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.
The previously reported analog of pregnenolone having a 3,4-dihydro-2H-pyran attached via a Cz.sbnd;C bond to the C-20 position (1), stereoselectively reacts with m-chloroperoxybenzoic acid in methanol at -5 degrees C. Acid-catalyzed hydrolysis of the isolated intermediates gives good yields of mostly a new 27-norcholesterol analog: (20R,23R)-3,20,23,26-tetrahydroxy-27-norcholest-5-en-22-one-3-acetate (2a, and a smaller amount of its 23S enantiomer 2b). Three different conditions of epoxidation and methanolysis followed by acid-catalyzed hydrolysis typically produce approximately 2:1 ratios of the 23R:23S diastereoisomers with a C-23 hydroxy group at the new asymmetric center. Bromine also reacts stereoselectively with (20R)-3,20-dihydroxy-(3',4'-dihydro-2'H-pyranyl)-5-pregnene (4) giving mostly (20R,23R)-23-bromo-3,20,26-trihydroxy-27-norcholest-5-en-22-one (7a). Thus both major steroidal products 2a and 7a have the same C-23R configuration. Assignment of molecular structures and the absolute configurations to 1 and 2a were based on elemental analysis, mass spectra, nuclear magnetic resonance, FTIR infrared spectroscopic analysis and X-ray crystallography. Mechanisms are discussed for stereochemical selectivity during epoxidation and bromination of the 3,4-dihydro-2H-pyranyl ring in 1 and 4.  相似文献   

2.
D Grobelny  R E Galardy 《Biochemistry》1985,24(22):6145-6152
The collagenase from Clostridium histolyticum is a mixture of several collagenases, all of which are zinc metalloproteases. This enzyme catalyzes the cleavage of the X-Gly peptide bond in the repeating sequence of collagen: -Gly-Pro-X-Gly-Pro-X-. Thus the S3, S2, and S1 subsites on the enzyme appear to be occupied by the sequence -Gly-Pro-X- and the S1', S2', and S3' subsites also by -Gly-Pro-X-. Short peptides up to and including N alpha-acyltetrapeptides containing the repeat sequence do not detectably inhibit the enzyme (IC50 greater than 10 mM). However, peptide aldehydes of the form aminoacyl-X-glycinal, presumably occupying the S1, S2, ..., Sn subsites, are inhibitors. The most potent of these was Pro6-Gly-Pro-glycinal, with an IC50 of 340 +/- 70 microM. The single peptide aldehyde investigated, which could occupy the S1' and S2' subsites, 4-oxobutanoyl-L-proline, did not inhibit collagenase (IC50 greater than 20 mM). The peptide ketone 5-benzamido-4-oxo-6-phenylhexanoyl-Pro-Ala (XXV), which could occupy the S1-S3' subsites, inhibits collagenase with an IC50 of 120 +/- 50 microM, over 80-fold more potently than its parent peptide analogue benzoyl-Phe-Gly-Pro-Ala (XXIII). The alcohol analogue of XXV, 5-benzamido-4-hydroxy-6-phenylhexanoyl-Pro-Ala (XXVI), is over 60-fold less potent with an IC50 of 8 +/- 2mM. Extending the peptide ketone XXV to occupy the S2-S3' subsites gave 5-(N alpha-carbobenzoxy-L-prolinamido)-4-oxo-6-phenylhexanoyl-Pro -Ala (XXVII). Surprisingly, XXVII had an IC50 of only 5.2 +/- 2 mM.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Several N-acetyllactosamine (LacNAc) derivatives were tested as acceptors for alpha 1,3-L-fucosyltransferase present in human ovarian cancer sera and ovarian tumor. The enzyme of the soluble fraction of tumor was purified to apparent homogeneity by chromatography on bovine IgG glycopeptide-Sepharose followed by Sephacryl S-200 (M(r) < 67,000). As compared with 2'-methyl LacNAc, 3'-sulfo LacNAc was about 5-fold more sensitive in measuring alpha 1,3-fucosyltransferase in sera (Km, 3'-sulfo LacNAc, 0.12 mM; 2'-methyl LacNAc, 6.67 mM). When ovarian cancer serum was the enzyme source, either the sulfate group or a sialyl moiety at C-3' of LacNAc enhanced the acceptor ability (341 and 242%, respectively), whereas the sulfate group at C-2' or C-6' reduced the activity (22-36%); sulfate at C-6 or fucose at C-2' increased the activity (172 and 253%). The