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 343 毫秒
1.
The substrate specificity of Escherichia coli N-acetylneuraminic acid lyase was previously switched from the natural condensation of pyruvate with N-acetylmannosamine, yielding N-acetylneuraminic acid, to the aldol condensation generating N-alkylcarboxamide analogues of N-acetylneuraminic acid. This was achieved by a single mutation of Glu192 to Asn. In order to analyze the structural changes involved and to more fully understand the basis of this switch in specificity, we have isolated all 20 variants of the enzyme at position 192 and determined the activities with a range of substrates. We have also determined five high-resolution crystal structures: the structures of wild-type E. coli N-acetylneuraminic acid lyase in the presence and in the absence of pyruvate, the structures of the E192N variant in the presence and in the absence of pyruvate, and the structure of the E192N variant in the presence of pyruvate and a competitive inhibitor (2R,3R)-2,3,4-trihydroxy-N,N-dipropylbutanamide. All structures were solved in space group P21 at resolutions ranging from 1.65 Å to 2.2 Å. A comparison of these structures, in combination with the specificity profiles of the variants, reveals subtle differences that explain the details of the specificity changes. This work demonstrates the subtleties of enzyme-substrate interactions and the importance of determining the structures of enzymes produced by directed evolution, where the specificity determinants may change from one substrate to another.  相似文献   

2.
Germinated, unpolished rice was found to contain a substantial amount of D-serine, with the ratio of the D-enantiomer to the L-enantiomer being higher for serine than for other amino acids. The relative amount of D-serine (D/(D + L)%) reached approximately 10% six days after germination. A putative serine racemase gene (serr, clone No. 001-110-B03) was found in chromosome 4 of the genomic DNA of Oryza sativa L. ssp. Japonica cv. Nipponbare. This was expressed as serr in Escherichia coli and its gene product (SerR) was purified to apparent homogeneity. SerR is a homodimer with a subunit molecular mass of 34.5 kDa, and is highly specific for serine. In addition to a serine racemase reaction, SerR catalyzes D- and L-serine dehydratase reactions, for which the specific activities were determined to be 2.73 and 1.42 nkatal/mg, respectively. The optimum temperature and pH were respectively determined for the racemase reaction (35 °C and pH 9.0) and for the dehydratase reaction (35 °C and pH 9.5). SerR was inhibited by PLP-enzyme inhibitors. ATP decreased the serine racemase activity of SerR but increased the serine dehydratase activity. Kinetic analysis showed that Mg2+ increases the catalytic efficiency of the serine racemase activity of SerR and decreases that of the serine dehydratase activity. Fluorescence-quenching analysis of the tryptophan residues in SerR indicated that the structure of SerR is distorted by the addition of Mg2+, and this structural change probably regulates the two enzymatic activities.  相似文献   

3.
The GroEL/GroES protein folding chamber is formed and dissociated by ATP binding and hydrolysis. ATP hydrolysis in the GroES-bound (cis) ring gates entry of ATP into the opposite unoccupied trans ring, which allosterically ejects cis ligands. While earlier studies suggested that hydrolysis of cis ATP is the rate-limiting step of the cycle (t½ ∼ 10 s), a recent study suggested that ADP release from the cis ring may be rate-limiting (t½ ∼ 15-20 s). Here we have measured ADP release using a coupled enzyme assay and observed a t½ for release of ?4-5 s, indicating that this is not the rate-limiting step of the reaction cycle.  相似文献   

4.
To clarify the mechanism for substrate recognition of α-aminoadipate aminotransferase (AAA-AT) from Thermus thermophilus, the crystal structure of AAA-AT complexed with N-(5′-phosphopyridoxyl)-l-glutamate (PPE) was determined at 1.67 Å resolution. The crystal structure revealed that PPE is recognized by amino acid residues the same as those seen in N-(5′-phosphopyridoxyl)-l-α-aminoadipate (PPA) recognition; however, to bind the γ-carboxyl group of Glu at a fixed position, the Cα atom of the Glu moiety moves 0.80 Å toward the γ-carboxyl group in the PPE complex. Markedly decreased activity for Asp can be explained by the shortness of the aspartyl side chain to be recognized by Arg23 and further dislocation of the Cα atom of bound Asp. Site-directed mutagenesis revealed that Arg23 has dual functions for reaction, (i) recognition of γ (δ)-carboxyl group of Glu (AAA) and (ii) rearrangement of α2 helix by changing the interacting partners to place the hydrophobic substrate at the suitable position.  相似文献   

