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.
P J Day  W V Shaw  M R Gibbs  A G Leslie 《Biochemistry》1992,31(17):4198-4205
The possible involvement of arginyl and lysyl side chains of chloramphenicol acetyltransferase (CAT) in binding coenzyme A (CoA) was studied by means of chemical modification, site-directed mutagenesis, variation in ionic strength, use of competitive inhibitors or substrate analogues, and X-ray crystallography. Unlike a number of enzymes, including citrate synthase, CAT does not employ specific ion pairs with the phosphoanionic centers of CoA to bind the acetyl donor, and arginyl residues play no role in recognition of the coenzyme. Although phenylglyoxal inactivates CAT reversibly, it does so by the formation of an unstable adduct with a thiol group, that of Cys-31 in the chloramphenicol binding site. The inhibitory effect of increasing ionic strength on kcat/Km(acetyl-CoA) can be explained by long-range electrostatic interactions between CoA and the epsilon-amino groups of Lys-54 and Lys-177, both of which are solvent-accessible. The epsilon-amino group of Lys-54 contributes 1.3 kcal.mol-1 to the binding of acetyl-CoA via interactions with both the 3'- and 5'-phosphoanions of CoA. Lys-177 contributes only 0.4 kcal.mol-1 to the productive binding of acetyl-CoA, mediated by long-range (approximately 14 A) interactions with the 5'-alpha- and -beta-phosphoanions of CoA. The combined energetic contribution of Lys-54 and Lys-177 to acetyl-CoA binding (1.7 kcal.mol-1) is less than that previously demonstrated (2.4 kcal.mol-1) for a simple hydrophobic interaction between Tyr-178 and the adenine ring of CoA (Day & Shaw, 1992).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Kim AR  Rylett RJ  Shilton BH 《Biochemistry》2006,45(49):14621-14631
Choline acetyltransferase (ChAT) catalyzes the synthesis of the neurotransmitter acetylcholine from choline and acetyl-CoA, and its presence is a defining feature of cholinergic neurons. We report the structure of human ChAT to a resolution of 2.2 A along with structures for binary complexes of ChAT with choline, CoA, and a nonhydrolyzable acetyl-CoA analogue, S-(2-oxopropyl)-CoA. The ChAT-choline complex shows which features of choline are important for binding and explains how modifications of the choline trimethylammonium group can be tolerated by the enzyme. A detailed model of the ternary Michaelis complex fully supports the direct transfer of the acetyl group from acetyl-CoA to choline through a mechanism similar to that seen in the serine hydrolases for the formation of an acyl-enzyme intermediate. Domain movements accompany CoA binding, and a surface loop, which is disordered in the unliganded enzyme, becomes localized and binds directly to the phosphates of CoA, stabilizing the complex. Interactions between this surface loop and CoA may function to lower the KM for CoA and could be important for phosphorylation-dependent regulation of ChAT activity.  相似文献   

