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.
Recombinant glycerol dehydratase of Klebsiella pneumoniae was purified to homogeneity. The subunit composition of the enzyme was most probably alpha 2 beta 2 gamma 2. When (R)- and (S)-propane-1,2-diols were used independently as substrates, the rate with the (R)-enantiomer was 2.5 times faster than that with the (S)-isomer. In contrast to diol dehydratase, an isofunctional enzyme, the affinity of the enzyme for the (S)-isomer was essentially the same or only slightly higher than that for the (R)-isomer (Km(R)/Km(S) = 1.5). The crystal structure of glycerol dehydratase in complex with cyanocobalamin and propane-1,2-diol was determined at 2.1 A resolution. The enzyme exists as a dimer of the alpha beta gamma heterotrimer. Cobalamin is bound at the interface between the alpha and beta subunits in the so-called 'base-on' mode with 5,6-dimethylbenzimidazole of the nucleotide moiety coordinating to the cobalt atom. The electron density of the cyano group was almost unobservable, suggesting that the cyanocobalamin was reduced to cob(II)alamin by X-ray irradiation. The active site is in a (beta/alpha)8 barrel that was formed by a central region of the alpha subunit. The substrate propane-1,2-diol and essential cofactor K+ are bound inside the (beta/alpha)8 barrel above the corrin ring of cobalamin. K+ is hepta-coordinated by the two hydroxyls of the substrate and five oxygen atoms from the active-site residues. These structural features are quite similar to those of diol dehydratase. A closer contact between the alpha and beta subunits in glycerol dehydratase may be reminiscent of the higher affinity of the enzyme for adenosylcobalamin than that of diol dehydratase. Although racemic propane-1,2-diol was used for crystallization, the substrate bound to glycerol dehydratase was assigned to the (R)-isomer. This is in clear contrast to diol dehydratase and accounts for the difference between the two enzymes in the susceptibility of suicide inactivation by glycerol.  相似文献   

2.
Shibata N  Masuda J  Morimoto Y  Yasuoka N  Toraya T 《Biochemistry》2002,41(42):12607-12617
Substrate binding triggers catalytic radical formation through the cobalt-carbon bond homolysis in coenzyme B12-dependent enzymes. We have determined the crystal structure of the substrate-free form of Klebsiella oxytoca diol dehydratase*cyanocobalamin complex at 1.85 A resolution. The structure contains two units of the heterotrimer consisting of alpha, beta, and gamma subunits. As compared with the structure of its substrate-bound form, the beta subunits are tilted by approximately 3 degrees and cobalamin is also tilted so that pyrrole rings A and D are significantly lifted up toward the substrate-binding site, whereas pyrrole rings B and C are only slightly lifted up. The structure revealed that the potassium ion in the substrate-binding site of the substrate-free enzyme is also heptacoordinated; that is, two oxygen atoms of two water molecules coordinate to it instead of the substrate hydroxyls. A modeling study in which the structures of both the cobalamin moiety and the adenine ring of the coenzyme were superimposed onto those of the enzyme-bound cyanocobalamin and the adenine ring-binding pocket, respectively, demonstrated that the distortions of the Co-C bond in the substrate-free form are already marked but slightly smaller than those in the substrate-bound form. It was thus strongly suggested that the Co-C bond becomes largely activated (labilized) when the coenzyme binds to the apoenzyme even in the absence of substrate and undergoes homolysis through the substrate-induced conformational changes of the enzyme. Kinetic coupling of Co-C bond homolysis with hydrogen abstraction from the substrate shifts the equilibrium to dissociation.  相似文献   