5.
Acid sensing ion channels (ASICs), Ca2+ and voltage-activated potassium channels (BK) are widely present throughout the central nervous system. Previous studies have shown that when expressed together in heterologous cells, ASICs inhibit BK channels, and this inhibition is relieved by acidic extracellular pH. We hypothesized that ASIC and BK channels might interact in neurons, and that ASICs may regulate BK channel activity. We found that ASICs inhibited BK currents in cultured wild-type cortical neurons, but not in ASIC1a/2/3 triple knockout neurons. The inhibition in the wild-type was partially relieved by a drop in extracellular pH to 6. To test the consequences of ASIC-BK interaction for neuronal excitability, we compared action potential firing in cultured cortical neurons from wild-type and ASIC1a/2/3 null mice. We found that in the knockout, action potentials were narrow and exhibited increased after-hyperpolarization. Moreover, the excitability of these neurons was significantly increased. These findings are consistent with increased BK channel activity in the neurons from ASIC1a/2/3 null mice. Our data suggest that ASICs can act as endogenous pH-dependent inhibitors of BK channels, and thereby can reduce neuronal excitability.  相似文献   

6.
Hydroxymandelate synthase (HMS) catalyzes the committed step in the formation of para-hydroxyphenylglycine, a recurrent substructure of polycyclic non-ribosomal peptide antibiotics such as vancomycin. HMS uses the same substrates as 4-hydroxyphenylpyruvate dioxygenase (HPPD), 4-hydroxyphenylpyruvate (HPP) and O2, and also conducts a dioxygenation reaction. The difference between the two lies in the insertion of the second oxygen atom, HMS directing this atom onto the benzylic carbon of the substrate while HPPD hydroxylates the aromatic C1 carbon. We have shown that HMS will bind NTBC, a herbicide/therapeutic whose mode of action is based on the inhibition of HPPD. This occurs despite residue differences at the active site of HMS from those known to contact the inhibitor in HPPD. Moreover, the minimal kinetic mechanism for association of NTBC to HMS differs only slightly from that observed with HPPD. The primary difference is that three charge-transfer species are observed to accumulate during association. The first reversible complex forms with a weak dissociation constant of 520 μM, the subsequent two charge-transfer complexes form with rate constants of 2.7 s−1 and 0.67 s−1. As was the case for HPPD, the final complex has the most intense charge-transfer, is not observed to dissociate, and is unreactive towards dioxygen.  相似文献   

7.
A new ligand LH (where LH = N-(picolinoyl)-biurate) has been prepared and characterized. The presence of three amide linkages make this ligand sufficiently flexible to act as N,N,O donor tridentate blocking ligand in the formation of a one dimensional metal-ligand layer like structure. Reaction of LH and dicyanamide (dca) with Co(NO3)2 · 6H2O gives [CoL(dca)]n (1). In this compound picolinamide modulated ligand L coordinated the central Co(II) ion in a meridonal-fashion. The single crystal X-ray crystallography revealed that in 1, dca acts as μ1,5− singly bridging ligand whereas μ1,5− doubly bridging is the more common type. This gives rise to the 1D undulated waves like structure. The Co(II) centre is surrounded in a distorted square pyramidal coordination geometry. The variable temperature magnetic (VTM) susceptibility measurements show that the global feature of the χMT versus T curve for 1 is characteristic of very weak antiferromagnetic interactions through the dicyanamide ligand and between 300 and 5 K the best fit parameter was determined as J = −3.52 cm−1. The X-ray structure, VTM study and UV-Vis spectrum of the compound show that 1 is a low-spin square-pyramidal compound whereas high-spin compounds are more common for the five coordinated cobalt (II) compounds. The X-band EPR spectrum of 1 at room temperature shows only one isotropic band centred at g = 2.08.  相似文献   