3.
J Ellis  C R Bagshaw  W V Shaw 《Biochemistry》1991,30(44):10806-10813
Chloramphenicol acetyltransferase (CAT) catalyzes the acetyl-CoA-dependent acetylation of chloramphenicol by a ternary complex mechanism with a rapid equilibrium and essentially random order of addition of substrates. Such a kinetic mechanism for a two-substrate reaction provides an opportunity to compare the affinity of enzyme for each substrate in the binary complexes (1/Kd) with corresponding values (1/Km) for affinities in the ternary complex where any effect of the other substrate should be manifest. The pursuit of such information for CAT involved the use of four independent methods to determine the dissociation constant (Kd) for chloramphenicol in the binary complex, techniques which included stopped-flow measurements of on and off rates, and a novel fluorometric titration method. The binary complex dissociation constant (Kd) for acetyl-CoA was measured by fluorescence enhancement and steady-state kinetic analysis. The ternary complex dissociation constant (Km) for each substrate (in the presence of the other) was determined by kinetic and fluorometric methods, using CoA or ethyl-CoA to form nonproductive ternary complexes. The results demonstrate an unequivocal decrease in affinity of CAT for each of its substrates on progression from the binary to the ternary complex, a phenomenon most economically described as negative cooperativity. The binary complex dissociation constants (Kd) for chloramphenicol and acetyl-CoA are 4 microM and 30 microM whereas the corresponding dissociation constants in the ternary complex (Km) are 12 microM and 90 microM, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Analogues of coenzyme A (CoA) and of CoA thioesters have been prepared in which the amide bond nearest the thiol group has been modified. An analogue of acetyl-CoA in which this amide bond is replaced with an ester linkage was a good substrate for the enzymes carnitine acetyltransferase, chloramphenicol acetyltransferase, and citrate synthase, with K(m) values 2- to 8-fold higher than those of acetyl-CoA and V(max) values from 14 to >80% those of the natural substrate. An analogue in which an extra methylene group was inserted between the amide bond and the thiol group showed less than 4-fold diminished binding to the three enzymes but exhibited less than 1% activity relative to acetyl-CoA with carnitine acetyltransferase and no measurable activity with the other two enzymes. Analogues of several CoA thioesters in which the amide bond was replaced with a hemithioacetal linkage exhibited no measurable activity with the appropriate enzymes. The results indicate that some aspects of the amide bond and proper distance between this amide and the thiol/thioester moiety are critical for activity of CoA ester-utilizing enzymes.  相似文献   

5.
Cody V  Galitsky N  Rak D  Luft JR  Pangborn W  Queener SF 《Biochemistry》1999,38(14):4303-4312
Structural data from two independent crystal forms (P212121 and P21) of the folate (FA) binary complex and from the ternary complex with the oxidized coenzyme, NADP+, and recombinant Pneumocystis carinii dihydrofolate reductase (pcDHFR) refined to an average of 2.15 A resolution, show the first evidence of ligand-induced conformational changes in the structure of pcDHFR. These data are also compared with the crystal structure of the ternary complex of methotrexate (MTX) with NADPH and pcDHFR in the monoclinic lattice with data to 2.5 A resolution. Comparison of the data for the FA binary complex of pcDHFR with those for the ternary structures reveals significant differences, with a >7 A movement of the loop region near residue 23 that results in a new "flap-open" position for the binary complex, and a "closed" position in the ternary complexes, similar to that reported for Escherichia coli (ec) DHFR complexes. In the orthorhombic lattice for the binary FA pcDHFR complex, there is also an unwinding of a short helical region near residue 47 that places hydrophobic residues Phe-46 and Phe-49 toward the outer surface, a conformation that is stabilized by intermolecular packing contacts. The pyrophosphate moiety of NADP+ in the ternary folate pcDHFR complexes shows significant differences in conformation compared with that observed in the MTX-NADPH-pcDHFR ternary complex. Additionally, comparison of the conformations among these four pcDHFR structures reveals evidence for subdomain movement that correlates with cofactor binding states. The larger binding site access in the new "flap-open" loop 23 conformation of the binary FA complex is consistent with the rapid release of cofactor from the product complex during catalysis as well as the more rapid release of substrate product from the binary complex as a result of the weaker contacts of the closed loop 23 conformation, compared to ecDHFR.  相似文献   

6.
The three-dimensional structures of D-Phe-Pro-Arg-chloromethyl ketone-inhibited thrombin in complex with Tyr-63-sulfated hirudin (ternary complex) and of thrombin in complex with the bifunctional inhibitor D-Phe-Pro-Arg-Pro-(Gly)4-hirudin (CGP 50,856, binary complex) have been determined by X-ray crystallography in crystal forms different from those described by Skrzypczak-Jankun et al. (Skrzypczak-Jankun, E., Carperos, V.E., Ravichandran, K.G., & Tulinsky, A., 1991, J. Mol. Biol. 221, 1379-1393). In both complexes, the interactions of the C-terminal hirudin segments of the inhibitors binding to the fibrinogen-binding exosite of thrombin are clearly established, including residues 60-64, which are disordered in the earlier crystal form. The interactions of the sulfate group of Tyr-63 in the ternary complex structure explain why natural sulfated hirudin binds with a 10-fold lower K(i) than the desulfated recombinant material. In this new crystal form, the autolysis loop of thrombin (residues 146-150), which is disordered in the earlier crystal form, is ordered due to crystal contacts. Interactions between the C-terminal fragment of hirudin and thrombin are not influenced by crystal contacts in this new crystal form, in contrast to the earlier form. In the bifunctional inhibitor-thrombin complex, the peptide bond between Arg-Pro (P1-P1') seems to be cleaved.  相似文献   