beta-benzylation of the reducing end, in general, increased the activity 2-3-fold. The enzyme of the soluble fraction of tumor exhibited more activity toward 3'-sulfo LacNAc (447%), 2'-fucosyl-LacNAc (436%), and 6-sulfo LacNAc (272%). Very low activity was observed with 3'-sialyl LacNAc (12.4%), 2'-sulfo LacNAc (33%), and 6'-sulfo LacNAc (5%); Fuc alpha 1,2Gal beta 1,3GlcNAc beta-O-p-nitrophenyl (166%), 2-methyl Gal beta 1,3GlcNAc beta-O-benzyl (204%), and 3-sulfo Gal beta 1,3GlcNAc (415%) also acted as acceptors, indicating the coexistence of alpha 1,3- and alpha 1,4-fucosyltransferase. The tumor particulate enzyme behaved entirely different, exhibiting low activity with 3'-sulfo LacNAc (39%) and 2'-fucosyl-LacNAc (148%); 3'-sialyl, 6'-sulfo, 6-sulfo, or 2'-sulfo LacNAc were 3, 43, 53, and 10% active, respectively. Thus, the ovarian cancer serum alpha 1,3-fucosyltransferase acts equally well on H-type 2,3'-sialyl LacNAc and 3'-sulfo LacNAc, but not on H-type 1. The enzyme of soluble tumor fraction acts on H-type 2,3'-sulfo LacNAc as well as H-type 1 but poorly on 3'-sialyl LacNAc. The tumor particulate enzyme acts on H-type 2 but poorly on 3'-sulfo or 3'-sialyl LacNAc and is inactive with H-type 1. When normal serum was examined with synthetic acceptors, > 80% activity was found as alpha 1,2-fucosyltransferase and the rest as alpha 1,3-fucosyltransferase. A screening of 21 ovarian cancer and 3 normal sera (3'-sulfo LacNAc as acceptor) showed 17-572% increase (average increase, 188%) of alpha 1,3-fucosyltransferase activity in cancer.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
(+)-Pisatin, the major phytoalexin of pea (Pisum sativum L.), is believed to be synthesized via two chiral intermediates, (-)-7,2'-dihydroxy-4',5'-methylenedioxyisoflavanone [(-)-sophorol] and (-)-7,2'-dihydroxy-4',5'-methylenedioxyisoflavanol [(-)-DMDI]; both have an opposite C-3 absolute configuration to that found at C-6a in (+)-pisatin. The expression of isoflavone reductase (IFR), which converts 7,2'-dihydroxy-4',5'-methylenedioxyisoflavone (DMD) to (-)-sophorol, sophorol reductase (SOR), which converts (-)-sophorol to (-)-DMDI, and hydroxymaackiain-3-O-methyltransferase (HMM), believed to be the last step of (+)-pisatin biosynthesis, were inactivated by RNA-mediated genetic interference (RNAi) in pea hairy roots. Some hairy root lines containing RNAi constructs of IFR and SOR accumulated DMD or (-)-sophorol, respectively, and were deficient in (+)-pisatin biosynthesis supporting the involvement of chiral intermediates with a configuration opposite to that found in (+)-pisatin in the biosynthesis of (+)-pisatin. Pea proteins also converted (-)-DMDI to an achiral isoflavene suggesting that an isoflavene might be the intermediate through which the configuration is changed to that found in (+)-pisatin. Hairy roots containing RNAi constructs of HMM also were deficient in (+)-pisatin biosynthesis, but did not accumulate (+)-6a-hydroxymaackiain, the proposed precursor to (+)-pisatin. Instead, 2,7,4'-trihydroxyisoflavanone (TIF), daidzein, isoformononetin, and liquiritigenin accumulated. HMM has a high amino acid similarity to hydroxyisoflavanone-4'-O-methyltransferase (HI4'OMT), an enzyme that methylates TIF, an early intermediate in the isoflavonoid pathway. The accumulation of these four compounds is consistent with the blockage of the synthesis of (+)-pisatin at the HI4'OMT catalyzed step resulting in the accumulation of liquiritigenin and TIF and the diversion of the pathway to produce daidzein and isoformononetin, compounds not normally made by pea. Previous results have identified two highly similar "HMMs" in pea. The current results suggest that both of these O-methyltransferases are involved in (+)-pisatin biosynthesis and that one functions early in the pathway as HI4'OMT and the second acts at the terminal step of the pathway.  相似文献   

5.