3.
Sun L  Warncke K 《Proteins》2006,64(2):308-319
The structure of the EutB protein from Salmonella typhimurium, which contains the active site of the coenzyme B12 (adenosylcobalamin)-dependent enzyme, ethanolamine ammonia-lyase, has been predicted by using structural proteomics techniques of comparative modelling. The 453-residue EutB protein displays no significant sequence identity with proteins of known structure. Therefore, secondary structure prediction and fold recognition algorithms were used to identify templates. Multiple three-dimensional template matching (threading) servers identified predominantly beta8alpha8, TIM-barrel proteins, and in particular, the large subunits of diol dehydratase (PDB: 1eex:A, 1dio:A) and glycerol dehydratase (PDB: 1mmf:A), as templates. Consistent with this identification, the dehydratases are, like ethanolamine ammonia-lyase, Class II coenzyme B12-dependent enzymes. Model building was performed by using MODELLER. Models were evaluated by using different programs, including PROCHECK and VERIFY3D. The results identify a beta8alpha8, TIM-barrel fold for EutB. The beta8alpha8, TIM-barrel fold is consistent with a central role of the alpha/beta-barrel structures in radical catalysis conducted by the coenzyme B12- and S-adenosylmethionine-dependent (radical SAM) enzyme superfamilies. The EutB model and multiple sequence alignment among ethanolamine ammonia-lyase, diol dehydratase, and glycerol dehydratase from different species reveal the following protein structural features: (1) a "cap" loop segment that closes the N-terminal region of the barrel, (2) a common cobalamin cofactor binding topography at the C-terminal region of the barrel, and (3) a beta-barrel-internal guanidinium group from EutB R160 that overlaps the position of the active-site potassium ion found in the dehydratases. R160 is proposed to have a role in substrate binding and radical catalysis.  相似文献   

4.
BACKGROUND: Adenosylcobalamin (coenzyme B(12)) serves as a cofactor for enzymatic radical reactions. The adenosyl radical, a catalytic radical in these reactions, is formed by homolysis of the cobalt-carbon bond of the coenzyme, although the mechanism of cleavage of its organometallic bond remains unsolved. RESULTS: We determined the three-dimensional structures of diol dehydratase complexed with adeninylpentylcobalamin and with cyanocobalamin at 1.7 A and 1.9 A resolution, respectively, at cryogenic temperatures. In the adeninylpentylcobalamin complex, the adenine ring is bound parallel to the corrin ring as in the free form and methylmalonyl-CoA-mutase-bound coenzyme, but with the other side facing pyrrole ring C. All of its nitrogen atoms except for N(9) are hydrogen-bonded to mainchain amide oxygen and amide nitrogen atoms, a sidechain hydroxyl group, and a water molecule. As compared with the cyanocobalamin complex, the sidechain of Seralpha224 rotates by 120 degrees to hydrogen bond with N(3) of the adenine ring. CONCLUSIONS: The structure of the adenine-ring-binding site provides a molecular basis for the strict specificity of diol dehydratase for the coenzyme adenosyl group. The superimposition of the structure of the free coenzyme on that of enzyme-bound adeninylpentylcobalamin demonstrated that the tight enzyme-coenzyme interactions at both the cobalamin moiety and adenine ring of the adenosyl group would inevitably lead to cleavage of the cobalt-carbon bond. Rotation of the ribose moiety around the glycosidic linkage makes the 5'-carbon radical accessible to the hydrogen atom of the substrate to be abstracted.  相似文献   

5.
The direct ion-dipolar interactions between potassium ion (K(+)) and the two hydroxyl groups of the substrate are the most striking feature of the crystal structure of coenzyme B(12)-dependent diol dehydratase. We carried out density-functional-theory computations to determine whether K(+) can assist the 1,2-shift of the hydroxyl group in the substrate-derived radical. Between a stepwise abstraction/recombination reaction proceeding via a direct hydroxide abstraction by K(+) and a concerted hydroxyl group migration assisted by K(+), only a transition state for the latter concerted mechanism was found from our computations. The barrier height for the transition state from the complexed radical decreases by only 2.3 kcal/mol upon coordination of the migrating hydroxyl group to K(+), which corresponds to a 42-fold rate acceleration at 37 degrees C. The net binding energy upon replacement of the K(+)-bound water for substrate was calculated to be 10.7 kcal/mol. It can be considered that such a large binding energy is at least partly used for the substrate-induced conformational changes in the enzyme that trigger the homolytic cleavage of the Co-C bond of the coenzyme and the subsequent catalysis by a radical mechanism. We propose here a new mechanism for diol dehydratase in which K(+) plays a direct role in the catalysis.  相似文献   

