomer (major species) was determined to be 59 414 Da and that of the tetramer as 64 239 Da. The N-terminal of the monomer and the tetramer was Glu, suggesting that the monomer is not a result of truncation at the N-terminal. The only differences detected were at the C-terminus. The tetramer yielded the expected C-terminus, CSDL, whereas the C-terminus of the monomer yielded a mixture of peptides, of which LLSATDTLD was the most abundant. These results suggest that monomeric FBS AChE is trimmed at the C-terminus, and the results are consistent with the involvement of C-terminal amino acids in the assembly of monomers into tetramers.  相似文献   

2.
Sepharose-bound tetrameric, dimeric and monomeric forms of yeast glyceraldehyde-3-phosphate dehydrogenase were prepared, as well as immobilized hybrid species containing (by selective oxidation of an active center cysteine residue with H2O2) one inactivated subunit per tetramer or dimer. The catalytic properties of these enzyme forms were compared in the forward reaction (glyceraldehyde-3-phosphate oxidation) and reverse reaction (1,3-bisphosphoglycerate reductive dephosphorylation) under steady-state conditions. In the reaction of glyceraldehyde-3-phosphate oxidation, immobilized monomeric and tetrameric forms exhibited similar specific activities. The hybrid-modified dimer contributed on half of the total activity of a native dimer. The tetramer containing one modified subunit possessed 75% of the activity of an unmodified tetramer. In the reaction of 1,3-bisphosphoglycerate reductive dephosphorylation, the specific activity of the monomeric enzyme species was nearly twice as high as that of the tetramer, suggesting that only one-half of the active centers of the oligomer were acting simultaneously. Subunit cooperativity in catalysis persisted in an isolated dimeric species. The specific activity of a monomer associated with a peroxide-inactivated monomer in a dimer was equal to that of an isolated monomeric species and twice as high as that of a native immobilized dimer. The specific activity of subunits associated with a peroxide-inactivated subunit in a tetramer did not differ from that of a native immobilized tetramer; this indicates that interdimeric interactions are involved in catalytic subunit cooperativity. A complex was formed between the immobilized glyceraldehyde-3-phosphate dehydrogenase and soluble phosphoglycerate kinase. Three monomers of phosphoglycerate kinase were bound per tetramer of the dehydrogenase and one per dimer. Evidence is presented that if the reductive dephosphorylation of 1,3-bisphosphoglycerate proceeds in the phosphoglycerate kinase - glyceraldehyde-3-phosphate dehydrogenase complex, all active sites of the latter enzyme act independently, i.e. subunit cooperativity is abolished.  相似文献   

3.
Human tryptase-beta (HTbeta) is a serine protease that is isolated as a tetramer of four identical, catalytically active subunits (HTbeta-AT). Tetramer activity is not affected by protein-based physiological inhibitors but instead may be regulated by an autoinactivation process we have called spontaneous inactivation. Unless stabilized by heparin or high salt, the active tetramer converts to an inactive state consisting of an inactive-destabilized tetramer that reversibly dissociates to inactive monomers upon dilution. We refer to this mixture of inactive species as siHTbeta and show in this study that previous reports of monomeric catalytic forms are derived from this mixture. siHTbeta itself did not hydrolyze model substrates but unlike the tetramer did react slowly with the serpin alpha2-antiplasmin (alpha2-AP), suggesting a highly limited catalytic potential. In the presence of heparin (or other highly charged polysaccharides), we demonstrate that siHTbeta formed a well-defined complex with the heparin (siHTbeta-HC) that reacted 70-fold faster with alpha2-AP than siHTbeta and also hydrolyzed model substrates and fibrinogen. Formation of siHTbeta-HC was limited to dilute subunit solutions since high subunit concentrations resulted in the reformation of the active tetramer. By compensating for changes in the strength of heparin binding, siHTbeta-HC could be formed over the pH range of 6.0-8.5. The activity dependence on pH was bell-shaped with highest activity between pH 6.8 and pH 7.5. In contrast, HTbeta-AT activity showed no dependence upon heparin, increased over the pH range of 6.0-8.5, and was much higher than that of siHTbeta-HC especially above pH 6.8. HTbeta-AT incubated with excess heparin of different size (3-15 kDa) was functionally stable at 25 degrees C but lost activity regardless of heparin size at 37 degrees C above pH 6.8. The change in stability, which is likely due to weakened heparin binding, did not result in the formation of a stable catalytic monomer. These results confirm that siHTbeta is for the most part an inactive species and that any active monomer is a consequence of heparin binding to siHTbeta under dilute conditions where unfavorable thermodynamics and/or kinetics restrict formation of active tetramer. Heparin binding under these conditions drives a limited reorganization of the active site to a conformation that is catalytic but not the equivalent of a subunit within the active tetramer.  相似文献   