13C and 2H NMR spectroscopy has been employed to probe the biosynthesis of vitamin B6 in Escherichia coli. The 13C NMR spectrum of a sample of pyridoxol derived biosynthetically from D-[1,2,3,4,5,6-13C6]glucose shows that the bonds, C(2)-C(3) and C(4)-C(5), of the pyridine nucleus are the only two carbon-carbon bonds of pyridoxol which are generated de novo in the course of its biosynthesis from glucose. It follows that the pyridoxol skeleton is generated from two intact triose units and a triose-derived two-carbon unit, all of which are supplied by glucose. From the 2H NMR spectra of samples of pyridoxol derived from (R)-[1,1-2H2]glycerol and (S)-[1,1-2H2]glycerol, respectively, it can be deduced that the rehydroxymethyl group of glycerol enters C-2', C-4', and C-5' of the pyridoxol skeleton. It follows that each of the three fragments is derived from glycerol in stereo-specific fashion. These results answer questions concerning the regiochemistry and the stereochemistry of pyridoxol biosynthesis.  相似文献   

6.
Twenty new 5-(hydroxyalkyl)-2-cyclopentenone derivatives (hygrophorones) could be isolated from Hygrophorus latitabundus, H. olivaceoalbus, H. persoonii, and H. pustulatus. Their fungicidal activity was exemplarily tested. The hygrophorones have structural similarities to the antibiotic pentenomycin. Chemically, hygrophorones are 2-cyclopentenones with hydroxy or acetoxy substituents at C-4 and/or C-5. An odd-numbered 1' oxidized alkyl chain (C(11), C(13), C(15), or C(17)) is attached at C-5. In addition, from H. persoonii the new gamma-butyrolactone derivative [5-(E)-2-hydroxytetradexylidene-5H-furan-2-one] could be isolated. Some hygrophorones are responsible for the color reaction of the stipes of these fungi upon treatment with potassium hydroxide solution. Structural elucidations are based on 1D ((1)H, (13)C) and 2D (COSY, NOESY, HSQC, HMBC) NMR spectroscopic analyses as well as HR-FT-ICR-MS investigations.  相似文献   

7.