6.
The three gldCDE genes from Lactobacillus diolivorans, that encode the three subunits of the glycerol dehydratase, were cloned and the proteins were co-expressed in soluble form in Escherichia coli with added sorbitol and betaine hydrochloride. The purified enzyme exists as a heterohexamer (α2β2γ2) structure with a native molecular mass of 210 kDa. It requires coenzyme B12 for catalytic activity and is subject to suicide inactivation by glycerol during catalysis. The enzyme had maximum activity at pH 8.6 and 37 °C. The apparent K m values for coenzyme B12, 1,2-ethanediol, 1,2-propanediol, and glycerol were 1.5 μM, 10.5 mM, 1.3 mM, and 5.8 mM, respectively. Together, these results indicated that the three genes gldCDE encoding the proteins make up a coenzyme B12-dependent diol dehydratase and not a glycerol dehydratase.  相似文献   

7.
N-terminal truncation of the Escherichia coli ethanolamine ammonia-lyase β-subunit does not affect the catalytic properties of the enzyme (Akita, K., Hieda, N., Baba, N., Kawaguchi, S., Sakamoto, H., Nakanishi, Y., Yamanishi, M., Mori, K., and Toraya, T. (2010) J. Biochem. 147, 83–93). The binary complex of the truncated enzyme with cyanocobalamin and the ternary complex with cyanocobalamin or adeninylpentylcobalamin and substrates were crystallized, and their x-ray structures were analyzed. The enzyme exists as a trimer of the (αβ)2 dimer. The active site is in the (β/α)8 barrel of the α-subunit; the β-subunit covers the lower part of the cobalamin that is bound in the interface of the α- and β-subunits. The structure complexed with adeninylpentylcobalamin revealed the presence of an adenine ring-binding pocket in the enzyme that accommodates the adenine moiety through a hydrogen bond network. The substrate is bound by six hydrogen bonds with active-site residues. Argα160 contributes to substrate binding most likely by hydrogen bonding with the O1 atom. The modeling study implies that marked angular strains and tensile forces induced by tight enzyme-coenzyme interactions are responsible for breaking the coenzyme Co–C bond. The coenzyme adenosyl radical in the productive conformation was modeled by superimposing its adenine ring on the adenine ring-binding site followed by ribosyl rotation around the N-glycosidic bond. A major structural change upon substrate binding was not observed with this particular enzyme. Gluα287, one of the substrate-binding residues, has a direct contact with the ribose group of the modeled adenosylcobalamin, which may contribute to the substrate-induced additional labilization of the Co–C bond.  相似文献   

8.
9.
Two previously undescribed flavonol tetraglycosides, isorhamnetin-3-O-α-l-rhamnopyranosyl-(1→6)-β-d-galactopyranosyl-(1→4)-α-l-rhamnopyranosyl-(1→6)-β-d-glucopyranoside (1) and isorhamnetin-3-O-α-l-rhamnopyranosyl-(1→6)-β-d-galactopyranosyl-(1→4)-α-l-rhamnopyranosyl-(1→6)-β-d-galactopyranoside (2), along with nine known compounds including seven flavonoids and two lignans, were isolated from the leaves of Opilia amentacea Roxb (Opiliaceae). Their structures were established on the basis of spectroscopic analysis. The DPPH radical scavenging activity of compounds 111 was evaluated. In addition, all compounds were evaluated for their tyrosinase inhibitions by using in vitro mushroom tyrosinase assay. Only 5,5-dimethoxylariciresinol-4-O-β-d-glucopyranoside (10) and eleutheroside E1 (11) exhibited significant tyrosinase inhibition (IC50 42.1 and 28 μM, respectively) and DPPH radical scavenging activity (IC50 85.1 and 42.1 μM, respectively) compared with the positive controls.  相似文献   