7.
J Ellis  I A Murray  W V Shaw 《Biochemistry》1991,30(44):10799-10805
Replacement by tyrosine or phenylalanine was used to assign the additive contributions of each of the three tryptophan residues of chloramphenicol acetyltransferase (CAT) to its intrinsic fluorescence on excitation at 295 nm. During the assessment of the fluorescence responses of the wild-type enzyme to the binding of ligands, it was found that the overlapping absorption spectra of chloramphenicol and tryptophan, with an attendant inner filter effect, required the use of a displacement technique involving an alternative substrate (the p-cyano analogue of chloramphenicol) without significant absorption at 295 nm. By the use of two-Trp, one-Trp, and Trp-less variants, in combination with this displacement technique, it was possible to demonstrate that Trp-86 and Trp-152 are involved in the fluorescence quenching associated with the binding of chloramphenicol, most likely via nonradiative energy transfer from these residues to the bound substrate. Trp-152 is mainly responsible for the fluorescence enhancement accompanying the binding of acetyl-CoA (and CoA) through proximity effects and solvent exclusion on substrate association.  相似文献   

8.
Recoverin, a member of the neuronal calcium sensor branch of the EF-hand superfamily, serves as a calcium sensor that regulates rhodopsin kinase (RK) activity in retinal rod cells. We report here the NMR structure of Ca(2+)-bound recoverin bound to a functional N-terminal fragment of rhodopsin kinase (residues 1-25, called RK25). The overall main-chain structure of recoverin in the complex is similar to structures of Ca(2+)-bound recoverin in the absence of target (<1.8A root-mean-square deviation). The first eight residues of recoverin at the N terminus are solvent-exposed, enabling the N-terminal myristoyl group to interact with target membranes, and Ca(2+) is bound at the second and third EF-hands of the protein. RK25 in the complex forms an amphipathic helix (residues 4-16). The hydrophobic face of the RK25 helix (Val-9, Val-10, Ala-11, Ala-14, and Phe-15) interacts with an exposed hydrophobic groove on the surface of recoverin lined by side-chain atoms of Trp-31, Phe-35, Phe-49, Ile-52, Tyr-53, Phe-56, Phe-57, Tyr-86, and Leu-90. Residues of recoverin that contact RK25 are highly conserved, suggesting a similar target binding site structure in all neuronal calcium sensor proteins. Site-specific mutagenesis and deletion analysis confirm that the hydrophobic residues at the interface are necessary and sufficient for binding. The recoverin-RK25 complex exhibits Ca(2+)-induced binding to rhodopsin immobilized on concanavalin-A resin. We propose that Ca(2+)-bound recoverin is bound between rhodopsin and RK in a ternary complex on rod outer segment disk membranes, thereby blocking RK interaction with rhodopsin at high Ca(2+).  相似文献   