D E Stevenson  M Akhtar  D Gani 《Biochemistry》1990,29(33):7631-7647
L-Methionine decarboxylase from the male fern Dryopteris filix-mas has been purified 256-fold from acetone powder extracts to very near homogeneity. The enzyme is membrane-associated and requires detergent for solubilization during the initial extraction. The enzyme is a homodimer of subunit Mr 57,000 and shows a pH optimum at approximately 5.0 with 20 mM (2S)-methionine as substrate. The specific activity, kcat, for methionine is approximately 50 mol s(-1) (mol of active site)(-1) at pH 4.5 and below. A wide range of straight- and branched-chain (2S)-alkylamino acids are substrates for the enzyme. The values for the rate of decarboxylation, Vmax, and for the apparent Michaelis constant, Km, however, vary with structure and with the chirality at C-3. The pH dependence of V and V/K has been examined for three substrates: (2S)-methionine, valine, and leucine. Pyridoxal 5'-phosphate (PLP) is required for activity, and in the absence of excess PLP, the activity of the enzyme in incubations reduced with respect to time. The addition of PLP fully restores the activity, indicating that an abortive decarboxylation-transamination accompanies the normal decarboxylation reaction. The occurrence of the abortive reaction was confirmed by showing that [35S]methionine is converted to labeled 3-(methylthio)propionaldehyde while [4'-3H]PLP is converted to labeled pyridoxamine 5'-phosphate (PMP). The decarboxylation of (2S)-methionine gave 3-(methylthio)-1-aminopropane. Preparation of the N-camphanamide derivative of the amine allowed the C-1 methylene protons to be distinguished by 1H NMR spectroscopy. Synthetic samples of the camphanamide were prepared in which each of the C-1 methylene protons was replaced by deuterium. When (2S)-methionine and the C-2 deuteriated isotopomer were incubated with the enzyme in deuterium oxide and protium oxide, respectively, and the products were converted to their camphanamide derivatives and analyzed by 1H NMR spectroscopy, it was evident that decarboxylation occurred with retention of configuration at C-2. When the decarboxylation of six other substrates was studied, examination of the N-camphanamide derivatives of the amines indicated that decarboxylation occurred stereospecifically and, by analogy, with retention of configuration at C-2. When tritiated pyridoxal phosphate was incubated with the enzyme, tritiated pyridoxamine phosphate was formed. Analysis of the chirality of the methylene group at C-4' indicated that, during abortive transamination, protonation occurred from the 4'-si face of the coenzyme, the same stereochemical result as that obtained for several bona fide transaminase enzymes.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

8.
Enzymatic activity which hydrolyzes diadenosine 5',5"'-P1,P4-tetraphosphate (Ap4A) yielding ADP has been identified in extracts of eubacteria, Escherichia coli and Acidaminococcus fermentans, and of a highly thermophilic archaebacterium, Pyrodictum occultum. Specific Ap4A (symmetric) pyrophosphohydrolase from Escherichia coli K12 has been purified almost 400-fold. The preparation was free of phosphatase, ATPase, phosphodiesterase, AMP-nucleosidase, and adenylate kinase. The Ap4A pyrophosphohydrolase molecular weight estimated by gel filtration is 27,000 +/- 1,000. Activity maximum is at pH 8.3. The Km value computed for Ap4A is 25 +/- 3 microM. The sulfhydryl group(s) is essential for enzyme activity. Metal chelators, EDTA, and o-phenanthroline, inhibit Ap4A hydrolysis; I0.5 values are 3 and 50 microM, respectively. Co2+ is a strong stimulator with an almost 100-fold increase in rate of Ap4A hydrolysis and a plateau in the range of 100-500 microM Co2+, when compared with the nonstimulated hydrolysis. Other transition metal ions, Mn2+, Cd2+, and Ni2+, stimulate by factors of 8, 3.5, and 3.5, respectively, with optimal concentrations in the range 200-500, 2-5, and 4-8 microM, respectively. Zn2+, Cu2+, and Fe2+, up to 30 microM, are without effect and they inhibit at higher concentrations. Mg2+ or Ca2+, in the absence of other divalent metal ions, are weak stimulators (1.5-fold stimulation occurs at 1-2 mM concentration), but act synergistically with Co2+ at its suboptimal concentrations. Stimulation in the presence of 10 microM Co2+ and either 1 mM MgCl2 or CaCl2 increases up to 75-fold. The same degree of synergy is found at 10 microM Co2+ and either 2-5 mM spermidine or 0.5-1.5 mM spermine. Besides Ap4A, bacterial Ap4A pyrophosphohydrolase hydrolyzes effectively Ap5A and Gp4G, and, to some extent, p4A, Ap6A, and Ap3A yielding in each case corresponding nucleoside diphosphate as one of the products.  相似文献   

9.