4.
Experimental conditions favouring the dissociation of tetrameric rabbit muscle D-glyceraldehyde-3-phosphate dehydrogenase into active monomers were elaborated. The urea-induced dissociation of the tetramer was shown to be a stepwise process (in 2 M urea only dimers are formed; an increase in urea concentration up to 3 M causes the splitting of the dimers into monomers). The specific activity of immobilized monomers in the glyceraldehyde-3-phosphate oxidation reaction does not differ from that of the parent immobilized tetrameric form. The tetrameric enzyme molecule binds the coenzyme with a negative cooperativity (the first two NAD+ molecules bind with KD below 0.1 microM; for the third and fourth molecules the dissociation constant was determined to be equal to 5.5 +/- 1.5 microM (50 mM medinal buffer, 10 mM sodium phosphate, pH 8.2). The cooperativity of NAD+ binding is preserved in the immobilized preparation of tetrameric dehydrogenase. The immobilized monomers bind NAD+ with KD of 1.6 +/- 1.0 microM. The experimental results are consistent with the hypothesis according to which the association of catalytically active subunits into a tetramer changes their coenzyme-binding properties in such a way that the first two NAD+ molecules bind more firmly to a tetramer than to a monomer, whereas the third and the fourth NAD+ molecules bind less firmly.  相似文献   

5.
In the presence of specific monovalent cations (K+, Cs+, NH4+), inactive monomers of formyltetrahydrofolate synthetase associate to a catalytically active tetramer. The rate and extent of association of enzyme monomers prepared from C. cylindrosporum are enhanced 3.3-and about 50-fold, respectively, by the substitution of D2O for H2O. Both rate and equilibrium solvent isotope effects are due to a decrease in D2O of the dissociation constant of the monomer-cation complex. Analysis of rate and equilibria data obtained in solvent mixtures of varying deuterium/protium ratios indicates that the isotope effect may be due to the change in bonding of a single monomer proton during the association process. The data are most consistent with a model in which this proton is in a very weak potential in the cation-free monomer and is converted to a "normal" water-like proton in the monomer-cation complex.  相似文献   

6.
The unfolding and dissociation of the tetrameric enzyme fructose-1,6-bisphosphatase from pig kidney by guanidine hydrochloride have been investigated at equilibrium by monitoring enzyme activity, ANS binding, intrinsic (tyrosine) protein fluorescence, exposure of thiol groups, fluorescence of extrinsic probes (AEDANS, MIANS), and size-exclusion chromatography. The unfolding is a multistate process involving as the first intermediate a catalytically inactive tetramer. The evidence that indicates the existence of this intermediate is as follows: (1) the loss of enzymatic activity and the concomitant increase of ANS binding, at low concentrations of Gdn.HCl (midpoint at 0.75 M), are both protein concentration independent, and (2) the enzyme remains in a tetrameric state at 0.9 M Gdn.HCl as shown by size-exclusion chromatography. At slightly higher Gdn.HCl concentrations the inactive tetramer dissociates to a compact dimer which is prone to aggregate. Further evidence for dissociation of tetramers to dimers and of dimers to monomers comes from the concentration dependence of AEDANS-labeled enzyme anisotropy data. Above 2.3 M Gdn.HCl the change of AEDANS anisotropy is concentration independent, indicative of monomer unfolding, which also is detected by a red shift of MIANS-labeled enzyme emission. At Gdn.HCl concentrations higher than 3.0 M, the protein elutes from the size-exclusion column as a single peak, with a retention volume smaller than that of the native protein, corresponding to the completely unfolded monomer. In the presence of its cofactor Mg(2+), the denaturated enzyme could be successfully reconstituted into the active enzyme with a yield of approximately 70-90%. Refolding kinetic data indicate that rapid refolding and reassociation of the monomers into a nativelike tetramer and reactivation of the tetramer are sequential events, the latter involving slow and small conformational rearrangements in the refolded enzyme.  相似文献   