9.
P Brandt  C Woodward 《Biochemistry》1987,26(11):3156-3167
Hydrogen exchange rates of six beta-sheet peptide amide protons in bovine pancreatic trypsin inhibitor (BPTI) have been measured in free BPTI and in the complexes trypsinogen-BPTI, trypsinogen-Ile-Val-BPTI, bovine trypsin-BPTI, and porcine trypsin-BPTI. Exchange rates in the complexes are slower for Ile-18, Arg-20, Gln-31, Phe-33, Tyr-35, and Phe-45 NH, but the magnitude of the effect is highly variable. The ratio of the exchange rate constant in free BPTI to the exchange rate constant in the complex, k/kcpIx, ranges from 3 to much greater than 10(3). Gln-31, Phe-45, and Phe-33 NH exchange rate constants are the same in each of the complexes. For Ile-18 and Tyr-35, k/kcpIx is much greater than 10(3) for the trypsin complexes but is in the range 14-43 for the trypsinogen complexes. Only the Arg-20 NH exchange rate shows significant differences between trypsinogen-BPTI and trypsinogen-Ile-Val-BPTI and between porcine and bovine trypsin-BPTI.  相似文献   

10.
The dissociation equilibrium constant for heparin binding to antithrombin III (ATIII) is a measure of the cofactor's binding to and activation of the proteinase inhibitor, and its salt dependence indicates that ionic and non-ionic interactions contribute approximately 40 and approximately 60% of the binding free energy, respectively. We now report that phenylalanines 121 and 122 (Phe-121 and Phe-122) together contribute 43% of the total binding free energy and 77% of the energy of non-ionic binding interactions. The large contribution of these hydrophobic residues to the binding energy is mediated not by direct interactions with heparin, but indirectly, through contacts between their phenyl rings and the non-polar stems of positively charged heparin binding residues, whose terminal amino and guanidinium groups are thereby organized to form extensive and specific ionic and non-ionic contacts with the pentasaccharide. Investigation of the kinetics of heparin binding demonstrated that Phe-122 is critical for promoting a normal rate of conformational change and stabilizing AT*H, the high affinity-activated binary complex. Kinetic and structural considerations suggest that Phe-122 and Lys-114 act cooperatively through non-ionic interactions to promote P-helix formation and ATIII binding to the pentasaccharide. In summary, although hydrophobic residues Phe-122 and Phe-121 make minimal contact with the pentasaccharide, they play a critical role in heparin binding and activation of antithrombin by coordinating the P-helix-mediated conformational change and organizing an extensive network of ionic and non-ionic interactions between positively charged heparin binding site residues and the cofactor.  相似文献   

11.
The imidazole of His-195 plays an essential role in the proposed general base mechanism of chloramphenicol acetyltransferase (CAT). The structure of the binary complex of CATIII and chloramphenicol suggests that two unusual interactions might determine the conformation of the side chain of His-195: (i) an intraresidue hydrogen bond between its main chain carbonyl and the protonated N delta 1 of the imidazole ring and (ii) face-to-face van der Waals contact between the His-195 imidazole group and the aromatic side chain of Tyr-25. Tyr-25 also makes a hydrogen bond, via its phenolic hydroxyl, to the carbonyl oxygen of the substrate chloramphenicol. Replacement of Tyr-25 of CATIII by phenylalanine results in a modest increase in the Km for chloramphenicol (from 11.6 to 14.6 microM) and a 2-fold fall in kcat (599 to 258 s-1), indicative of a free energy contribution to transition state binding of 0.6 kcal mol-1 for the hydrogen bond between Tyr-25 and chloramphenicol. In contrast, substitution of Tyr-25 by alanine yields an enzyme that is dramatically impaired in its ability to bind chloramphenicol (Km = 173 microM). As kcat for Ala-25 CAT is also reduced (130 s-1), the loss of the aryl group results in a 69-fold decrease in kcat/Km, corresponding to a free energy contribution to binding and catalysis of 2.5 kcal mol-1. In addition to the loss of the hydrogen bond between Tyr-25 and chloramphenicol, the loss of substrate affinity in Ala-25 CAT may be a direct consequence of reduced hydrophobicity of the chloramphenicol-binding site and/or the loss of critical constraints on the precise conformation of the catalytic imidazole. However, as with wild type CAT, inactivation of Ala-25 CAT by the affinity reagent 3-(bromoacetyl) chloramphenicol is accompanied by modification solely at N epsilon 2 of His-195. Hence, the results demonstrate that tautomeric stabilization of the imidazole ring persists in the absence of van der Waals interactions with the side chain of Tyr-25, probably as a consequence of hydrogen bonding between the protonated N delta 1 and the carbonyl oxygen of His-195.  相似文献   