8.
Previous in vitro motility assays using bipolar myosin thick filaments demonstrated that actin filaments were capable of moving in both directions along the myosin filament tracks. The movements; however, were slower in the direction leading away from the central bare zone than towards it. To understand the mechanism underlying these different direction-dependent motilities, we have examined the effects of temperature on the velocities of the bidirectional movements along reconstituted myosin filaments. Activation energies of the movements were determined by Arrhenius plots at high and low concentrations of ATP. As a result, the thermal activation energy of the movement away from the central bare zone was significantly higher than that of the movement toward the zone. Given that the backward movement away from the central bare zone would cause the myosin heads to be constrained and the stiffness of the cross-bridges to increase, these results suggest that elastic energy required for the cross-bridge transition is supplied by thermal fluctuations.  相似文献   

9.
Electron transfer reactions between optically-active RuII/III complexes incorporating (S)-/(R)-amino acids, and the two azurins, azurin-1 (az-1Cu) and azurin-2 (az-2Cu) isolated from Alcaligenes xylosoxidans GIFU 1051, have been studied to probe molecular recognition sites on the two azurins. The RuII/III complexes are K[RuII(L)(bpy)] and [RuIII(L)(bpy)], and have a tripodal ligand (L) derived from the (S)-/(R)-amino acids, which are in turn exchanged for other functional substituent groups, such as (S)-/(R)-phenylalanine, -leucine, -valine, -alanine, and -glutamic acid (L = (S)-/(R)-BCMPA, -BCMLE, -BCMVA, -BCMAL, and -BCMGA). In the oxidation reaction of az-1CuI promoted by the RuIII complexes, the kinetic parameters exhibited enantio- and stereo-selectivities, while the same reaction of az-2CuI was less enantio- and stereo-selective. These differences suggest that the processes of formation of the activated states are different for the two azurins. On the other hand, such a difference has not been observed for az-1 and az-2 with respect to the reduction reactions promoted by both azurins CuII by the RuII complexes within the experimental error. This suggests that the neutrality of the Ru complexes is important for precise molecular recognition of azurins. His117 has been proposed as the electron transfer site. The local structures in the vicinity of the His117 side chain in the two azurins, are essentially identical with the exception of the 43rd residue, Val43 and Ala43 for az-1 and az-2, respectively. Electron transfer reactions between RuIII complexes and a mutant azurin, V43A-az-1, were also carried out. Interestingly, the activation parameters estimated were very similar to those of az-2, indicating that the 43rd residue acts as the electron transfer site in azurins and provides rationalization for the different mechanisms of az-1 and az-2 in redox reactions.  相似文献   

10.
Petr Paucek  Martin Jab?rek 《BBA》2004,1659(1):83-91
The Na+/Ca2+ antiporter was purified from beef heart mitochondria and reconstituted into liposomes containing fluorescent probes selective for Na+ or Ca2+. Na+/Ca2+ exchange was strongly inhibited at alkaline pH, a property that is relevant to rapid Ca2+ oscillations in mitochondria. The effect of pH was mediated entirely via an effect on the Km for Ca2+. When present on the same side as Ca2+, K+ activated exchange by lowering the Km for Ca2+ from 2  to 0.9 μM. The Km for Na+ was 8 mM. In the absence of Ca2+, the exchanger catalyzed high rates of Na+/Li+ and Na+/K+ exchange. Diltiazem and tetraphenylphosphonium cation inhibited both Na+/Ca2+ and Na+/K+ exchange with IC50 values of 10 and 0.6 μM, respectively. The Vmax for Na+/Ca2+ exchange was increased about fourfold by bovine serum albumin, an effect that may reflect unmasking of an autoregulatory domain in the carrier protein.  相似文献   