7.
On expression of the cDNA encoding the precursor of chicken mitochondrial aspartate aminotransferase (pmAspAT) in Escherichia coli, the bulk of pmAspAT was found to be associated with the 70-kDa heat-shock protein DnaK which is closely related to mitochondrial 70-kDa heat-shock protein (HSP70). Purification protocols for the DnaK/pmAspAT complex and its individual components were elaborated. The complex dissociated on treatment with MgATP or at pH 5.5. Like the mature enzyme, pmAspAT is a dimer (2 x 47 kDa) and exhibits about a third of its enzyme activity. In the DnaK/pmAspAT complex, one DnaK molecule is bound to each subunit of pmAspAT; this tetramer may further aggregate to an octamer. The complex is catalytically almost as active as free pmAspAT. It could be reconstituted from isolated DnaK and pmAspAT. No complex was formed with mAspAT. Apparently, DnaK binds to the solvent-exposed presequence of folded pmAspAT without significantly changing the structure and functional properties of its mature moiety.  相似文献   

8.
The oligomerization state and mode of binding to DNA of the GIY-YIG endonuclease II (EndoII) from bacteriophage T4 was studied using gel filtration and electrophoretic mobility shift assays with a set of mutants previously found to have altered enzyme activity. At low enzyme/DNA ratios all mutants except one bound to DNA only as tetramers to two DNA substrates. The putatively catalytic E118 residue actually interfered with DNA binding (possibly due to steric hindrance or repulsion between the glutamate side chain and DNA), as shown by the ability of E118A to bind stably also as monomer or dimer to a single substrate. The tetrameric structure of EndoII in the DNA–protein complex is surprising considering the asymmetry of the recognized sequence and the predominantly single-stranded nicking. Combining the results obtained here with those from our previous in vivo studies and the recently obtained crystal structure of EndoII E118A, we suggest a model where EndoII translocates DNA between two adjacent binding sites and either nicks one strand of one or both substrates bound by the tetramer, or nicks both strands of one substrate. Thus, only one or two of the four active sites in the tetramer is catalytically active at any time.  相似文献   

9.
Tetrameric D-glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12) isolated from rabbit skeletal muscle was covalently bound to CNBr-activated Sepharose 4B via a single subunit. Catalytically active immobilized dimer and monomeric forms of the enzyme were prepared after urea-induced dissociation of the tetramer. A study of the coenzyme-binding properties of matrix-bound tetrameric, dimeric and monomeric species has shown that: (1) an immobilized tetramer binds NAD+ with negative cooperativity, the dissociation constants being 0.085 microM for the first two coenzyme molecules and 1.3 microM for the third and the fourth one; (2) coenzyme binding to the dimeric enzyme form also displays negative cooperativity with Kd values of 0.032 microM and 1.1 microM for the first and second sites, respectively; (3) the binding of NAD+ to a monomer can occur with a dissociation constant of 1.6 microM which is close to the Kd value for low-affinity coenzyme binding sites of the tetrameric or dimeric enzyme forms. In the presence of NAD+ an immobilized monomer acquires a stability which is not inferior to that of a holotetramer. The catalytic properties of monomeric and tetrameric enzyme forms were compared and found to be different under certain conditions. Thus, the monomers of rabbit muscle D-glyceraldehyde-3-phosphate dehydrogenase displayed a hyperbolic kinetic saturation curve for NAD+, whereas the tetramers exhibited an intermediary plateau region corresponding to half-saturating concentrations of NAD+. At coenzyme concentrations below half-saturating a monomer is more active than a tetramer. This difference disappears at saturating concentrations of NAD+. Immobilized monomeric and tetrameric forms of D-glyceraldehyde-3-phosphate dehydrogenase from baker's yeast were also used to investigate subunit interactions in catalysis. The rate constant of inactivation due to modification of essential arginine residues in the holoenzyme decreased in the presence of glyceraldehyde 3-phosphate, probably as a result of conformational changes accompanying catalysis. This effect was similar for monomeric and tetrameric enzyme forms at saturating substrate concentrations, but different for the two enzyme species under conditions in which about one-half of the active centers remained unsaturated. Taken together, the results indicate that association of D-glyceraldehyde-3-phosphate dehydrogenase monomers into a tetramer imposes some constraints on the functioning of the active centers.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