12.
The mechanism of the enzymic reaction responsible for chloramphenicol resistance in bacteria was examined by steady-state kinetic methods. The forward reaction catalysed by chloramphenicol acetyltransferase leads to inactivation of the antibiotic. Use of alternative acyl donors and acceptors, as well as the natural substrates, has yielded data that favour the view that the reaction proceeds to the formation of a ternary complex by a rapid-equilibrium mechanism wherein the addition of substrates may be random but a preference for acetyl-CoA as the leading substrate can be detected. Chloramphenicol and acetyl-CoA bind independently, but the correlation between directly determined and kinetically derived dissociation constants is imperfect because of an unreliable slope term in the rate equation. The reverse reaction, yielding acetyl-CoA and chloramphenicol, was studied in a coupled assay involving citrate synthase and malate dehydrogenase, and is best described by a rapid-equilibrium mechanism with random addition of substrates. The directly determined dissociation constant for CoA is in agreement with that derived from kinetic measurements under the assumption of an independent-sites model.  相似文献   

13.
1H-NMR spectroscopy is employed to study the interaction between rabbit skeletal muscle troponin (C (TnC) and wasp venom tetradecapeptide mastoparan. We monitored the spectral change of the following species of TnC as a function of mastoparan concentration: apoTnC, Ca(2+)-saturated TnC (Ca4TnC) and Ca(2+)-half loaded TnC (Ca2TnC). When apo-TnC is titrated with mastoparan, line-broadening is observed for the ring-current shifted resonance of Phe-23, Ile-34, Val-62 and Phe-72 and the downfield-shifted CH alpha-resonances of Asp-33, Thr-69 and Asp-71; these residues are located in the N-domain. When Ca4TnC is titrated with mastoparan, chemical shift change is observed for the ring-current shifted resonances of Phe-99, Ile-110 and Phe-148 and the downfield-shifted CH alpha-resonances of Asn-105, Ala-106, Ile-110 and Ile-146 and aromatic resonance of Tyr-109 and His-125; these residues are located in the C-domain. The resonance of Phe-23, Asp-33, Asp-71, Phe-72, Phe-99, Tyr-109, Ile-146, His-125 and Phe-148 in both N- and C-domains changes when Ca2TnC is titrated with mastoparan. These results suggest that mastoparan binds to the N-domain of apo-TnC, the C-domain of Ca4TnC and the N- and C-domains of Ca2TnC; the hydrophobic cluster in each domain is involved in binding. As mastoparan binds to TnC, the above resonances shift to their normal chemical shift positions. The stability of the cluster and the beta-sheet is reduced by mastoparan-binding. These results suggest that the conformation of the hydrophobic cluster and the neighboring beta-sheet change to a loose form. The stability of the N-domain of Ca2TnC and Ca4TnC increases when these species bind 1 mol of mastoparan at the C-domain. These results suggest a mastoparan-induced interaction between the N- and C-domains of TnC.  相似文献   

14.
The carbon-13 NMR spectrum of oxaloacetate bound in the active site of citrate synthase has been obtained at 90.56 MHz. In the binary complex with enzyme, the positions of the resonances of oxaloacetate are shifted relative to those of the free ligand as follows: C-1 (carboxylate), -2.5 ppm; C-2 (carbonyl), +4.3 ppm; C-3 (methylene), -0.6 ppm; C-4 (carboxylate), +1.3 ppm. The change observed in the carbonyl chemical shift is successively increased in ternary complexes with the product [coenzyme A (CoA)], a substrate analogue (S-acetonyl-CoA), and an acetyl-CoA enolate analogue (carboxymethyl-CoA), reaching a value of +6.8 ppm from the free carbonyl resonance. Binary complexes are in intermediate to fast exchange on the NMR time scale with free oxaloacetate; ternary complexes are in slow exchange. Line widths of the methylene resonance in the ternary complexes suggest complete immobilization of oxaloacetate in the active site. Analysis of line widths in the binary complex suggests the existence of a dynamic equilibrium between two or more forms of bound oxaloacetate, primarily involving C-4. The changes in chemical shifts of the carbonyl carbon indicate strong polarization of the carbonyl bond or protonation of the carbonyl oxygen. Some of this carbonyl polarization occurs even in the binary complex. Development of positive charge on the carbonyl carbon enhances reactivity toward condensation with the carbanion/enolate of acetyl-CoA in the mechanism which has been postulated for this enzyme. The very large change in the chemical shift of the reacting carbonyl in the presence of an analogue of the enolate of acetyl-CoA supports this interpretation.  相似文献   