11.
The 5′-nucleotidase (5′-ribonucleoside phosphohydrolase, EC 3.1.3.5) from bovine milk fat globule membranes was partially purified. Two separate peaks of activity were obtained from a Sepharose column and the two fractions, designated V and VI in order of elution, were collected and characterized separately. Both V and VI exhibited pH optima between 7.0 and 7.5 for AMP, GMP and CMP in the absence of metal ions. In the presence of Mg2+, a second pH optimum at 10.0 was observed with both fractions. Low concentrations of MnCl2 activated Fraction V but not Fraction VI. HgCl2 was a potent inhibitor of both fractions. The relative rates of hydrolysis of various 5′-mononucleotides differed comparing the two fractions. Optimum temperature for Pi release was 69 °C for both fractions. Activation energies were 10 400 cal/mole and 9600 cal/mole for Fractions V and VI, respectively. For V, calculated Km values for AMP, GMP and CMP were 0.94, 2.5 and 1.16 mM, respectively. Calculated Km values for Fraction VI for AMP, GMP and CMP were 5.0, 3.95 and 1.73 mM, respectively. ATP was a competitive inhibitor of AMP hydrolysis by Fraction V and a noncompetitive inhibitor of AMP hydrolysis by Fraction VI. Both fractions contained chloroform-methanol-extractable phospholipid. The phospholipid distribution pattern of Fraction VI was similar to that of milk fat globule membranes. Fraction V contained only sphingomyelin and phosphatidylcholine. It is proposed that milk fat globule membranes contain two separate 5′-nucleotidases.  相似文献   

12.
The squid Watasenia scintillans emits blue light from numerous photophores. According to Tsuji [F.I. Tsuji, Bioluminescence reaction catalyzed by membrane-bound luciferase in the “firefly squid”, Watasenia scintillans, Biochim. Biophys. Acta 1564 (2002) 189–197.], the luminescence from arm light organs is caused by an ATP-dependent reaction involving Mg2+, coelenterazine disulfate (luciferin), and an unstable membrane-bound luciferase. We stabilized and partially purified the luciferase in the presence of high concentrations of sucrose, and obtained it as particulates (average size 0.6–2 µm). The ATP-dependent luminescence reaction of coelenterazine disulfate catalyzed by the particulate luciferase was investigated in detail. Optimum temperature of the luminescence reaction is about 5 °C. Coelenterazine disulfate is a strictly specific substrate in this luminescence system; any modification of its structure resulted in a very heavy loss in its light emission capability. The light emitter is the excited state of the amide anion form of coelenteramide disulfate. The quantum yield of coelenterazine disulfate is calculated at 0.36. ATP could be replaced by ATP-γ-S, but not by any other analogues tested. The amount of AMP produced in the luminescence reaction was much smaller than that of coelenteramide disulfate, suggesting that the reaction mechanism of the Watasenia bioluminescence does not involve the formation of adenyl luciferin as an intermediate.  相似文献   

13.
The biotransformation and cytotoxic effects of hydroxychavicol (HC; 1-allyl-3,4-dihydroxybenzene), which is a catecholic component in piper betel leaf and a major intermediary metabolite of safrole in rats and humans, was studied in freshly isolated rat hepatocytes. The exposure of hepatocytes to HC caused not only concentration (0.25-1.0 mM)- and time (0-3 h)-dependent cell death accompanied by the loss of cellular ATP, adenine nucleotide pools, reduced glutathione, and protein thiols, but also the accumulation of glutathione disulfide and malondialdehyde, indicating lipid peroxidation. At a concentration of 1 mM, the cytotoxic effects of safrole were less than those of HC. The loss of mitochondrial membrane potential and generation of oxygen radical species assayed using 2′,7′-dichlorodihydrofluoresein diacetate (DCFH-DA) in hepatocytes treated with HC were greater than those with safrole. HC at a weakly toxic level (0.25 and/or 0.50 mM) was metabolized to monoglucuronide, monosulfate, and monoglutathione conjugates, which were identified by mass spectra and/or 1H nuclear magnetic resonance spectra. The amounts of sulfate rather than glucuronide or glutathione conjugate predominantly increased, accompanied by a loss of the parent compound, with time. In hepatocytes pretreated with either diethyl maleate or salicylamide, HC-induced cytotoxicity was enhanced, accompanied by a decrease in the formation of these conjugates and by the inhibition of HC loss. Taken collectively, our results indicate that (a) mitochondria are target organelles for HC, which elicits cytotoxicity through mitochondrial failure related to mitochondrial membrane potential at an early stage and subsequently lipid peroxidation through oxidative stress at a later stage; (b) the onset of cytotoxicity depends on the initial and residual concentrations of HC rather than those of its metabolites; (c) the toxicity of HC is greater than that of safrole, suggesting the participation of a catecholic intermediate in safrole cytotoxicity in rat hepatocytes.  相似文献   