10.
The 1918 pandemic flu virus caused one of the most deadly pandemics in human history. To search for unique structural features of the neuraminidase from this virus that might have contributed to its unusual virulence, we expressed this enzyme. The purified enzyme appeared as a monomer, a dimer and a tetramer, with only the tetramer being active and therefore biologically relevant. The monomer and the dimer could not be oligomerized into the tetramer in solution, suggesting that some unique structural features were required for oligomerization and activation. These features could be related to N-glycosylation, because the tetramer displayed different N-glycans than the monomer and the dimer. Furthermore, the tetramer was found to be resistant to trypsin digestion, which may give the virus the capability to invade tissues that are normally not infected by influenza viruses and make the virus more robust for infection.  相似文献   

11.
The effect of aldolase on the concentration-dependent kinetic behaviour of phosphofructokinase was investigated by means of covalently attached fluorescent probe and by using a kinetic approach. The dimeric form of kinase in equilibrium with the active tetramer interacts with the native aldolase with an apparent dissociation constant of 2.5 microM. Within this heterologous enzyme complex the phosphofructokinase is catalytically active probably because the aldolase binding to nascent kinase dimers might protect them against inactivation.  相似文献   

12.
The presence of two distinct molecular structures for tyrosinase in fungi is confirmed. The enzyme from Agaricus bisporus is acidic and comprises two dissimilar subunits which aggregate to form a tetramer. This tetramer constitutes the majority both in the resting and functional states. In Neurospora crassa, tyrosinase is slightly basic and contains only one subunit, similar in size to the larger subunit of the Agaricus enzyme. In the resting state Neurospora tyrosinase is distributed among a number of forms, from the monomer to the tetramer. In this case it was possible to show that a species smaller than the tetramer, probably the monomer, was fully active.  相似文献   

13.
The structure of the recombinant medium chain alcohol dehydrogenase (ADH) from the hyperthermophilic archaeon Aeropyrum pernix has been solved by the multiple anomalous dispersion technique using the signal from the naturally occurring zinc ions. The enzyme is a tetramer with 222 point group symmetry. The ADH monomer is formed from a catalytic and a cofactor-binding domain, with the overall fold similar to previously solved ADH structures. The 1.62 A resolution A.pernix ADH structure is that of the holo form, with the cofactor NADH bound into the cleft between the two domains. The electron density found in the active site has been interpreted to be octanoic acid, which has been shown to be an inhibitor of the enzyme. This inhibitor is positioned with its carbonyl oxygen atom forming the fourth ligand of the catalytic zinc ion. The structural zinc ion of each monomer is present at only partial occupancy and in its absence a disulfide bond is formed. The enhanced thermal stability of the A.pernix ADH is thought to arise primarily from increased ionic and hydrophobic interactions on the subunit interfaces.  相似文献   

14.
A previously unrecognized erythrocyte phosphodiesterase I with activity against thymidine-5'-monophospho-p-nitrophenyl ester is described. The enzyme is present in the soluble fraction of the erythrocyte, and was purified about 500-fold by chromatography using DEAE-cellulose, followed by gel chromatography with Sephadex G-200. Erythrocyte phosphodiesterase I has a molecular weight of about 70 000, when fully active as a monomer. Its pI is 5.4 and the pH optimum is 8.5. The Km value for thymidine-5'-monophospho-p-nitrophenyl ester is rather high, about 4 mmol/l. The enzyme has a barely detectable nucleotide pyrophosphatase activity. It is extremely sensitive to SH-inhibitors such as N-ethyl-maleimide, p-chloromercuribenzoate and disulphides (a reversible 50% inhibition was obtained by cystamine, 0.01 mmol/l). It is a metalloenzyme with loosely bound metal, and is stimulated by Mg2+. This activation by Mg2+ is counteracted by Zn2+. Gel chromatography revealed that the enzyme is a monomer in the presence of Mg2+. When inhibited by Zn2+, it forms polymers that can be reconverted to the monomer by thiols. All of the above properties of the erythrocyte enzyme support the conclusion that it is different from plasma membrane phosphodiesterase I (oligonucleate 5'-nucleotidohydrolase, EC 3.1.4.1).  相似文献   

15.
If acetyl-CoA carboxylase in epididymal fat tissue is subject to control by convalent modification as in the case of the liver enzyme, catalytically different forms of carboxylase should exist, independent of polymerization. By treating epididymal fat tissue in culture with epinephrine, we have demonstrated catalytically less active forms of acetyl-CoA carboxylase. The catalytically less active forms of the enzyme reacted to antibody with the same efficiency as the active form of carboxylase. However, the less active enzyme formed by epinephrine treatment of tissues has a sedimentation constant of 30 to 35 S, whereas that of the enzyme from control tissue is 45 S. Incubation of the less active forms of the carboxylase with 10 mM citrate and up to 10 mg/ml of bovine serum albumin activated the enzyme without any change in the sedimentation constant. Therefore, the less active forms of the carboxylase formed as a result of epinephrine treatment are not due to the depolymerization of polymeric forms (45 S) to the protomeric forms (17 to 20 S), but to the formation of intermediate species of carboxylase which cannot form polymeric enzyme (45 S) in the presence of high concentrations of citrate.  相似文献   