10.
Kajiura H  Mori K  Shibata N  Toraya T 《The FEBS journal》2007,274(21):5556-5566
Adenosylcobalamin-dependent diol and glycerol dehydratases are isofunctional enzymes and undergo mechanism-based inactivation by a physiological substrate glycerol during catalysis. Inactivated holoenzymes are reactivated by their own reactivating factors that mediate the ATP-dependent exchange of an enzyme-bound, damaged cofactor for free adenosylcobalamin through intermediary formation of apoenzyme. The reactivation takes place in two steps: (a) ADP-dependent cobalamin release and (b) ATP-dependent dissociation of the resulting apoenzyme-reactivating factor complexes. The in vitro experiments with purified proteins indicated that diol dehydratase-reactivating factor (DDR) cross-reactivates the inactivated glycerol dehydratase, whereas glycerol dehydratase-reactivating factor (GDR) did not cross-reactivate the inactivated diol dehydratase. We investigated the molecular basis of their specificities in vitro by using purified preparations of cognate and noncognate enzymes and reactivating factors. DDR mediated the exchange of glycerol dehydratase-bound cyanocobalamin for free adeninylpentylcobalamin, whereas GDR cannot mediate the exchange of diol dehydratase-bound cyanocobalamin for free adeninylpentylcobalamin. As judged by denaturing PAGE, the glycerol dehydratase-DDR complex was cross-formed, although the diol dehydratase-GDR complex was not formed. There were no specificities of reactivating factors in the ATP-dependent dissociation of enzyme-reactivating factor complexes. Thus, it is very likely that the specificities of reactivating factors are determined by the capability of reactivating factors to form complexes with apoenzymes. A modeling study based on the crystal structures of enzymes and reactivating factors also suggested why DDR cross-forms a complex with glycerol dehydratase, and why GDR does not cross-form a complex with diol dehydratase.  相似文献   

11.
In extracts of polyethylene glycol (PEG)-grown cells of the strictly anaerobically fermenting bacterium Pelobacter venetianus, two different enzyme activities were detected, a diol dehydratase and a PEG-degrading enzyme which was characterized as a PEG acetaldehyde lyase. Both enzymes were oxygen sensitive and depended on a reductant, such as titanium citrate or sulfhydryl compounds, for optimal activity. The diol dehydratase was inhibited by various corrinoids (adenosylcobalamin, cyanocobalamin, hydroxocobalamin, and methylcobalamin) by up to 37% at a concentration of 100 μM. Changes in ionic strength and the K+ ion concentration had only limited effects on this enzyme activity; glycerol inhibited the enzyme by 95%. The PEG-degrading enzyme activity was stimulated by the same corrinoids by up to 80%, exhibited optimal activity in 0.75 M potassium phosphate buffer or in the presence of 4 M KCI, and was only slightly affected by glycerol. Both enzymes were located in the cytoplasmic space. Also, another PEG-degrading bacterium, Bacteroides strain PG1, contained a PEG acetaldehyde lyase activity analogous to the corresponding enzyme of P. venetianus but no diol dehydratase. Our results confirm that corrinoid-influenced PEG degradation analogous to a diol dehydratase reaction is a common strategy among several different strictly anaerobic PEG-degrading bacteria.  相似文献   

12.
Biotransformation of the anabolic steroid dianabol (1) by suspended-cell cultures of the filamentous fungi Cunninghamella elegans and Macrophomina phaseolina was studied. Incubation of 1 with C. elegans yielded five hydroxylated metabolites 26, while M. phaseolina transformed compound 1 into polar metabolites 711. These metabolites were identified as 6β,17β-dihydroxy-17α-methylandrost-1,4-dien-3-one (2), 15α,17β-dihydroxy-17α-methylandrost-1,4-dien-3-one (3), 11α,17β-dihydroxy-17α-methylandrost-1,4-dien-3-one (4), 6β,12β,17β-trihydroxy-17α-methylandrost-1,4-dien-3-one (5), 6β,15α,17β-trihydroxy-17α-methylandrost-1,4-dien-3-one (6), 17β-hydroxy-17α-methylandrost-1,4-dien-3,6-dione (7), 7β,17β,-dihydroxy-17α-methylandrost-1,4-dien-3-one (8), 15β,17β-dihydroxy-17α-methylandrost-1,4-dien-3-one (9), 17β-hydroxy-17α-methylandrost-1,4-dien-3,11-dione (10), and 11β,17β-dihydroxy-17α-methylandrost-1,4-dien-3-one (11). Metabolite 3 was also transformed chemically into diketone 12 and oximes 13, and 14. Compounds 6 and 1214 were identified as new derivatives of dianabol (1). The structures of all transformed products were deduced on the basis of spectral analyses. Compounds 114 were evaluated for β-glucuronidase enzyme inhibitory activity. Compounds 7, 13, and 14 showed a strong inhibition of β-glucuronidase enzyme, with IC50 values between 49.0 and 84.9 μM.  相似文献   