14.
Ketol-acid reductoisomerase (KARI; EC 1.1.1.86) is an enzyme in the branched-chain amino acid biosynthesis pathway where it catalyzes the conversion of 2-acetolactate into (2R)-2,3-dihydroxy-3-isovalerate or the conversion of 2-aceto-2-hydroxybutyrate into (2R,3R)-2,3-dihydroxy-3-methylvalerate. KARI catalyzes two reactions—alkyl migration and reduction—and requires Mg2+ and NADPH for activity. To date, the only reported structures for a plant KARI are those of the spinach enzyme-Mn2+-(phospho)ADP ribose-(2R,3R)-2,3-dihydroxy-3-methylvalerate complex and the spinach KARI-Mg2+-NADPH-N-hydroxy-N-isopropyloxamate complex, where N-hydroxy-N-isopropyloxamate is a predicted transition-state analog. These studies demonstrated that the enzyme consists of two domains, N-domain and C-domain, with the active site at the interface of these domains. Here, we have determined the structures of the rice KARI-Mg2+ and rice KARI-Mg2+-NADPH complexes to 1.55 Å and 2.80 Å resolutions, respectively. In comparing the structures of all the complexes, several differences are observed. Firstly, the N-domain is rotated up to 15° relative to the C-domain, expanding the active site by up to 4 Å. Secondly, an α-helix in the C-domain that includes residues V510-T519 and forms part of the active site moves by ∼ 3.9 Å upon binding of NADPH. Thirdly, the 15 C-terminal amino acid residues in the rice KARI-Mg2+ complex are disordered. In the rice KARI-Mg2+-NADPH complex and the spinach KARI structures, many of the 15 residues bind to NADPH and the N-domain and cover the active site. Fourthly, the location of the metal ions within the active site can vary by up to 2.7 Å. The new structures allow us to propose that an induced-fit mechanism operates to (i) allow substrate to enter the active site, (ii) close over the active site during catalysis, and (iii) open the active site to facilitate product release.  相似文献   

15.
The direct induction of apoptosis has emerged as a powerful anticancer strategy, and small molecules that either inhibit or activate certain proteins in the apoptotic pathway have great potential as novel chemotherapeutic agents. Central to apoptosis is the activation of the zymogen procaspase-3 to caspase-3. Caspase-3 is the key “executioner” caspase, catalyzing the hydrolysis of a multitude of protein substrates within the cell. Interestingly, procaspase-3 levels are often elevated in cancer cells, suggesting a compound that directly stimulates the activation of procaspase-3 to caspase-3 could selectively induce apoptosis in cancer cells. We recently reported the discovery of a compound, PAC-1, which enhances procaspase-3 activity in vitro and induces apoptotic death in cancer cells in culture and in mouse xenograft models. Described herein is the mechanism by which PAC-1 activates procaspase-3 in vitro. We show that zinc inhibits the enzymatic activity of procaspase-3 and that PAC-1 strongly activates procaspase-3 in buffers that contain zinc. PAC-1 and zinc form a tight complex with one another, with a dissociation constant of approximately 42 nM. The combined data indicate that PAC-1 activates procaspase-3 in vitro by sequestering inhibitory zinc ions, thus allowing procaspase-3 to autoactivate itself to caspase-3. The small-molecule-mediated activation of procaspases has great therapeutic potential and thus this discovery of the in vitro mechanism of action of PAC-1 is critical to the development and optimization of other procaspase-activating compounds.  相似文献   