16.
BETA-Galactosidase (EC 3.2.1.23), prepared from strains ML 308 and K12 3300 of Escherichia coli, dissociated into an inactive monomer in the presence of Ag+. When such a monomer preparation is treated with excess of thiol an enzymically active dimer is formed in addition to an active tetramer. It is suggested that Ag+ may be of value in studies on other multimeric proteins as a mild dissociating agent.  相似文献   

17.
Beta-Glucuronidase (EC 3.2.1.31) has been isolated from rat-liver microsomes by a novel chromatographic method employing antibody to rat preputial gland beta-glucuronidase coupled to Sepharose. The purified enzyme, homogeneous by several methods, was purified some 1700-fold. The microsomal beta-glucuronidase has been characterized with respect to catalysis, stability, and molecular weight. The purified enzyme is a tetramer of 290 000 daltons. Comparative studies with lysosomal beta-glucuronidase indicate that while these two enzymes are electrophoretically distinct, they are catalytically and immunologically identical and have indistinguishable molecular dimensions. The results suggest that microsomal and lysosomal beta-glucuronidase are charge isomers.  相似文献   

18.
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   

19.
S-adenosylmethionine (SAM) synthetase (EC 2.5.1.6) catalyzes the synthesis of S-adenosylmethionine using l-methionine and ATP as substrates. SAM synthetase gene (metE) from Bacillus subtilis was cloned and over-expressed, for the first time, in the heterologus host Escherichia coli as an active enzyme. Size-exclusion chromatography (SEC) revealed a molecular weight of ~180 kDa, suggesting that the enzyme is a homotetramer stabilized by non-covalent interactions. SAM synthetase exhibited optimal activity at pH 8.0 and 45 degrees C with the requirement of divalent cation Mg(2+), and stimulated by the monovalent cation K(+). The enzyme followed sequential mechanism with a V(max) of 0.362 micromol/min/mg, and a K(m) of 920 microM and 260 microM for ATP and l-methionine, respectively. The urea-induced unfolding equilibrium of the recombinant enzyme revealed a multistate process, comprising partially unfolded tetramer, structural dimer, structural monomer and completely unfolded monomer, as evidenced by intrinsic and extrinsic fluorescence, circular dichroism (CD) and SEC. Absence of trimer in the SEC implicates that the enzyme is a dimer of dimer. Concordance between results of SEC and enzyme activity in the presence of urea amply establishes that tetramer alone with intersubunit active site(s) exhibits enzyme activity.  相似文献   

20.
The first step of the shikimate pathway for aromatic amino acid biosynthesis is catalyzed by 3-deoxy-d-arabino-heptulosonate 7-phosphate synthase (DAH7PS). Thermotoga maritima DAH7PS (TmaDAH7PS) is tetrameric, with monomer units comprised of a core catalytic (β/α)8 barrel and an N-terminal domain. This enzyme is inhibited strongly by tyrosine and to a lesser extent by the presence of phenylalanine. A truncated mutant of TmaDAH7PS lacking the N-terminal domain was catalytically more active and completely insensitive to tyrosine and phenylalanine, consistent with a role for this domain in allosteric inhibition. The structure of this protein was determined to 2.0 Å. In contrast to the wild-type enzyme, this enzyme is dimeric. Wild-type TmaDAH7PS was co-crystallized with tyrosine, and the structure of this complex was determined to a resolution of 2.35 Å. Tyrosine was found to bind at the interface between two regulatory N-terminal domains, formed from diagonally located monomers of the tetramer, revealing a major reorganization of the regulatory domain with respect to the barrel relative to unliganded enzyme. This significant conformational rearrangement observed in the crystal structures was also clearly evident from small angle X-ray scattering measurements recorded in the presence and absence of tyrosine. The closed conformation adopted by the protein on tyrosine binding impedes substrate entry into the neighboring barrel, revealing an unusual tyrosine-controlled gating mechanism for allosteric control of this enzyme.  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