13.
Two models for the active site of liver alcohol dehydrogenase (EC 1.1.1.1) have been proposed. Results of X-ray diffraction studies (B.V. Plapp, H. Eklund, and C.-I. Brändén, 1978, J. Mol. Biol.122, 23–32) on the native enzyme indicate that substrates are directly coordinated to the active site zinc ion, while NMR studies (D. L. Sloan, J. M. Young, and A. S. Mildvan, Biochemistry14, 1998–2008) on the Co II enzyme indicate that substrates are not bound directly to the metal. It was unclear whether the basis for this difference was structural or technical. Therefore, this NMR study has been done with wellcharacterized zinc and cobalt enzymes, and to facilitate comparison with X-ray diffraction data, the substrate analogs chosen were dimethyl sulfoxide and trifluoroethanol. Binding of either analog to the zinc enzyme in the presence of the appropriate cofactor produced unique changes in the T1 and T2 relaxation rates of the 1H and 19F nuclei. Similar results were obtained when cobalt enzyme was used for T1 measurements, but relaxation was more rapid due to the presence of the paramagnetic ion. From these data, the distances between the analog nuclei and the catalytic site cobalt ion were calculated to be 8.9 ± 0.9 and 10.5 ± 1.2 Å for the cobalt enzyme-NADH-dimethyl sulfoxide and the cobalt enzyme-NAD+-F3CCH2OH complexes, respectively. The distances are comparable and the magnitudes indicate that the functional groups are not directly coordinated to the active site cobalt ion. These values are in good agreement with those previously reported by Sloan et al. (1975) for the cobalt enzyme-NADH-isobutyramide complex, and are consistent with their model in which a metal water ligand forms a bridge between the substrate and the metal. Therefore, there must be a structural basis for the differences observed in magnetic resonance versus X-ray diffraction studies.  相似文献   

14.
《Steroids》1996,61(2):50-57
The syntheses of three classes of C-3 perfluoroalkyl substituted steroids are described. They are the 3β-hydroxy-3α-perfluoroalkylandrost-4-en-17-ones (5a-c), 3-perfluoroalkylandrosta-3,5-dien-3-ones (8a-c) and 3β-hydroxy-3α-perfluoroalkylandrost-5-en-17-ones (12a-c). Addition of a series of perfluroalkylorganometallic reagents (RFLi; RF = C2F5, C3F7, or C4F9) to the 3 position of silylated testosterone 2b afforded Δ4 perfluoroalkyl carbinols 3. In Scheme 1, deprotection with HF and oxidation at the C-17 carbon with PCC produced Δ4 ketones 5. In Scheme 2 dehydration of 3 with 1,2-phenylenephosphorochloridite and iodine afforded Δ3,5 dienes 6 which were deprotected and oxidized as above to the C-17 ketones 8. In Scheme 3 isomerization of the double bond of 3 from the C-4 to the C-5 position using the allylic halogenation followed by treatment with lithium aluminum hydride led to the synthesis of the double bond isomer series 12. A new method for dehydration was developed. On average and within experimental error, 3β-hydroxy-3α-perfluoroalkylandrost-5-en-17 ones (12a-c) were better than the 3-perfluoroalkylandrosta-3,5-dien-17-ones (8a-c) and 3β-hydroxy-3α-perfluoroalkylandrost-4-en-17-ones (5a-c) at inhibiting glucose-6-phosphate dehydrogenase.  相似文献   