16.
17.
Unlike superoxide dismutases (SODs), superoxide reductases (SORs) eliminate superoxide anion (O2•−) not through its dismutation, but via reduction to hydrogen peroxide (H2O2) in the presence of an electron donor. The microaerobic protist Giardia intestinalis, responsible for a common intestinal disease in humans, though lacking SOD and other canonical reactive oxygen species-detoxifying systems, is among the very few eukaryotes encoding a SOR yet identified. In this study, the recombinant SOR from Giardia (SORGi) was purified and characterized by pulse radiolysis and stopped-flow spectrophotometry. The protein, isolated in the reduced state, after oxidation by superoxide or hexachloroiridate(IV), yields a resting species (Tfinal) with Fe3+ ligated to glutamate or hydroxide depending on pH (apparent pKa = 8.7). Although showing negligible SOD activity, reduced SORGi reacts with O2•− with a pH-independent second-order rate constant k1 = 1.0 × 109 M− 1 s− 1 and yields the ferric-(hydro)peroxo intermediate T1; this in turn rapidly decays to the Tfinal state with pH-dependent rates, without populating other detectable intermediates. Immunoblotting assays show that SORGi is expressed in the disease-causing trophozoite of Giardia. We propose that the superoxide-scavenging activity of SOR in Giardia may promote the survival of this air-sensitive parasite in the fairly aerobic proximal human small intestine during infection.  相似文献   

18.
The molecular mechanism responsible for the regulation of the mitochondrial membrane proton conductance (G) is not clearly understood. This study investigates the role of the transmembrane potential (ΔΨm) using planar membranes, reconstituted with purified uncoupling proteins (UCP1 and UCP2) and/or unsaturated FA. We show that high ΔΨm (similar to ΔΨm in mitochondrial State IV) significantly activates the protonophoric function of UCPs in the presence of FA. The proton conductance increases nonlinearly with ΔΨm. The application of ΔΨm up to 220 mV leads to the overriding of the protein inhibition at a constant ATP concentration. Both, the exposure of FA-containing bilayers to high ΔΨm and the increase of FA membrane concentration bring about the significant exponential Gm increase, implying the contribution of FA in proton leak. Quantitative analysis of the energy barrier for the transport of FA anions in the presence and absence of protein suggests that FA remain exposed to membrane lipids while crossing the UCP-containing membrane. We believe this study shows that UCPs and FA decrease ΔΨm more effectively if it is sufficiently high. Thus, the tight regulation of proton conductance and/or FA concentration by ΔΨm may be key in mitochondrial respiration and metabolism.  相似文献   

19.
We have compared the physical properties of a 15.51-kb constitutive heterochromatin segment and a 16.17-kb facultative heterochromatin segment that form part of the chicken β-globin locus. These segments were excised from an avian erythroleukemia cell line by restriction enzyme digestion and released from the nucleus, thus allowing measurement of the sedimentation coefficients by use of calibrated sucrose gradients. A determination of the buoyant density of the cross-linked particle in CsCl led to the total mass of the particles and their frictional coefficients, f. Despite the slight differences in nucleosome density, the measured value of f for both fragments was consistent with a rodlike particle having a diameter of 33-45 nm and a length corresponding to approximately six to seven nucleosomes per 11-nm turn. At higher ionic strengths we found no evidence of any abrupt conformational change, demonstrating that these chromatin fragments released from the nucleus did not assume the more compact conformations recently described for some reconstituted structures.  相似文献   

20.
Lu Lu  Jie Nan  Lan-Fen Li  Xiao-Dong Su  Yi Li 《FEBS letters》2010,584(16):3533-3539
Microtubules are composed of polymerized α/β-tubulin heterodimers. Biogenesis of assembly-competent tubulin dimers is a complex multistep process that requires sequential actions of distinct molecular chaperones and cofactors. Tubulin folding cofactor A (TFCA), which captures β-tubulin during the folding pathway, has been identified in many organisms. Here, we report the crystal structure of Arabidopsis thaliana TFC A (KIESEL, KIS), which forms a monomeric three-helix bundle. The functional binding analysis demonstrated that KIS interacts with β-tubulin in plant. Furthermore, mutagenesis studies indicated that the α-helical regions of KIS participate in β-tubulin binding. Unlike the budding yeast TFC A, the two loop regions of KIS are not required for this interaction suggesting a distinct binding mechanism of TFC A to β-tubulin in plants.

Structured summary

MINT-7968902, MINT-7968915, MINT-7968951, MINT-7968966: KIS (uniprotkb:O04350) physically interacts (MI:0915) with Tub9 (uniprotkb:P29517) by anti tag coimmunoprecipitation (MI:0007)MINT-7968928: KIS (uniprotkb:O04350) and Tub9 (uniprotkb:P29517) physically interact (MI:0915) by bimolecular fluorescence complementation (MI:0809)  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