Equilibrium dialysis and protection from heat inactivation and proteolysis show that initiation factor 2 (IF-2) interacts not only with GTP but also with GDP and that its conformation is changed upon binding of either nucleotide. The apparent Ka (at 25 degrees C) for the IF-2 X GDP and IF-2 X GTP complexes was 8.0 X 10(4) and 7.0 X 10(3) M(-1), respectively. The lower affinity for GTP is associated with a more negative delta S0. The interaction, monitored by 1HNMR spectroscopy, is characterized by fast exchange and results in line broadening and downfield shift of the purine C-8 and ribose C-1' protons of GTP as well as of the beta, gamma-methylene protons of (beta-gamma-methylene)guanosine 5'-triphosphate. The interaction of guanosine nucleotides with IF-2 requires an H bond donor (or acceptor) group at position C-2 of the purine and involves the beta- and/or gamma-phosphate of the nucleotide while the ribose 2'-OH group or the integrity of the furan ring are less critical. IF-2 binds to ribosomal particles with decreasing affinity: 30 S greater than 70 S greater than 50 S. GTP and GDP have no effect on the binding to 70 S. GTP stimulates the binding to the 30 S and depresses somewhat the binding to the 50 S subunits; GDP has the opposite effect. These results seem to rule out that the release of IF 2 from 70 S is due to a "GDP-conformation" of the factor incompatible with its permanence on the ribosome. The rate and the extent of 30 S initiation complex formation are approximately 2-fold higher with IF-2 X GTP than with IF-2 alone. At low concentrations of IF-2 and 30 S subunits, GDP inhibits this reaction, acting as a strong competitive inhibitor of GTP (Ki = 1.25 X 10(-5)m) and preventing IF-2 from binding to the ribosomal subunit.  相似文献   

10.
Some effects of salts on the adenylate cyclase of partially purified plasma membranes from rat liver have been studied. Under conditions where cyclic adenosine 3':5'-monophosphate formation was linear with respect to time and protein concentration, the enzyme was stimulated 3- to 6-fold by 10 mM NaF, 10- to 30-fold by 1 muM glucagon, 4- to 5-fold by 0.1 mM 5'-guanylylimidodiphosphate, and in the presence of 3 muM GTP, 2-fold by 10 mug/ml of prostaglandin E1. Various salts were found to stimulate basal activity slightly, but enhanced the response to NaF 3- to 4-fold, to glucagon 1.5- to 2-fold, to 5'-guanylylimidodiphosphate 2- to 3-fold, and to prostaglandin E1 1.5-fold. This enhancement was observed at maximally effective concentrations of each of the respective activators. Of the salts tested, NaN3 and the Na- or K-halides were most effective. Their action appeared to be due to the respective anions. Stimulation was detectable with 1.5 mM NaN3 or 3 mM NaCl and was maximal with 30 mM NaN3 or 60 mM NaCl. The stimulatory effect of NaN3 was not due to ATP-sparing, nor to an altered cyclic adenosine 3':5'-monophosphate recovery. It was independent of the chromatography and assay methods used, and was therefore not due to procedural artifact. Fluoride-stimulated cyclase activity was enhanced by salts to a greater degree than were 5'-guanylylimidodiphosphate-, glucagon-, or (prostaglandin E1 + GTP)-stimulated activities. The effects of NaN3 were not the result of significant changes in the enzyme's responses to GTP, which increased basal and glucagon-stimulated activities but inhibited F--stimulated activity. The effects of NaN3 were greater when cyclase was assayed with Mn2+ than with Mg2+. The facilitatory effect of NaN3 or NaCl on fluoride-stimulated adenylate cyclase activity was partially reversible as was the stimulatory effect of fluoride in the presence of NaN3. Enhancement of hormonal stimulation by NaN3 was also demonstrable with cardiac and adipose tissue adenylate cyclase. However, NaN3 did not stimulate detergent-dispersed adenylate cyclases from either liver plasma membranes or brain. The data suggest that stimulation of adenylate cyclase by salts may require the added presence of other stimulatory agents and an intact membrane structure.  相似文献   

11.