15.
Three saponins, including two dimeric triterpenoid glucosides possessing an unusual skeleton, ivorenosides A and B, and a monomeric triterpenoid saponin (ivorenoside C), together with the known sericoside, were isolated from the bark of Terminalia ivorensis. Their structures were established on the basis of 1D and 2D NMR data, chemical methods and tandem MS–MS spectrometry as a dimer of β-d-glucopyranosyl-18,19-seco-2α,3β,19,19,24-pentahydroxyolean-12-en-28-oate and β-d-glucopyranosyl-2α,3β,19α,24-tetrahydroxyolean-12-en-28-oate (ivorenoside A, 1), a dimer of β-d-glucopyranosyl-18,19-seco-24-carboxyl-2α,3β,19,19-tetrahydroxyolean-12-en-28-oate and β-d-glucopyranosyl-2α,3β,19α,24-tetrahydroxyolean-12-en-28-oate (ivorenoside B, 2) and β-d-glucopyranosyl-2α,3β,19β,24-tetrahydroxyolean-11-oxo-olean-12-en-28-oate (ivorenoside C, 3). Ivorenosides A and B are the first examples in nature of dimeric triterpenoid saponins with a 18,19-seco E ring of one of the two units. These isolated compounds were evaluated for their antioxidant properties and further for their cytotoxic activity against four human cancer cell lines. Ivorenoside B and C exhibited scavenging activity against DPPH and ABTS+ radicals with IC50 values comparable with that of the standard drug Trolox and ivorenoside A showed antiproliferative activity against MDA-MB-231 and HCT116 human cancer cell lines with IC50 values of 3.96 and 3.43 μM, respectively.  相似文献   

16.
Thirteen (Z)-4-(substituted benzylidene)-3-phenylisoxazol-5(4H)-ones were designed to confirm the geometric effect of the double bond of the β-phenyl-α, β-unsaturated carbonyl scaffold on tyrosinase inhibitory activity. Compounds 1a1m, which all possessed the (Z)-β-phenyl-α, β-unsaturated carbonyl scaffold, were synthesized using a tandem reaction consisting of an isoxazolone ring formation and a Knoevenagel condensation, and three starting materials, ethyl benzoylacetate, hydroxylamine and benzaldehydes. Some of the compounds showed inhibitory activity against mushroom tyrosinase as potent as compounds containing the “(E)”-β-phenyl-α, β-unsaturated carbonyl scaffold. Compounds 1c and 1m showed greater inhibitory activity than kojic acid: IC50?=?32.08?±?2.25?μM for 1c; IC50?=?14.62?±?1.38?μM for 1m; and IC50?=?37.86?±?2.21?μM for kojic acid. A kinetic study indicated that 1m inhibited tyrosinase in a competitive manner and that it probably binds to the enzyme’s active site. In silico docking simulation supported binding of 1m (?7.6?kcal/mol) to the active site of tyrosinase with stronger affinity than kojic acid (?5.7?kcal/mol). Similar results were obtained using cell-based assays, and in B16F10 cells, compound 1m dose-dependently inhibited tyrosinase activity and melanogenesis. These results indicate the anti-melanogenic effect of compound 1m is due to the inhibition of tyrosinase and (Z)-isomer of the β-phenyl-α, β-unsaturated carbonyl scaffold can, like its congener the (E)-isomer, act as an excellent scaffold for tyrosinase inhibition.  相似文献   

17.
Two new steroids, (2β,3β,4α,5α,8β)-4-methylergost-24(28)-ene-2,3,8-triol (1) and (3β,7α)-24-methyl-7-hydroperoxycholest-5,24(28)-diene-3-ol (2), together with 13 known analogues (315) were isolated from the soft coral Sinularia depressa Tixier-Durivault. The structures of the new compounds were elucidated by detailed spectroscopic analysis and comparison with reported data. In the bioassay in vitro, compounds 3a, 4, and 14 exhibited potent PTP1B inhibitory activity, being similar as that of positive control oleanolic acid. Compound 14 also displayed a notable neuroprotective activity against both amyloid-β25–35- and serum deprivation-induced injuries in SH-SY5Y cells while compound 11 showed a considerable antibacterial activity against Staphylococcus aureus. Preliminary structure–activity relationships of these steroids were discussed.  相似文献   