15.
Grayanotoxin (GTX) exerts selective effects on voltage-dependent sodium channels by eliminating fast sodium inactivation and causing a hyperpolarizing shift in voltage dependence of channel activation. In this study, we adopted a newly developed protocol that provides independent estimates of the binding and unbinding rate constants of GTX (k(on) and k(off)) to GTX sites on the sodium channel protein, important in the molecular analysis of channel modification. Novel GTX sites were determined in D2S6 (Asn-784) and D3S6 (Ser-1276) by means of site-directed mutagenesis; the results suggested that the GTX receptor consists of the S6 transmembrane segments of four homologous domains facing the ion-conducting pore. We systematically introduced at two sites in D4S6 (Na(v)1.4-Phe-1579 and Na(v)1.4-Tyr-1586) amino acid substituents with residues containing hydrophobic, aromatic, charged, or polar groups. Generally, substitutions at Phe-1579 increased both k(on) and k(off), resulting in no prominent change in dissociation constant (K(d)). It seems that the smaller the molecular size of the residue at Na(v)1.4-Phe-1579, the larger the rates of k(on) and k(off), indicating that this site acts as a gate regulating access of toxin molecules to a receptor site. Substitutions at Tyr-1586 selectively increased k(off) but had virtually no effect on k(on), thus causing a drastic increase in K(d). At position Tyr-1586, a hydrophobic or aromatic amino acid side chain was required to maintain normal sensitivity to GTX. These results suggest that the residue at position Tyr-1586 has a more critical role in mediating GTX binding than the one at position Phe-1579. Here, we propose that the affinity of GTX to Na(v)1.4 sodium channels might be regulated by two residues (Phe and Tyr) at positions Phe-1579 and Tyr-1586, which, respectively, control access and binding of GTX to its receptor.  相似文献   

16.
We have previously shown that an antigenic site in native lysozyme resides around the disulphide bridge 30-115 and incorporates Lys-33 and Lys-116 and one or both of Tyr-20 and Tyr-23. These residues fall in an imaginary line circumscribing part of the surface of the molecule and passing through the spatially adjacent residues Tyr-20, Arg-21, Tyr-23, Lys-116, Asn-113, Arg-114, Phe-34 and Lys-33. The identity of the site was confirmed by demonstrating that the synthetic peptide Tyr-Arg-Tyr-Gly-Lys-Asn-Arg-Gly-Phe-Lys (which does not exist in lysozyme but simulates a surface region of it), and an analogue in which glycine replaced Tyr-23, possessed remarkable immuno-chemical reactivity that accounted entirely for the expected reactivity of the site in native lysozyme. Tyr-23 is not part of the site, and its contribution was satisfied by a glycine spacer. The novel approach presents a powerful technique for the delineation of antigenic (and other binding) sites in native proteins and confirms that these need not always comprise residues in direct peptide linkage.  相似文献   