18.
Natural products containing carbon–phosphorus bonds elicit important bioactivity in many organisms. l-Phosphinothricin contains the only known naturally-occurring carbon–phosphorus–carbon bond linkage. In actinomycetes, the cobalamin-dependent radical S-adenosyl-l-methionine (SAM) methyltransferase PhpK catalyzes the formation of the second C–P bond to generate the complete C–P–C linkage in phosphinothricin. Here we use electron paramagnetic resonance and nuclear magnetic resonance spectroscopies to characterize and demonstrate the activity of a cobalamin-dependent radical SAM methyltransferase denoted SD_1168 from Shewanella denitrificans OS217, a marine bacterium that has not been reported to synthesize phosphinothricin. Recombinant, refolded, and reconstituted SD_1168 binds a four-iron, four-sulfur cluster that interacts with SAM and cobalamin. In the presence of SAM, a reductant, and methylcobalamin, SD_1168 surprisingly catalyzes the P-methylation of N-acetyl-demethylphosphinothricin and demethylphosphinothricin to produce N-acetyl-phosphinothricin and phosphinothricin, respectively. In addition, this enzyme is active in the absence of methylcobalamin if the strong reductant titanium (III) citrate and hydroxocobalamin are provided. When incubated with [methyl-13C] cobalamin and titanium citrate, both [methyl-13C] and unlabeled N-acetylphosphinothricin are produced. Our results suggest that SD_1168 catalyzes P-methylation using radical SAM-dependent chemistry with cobalamin as a coenzyme. In light of recent genomic information, the discovery of this P-methyltransferase suggests that S. denitrificans produces a phosphinate natural product.  相似文献   

19.
This study investigated the anti-inflammatory effects of novel pseudotripeptides (GPE 13) as potential candidates to counteract neuroinflammation processes in Alzheimer’s disease.GPE 13 pseudotripeptides are synthetic derivatives of Gly-l-Pro-l-Glu (GPE), the N-terminal tripeptide of IGF-1, obtained through the introduction of isosteres of the amidic bond (aminomethylene unit) to increase the metabolic stability of the native tripeptide. The results showed that all synthetic derivatives possessed higher half-lives (t1/2?>?4?h) than GPE (t1/2?=?30?min) in human plasma and had good water solubility. The biological results demonstrated that GPE 13 had protective properties in several experimental models of treated THP-1 cells. Notably, the novel pseudotripeptides influenced inflammatory cytokine expression (IL-1β, IL-18, and TNF-α) in Aβ25–35-, PMA-, and LPS-treated THP-1 cells. In PMA-differentiated THP-1 macrophages, both GPE 1 and GPE 3 reduced the expression levels of all selected cyto-chemokines, even though GPE 3 showed the best neuroprotective properties.  相似文献   

20.
Little is known about the acquisition of cobalamin by the mammary gland and its secretion into milk. Human milk and plasma contain at least two types of cobalamin binding proteins: transcobalamin II (TC) and haptocorrin (HC). In plasma, TC is responsible for the transport of cobalamin to tissues and cells; however, cobalamin in milk is present exclusively bound to HC. We show that human mammary epithelial cells (HMEC) exhibit high affinity for TC; Scatchard analysis revealed a single class of binding sites for the TC–[57Co]cyanocobalamin complex with a dissociation constant (Kd) of 4.9×10−11 M. Uptake of the TC–[57Co]cyanocobalamin complex at 37°C was saturable by 24 h. Binding of free [57Co]cyanocobalamin to HMEC was not saturable and very limited binding of the HC–[57Co]cyanocobalamin complex was observed. Expression of the haptocorrin gene by HMEC was confirmed by Northern blot and PCR analysis. Thus, a specific cell surface receptor for the TC–cobalamin complex exists in the mammary gland and once cobalamin is internalized, it may be transferred to HC and subsequently secreted into milk as a HC–cobalamin complex.  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