17.
We present a mutational analysis of vaccinia topoisomerase that highlights the contributions of five residues in the catalytic domain (Phe-88 and Phe-101 in helix alpha1, Ser-204 in alpha5, and Lys-220 and Asn-228 in alpha6) to the DNA binding and transesterification steps. When augmented by structural information from exemplary type IB topoisomerases and tyrosine recombinases in different functional states, the results suggest how closure of the protein clamp around duplex DNA and assembly of a functional active site might be orchestrated by internal conformational changes in the catalytic domain. Lys-220 is a constituent of the active site, and a positive charge at this position is required for optimal DNA cleavage. Ser-204 and Asn-228 appear not to be directly involved in reaction chemistry at the scissile phosphodiester. We propose that (i) Asn-228 recruits the Tyr-274 nucleophile to the active site by forming a hydrogen bond to the main chain of the tyrosine-containing alpha8 helix and that (ii) contacts between Ser-204 and the DNA backbone upstream of the cleavage site trigger a separate conformational change required for active site assembly. Mutations of Phe-88 and Phe-101 affect DNA binding, most likely at the clamp closure step, which we posit to entail a distortion of helix alpha1.  相似文献   

18.
The biosynthesis of UDP-GlcNAc in bacteria is carried out by GlmU, an essential bifunctional uridyltransferase that catalyzes the CoA-dependent acetylation of GlcN-1-PO(4) to form GlcNAc-1-PO(4) and its subsequent condensation with UTP to form pyrophosphate and UDP-GlcNAc. As a metabolite, UDP-GlcNAc is situated at a branch point leading to the biosynthesis of lipopolysaccharide and peptidoglycan. Consequently, GlmU is regarded as an important target for potential antibacterial agents. The crystal structure of the Escherichia coli GlmU acetyltransferase active site has been determined in complexes with acetyl-CoA, CoA/GlcN-1-PO(4), and desulpho-CoA/GlcNAc-1-PO(4). These structures reveal the enzyme groups responsible for binding the substrates. A superposition of these complex structures suggests that the 2-amino group of GlcN-1-PO(4) is positioned in proximity to the acetyl-CoA to facilitate direct attack on its thioester by a ternary complex mechanism.  相似文献   

19.
The structure of the ternary complex of mycothiol synthase from Mycobacterium tuberculosis with bound desacetylmycothiol and CoA was determined to 1.8 A resolution. The structure of the acetyl-CoA-binary complex had shown an active site groove that was several times larger than its substrate. The structure of the ternary complex reveals that mycothiol synthase undergoes a large conformational change in which the two acetyltransferase domains are brought together through shared interactions with the functional groups of desacetylmycothiol, thereby decreasing the size of this large central groove. A comparison of the binary and ternary structures illustrates many of the features that promote catalysis. Desacetylmycothiol is positioned with its primary amine in close proximity and in the proper orientation for direct nucleophilic attack on the si-face of the acetyl group of acetyl-CoA. Glu-234 and Tyr-294 are positioned to act as a general base and general acid to promote acetyl transfer. In addition, this structure provides further evidence that the N-terminal acetyltransferase domain no longer has enzymatic activity and is vestigial in nature.  相似文献   

20.
Spermidine acetyltransferase (SAT) from Escherichia coli, which catalyses the transfer of acetyl groups from acetyl-CoA to spermidine, is a key enzyme in controlling polyamine levels in prokaryotic cells. In this study, we determined the crystal structure of SAT in complex with spermidine (SPD) and CoA at 2.5 Å resolution. SAT is a dodecamer organized as a hexamer of dimers. The secondary structural element and folding topology of the SAT dimer resemble those of spermidine/spermine N1-acetyltransferase (SSAT), suggesting an evolutionary link between SAT and SSAT. However, the polyamine specificity of SAT is distinct from that of SSAT and is promiscuous. The SPD molecule is also located at the inter-dimer interface. The distance between SPD and CoA molecules is 13 Å. A deep, highly acidic, water-filled cavity encompasses the SPD and CoA binding sites. Structure-based mutagenesis and in-vitro assays identified SPD-bound residues, and the acidic residues lining the walls of the cavity are mostly essential for enzymatic activities. Based on mutagenesis and structural data, we propose an acetylation mechanism underlying promiscuous polyamine recognition for SAT.  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