The biosynthetic pathways to abscisic acid (ABA) were investigated by feeding [1-(13)C]-D-glucose to cuttings from young tulip tree shoots and to two ABA-producing phytopathogenic fungi. 13C-NMR spectra of the ABA samples isolated showed that the carbons at 1, 5, 6, 4', 7' and 9' of ABA from the tulip tree were labeled with 13C, while the carbons at 2, 4, 6, 1', 3', 5', 7', 8' and 9' of ABA from the fungi were labeled with 13C. The former corresponds to C-1 and -5 of isopentenyl pyrophosphate, and the latter to C-2, -4 and -5 of isopentenyl pyrophosphate. This finding reveals that ABA was biosynthesized by the non-mevalonate pathway in the plant, and by the mevalonate pathway in the fungi. 13C-Labeled beta-carotene from the tulip tree showed that the positions of the labeled carbons were the same as those of ABA, being consistent with the biosynthesis of ABA via carotenoids. Lipiferolide of the tulip tree was also biosynthesized by the non-mevalonate pathway.  相似文献   

12.
Hyperforin and adhyperforin contribute to the antidepressant effects of Hypericum perforatum. The involvement of branched-chain amino acids in the biosynthesis of hyperforin and adhyperforin was demonstrated in H. perforatum shoot cultures. L-[U-(13)C(5)]Valine and L-[U-(13)C(6)]isoleucine, upon administration to the shoot cultures, were incorporated into acyl side chain of hyperforin and adhyperforin, respectively. Feeding the shoot cultures with unlabelled L-isoleucine at a concentration of 2mM induced a 3.7-fold increase in the production of adhyperforin. The addition of 3mM L-threonine, a precursor of isoleucine, stimulated a 2.0-fold increase in the accumulation of adhyperforin. The administration of L-valine at concentrations of 0-5mM had no stimulating effect on the hyperforin production in H. perforatum shoot cultures.  相似文献   

13.
The activity of enzyme I (EI), the first protein in the bacterial PEP:sugar phosphotransferase system, is regulated by a monomer-dimer equilibrium where a Mg(2+)-dependent autophosphorylation by PEP requires the homodimer. Using inactive EI(H189A), in which alanine is substituted for the active-site His189, substrate-binding effects can be separated from those of phosphorylation. Whereas 1 mM PEP (with 2 mM Mg(2+)) strongly promotes dimerization of EI(H189A) at pH 7.5 and 20 degrees C, 5 mM pyruvate (with 2 mM Mg(2+)) has the opposite effect. A correlation between the coupling of N- and C-terminal domain unfolding, measured by differential scanning calorimetry, and the dimerization constant for EI, determined by sedimentation equilibrium, is observed. That is, when the coupling between N- and C-terminal domain unfolding produced by 0.2 or 1.0 mM PEP and 2 mM Mg(2+) is inhibited by 5 mM pyruvate, the dimerization constant for EI(H189A) decreases from > 10(8) to < 5 x 10(5) or 3 x 10(7) M(-1), respectively. Incubation of the wild-type, dephospho-enzyme I with the transition-state analog phosphonopyruvate and 2 mM Mg(2+) also increases domain coupling and the dimerization constant approximately 42-fold. With 2 mM Mg(2+) at 15-25 degrees C and pH 7.5, PEP has been found to bind to one site/monomer of EI(H189A) with K(A)' approximately 10(6) M(-1) (deltaG' = -8.05 +/- 0.05 kcal/mole and deltaH = +3.9 kcal/mole at 20 degrees C); deltaC(p) = -0.33 kcal K(-1) mole(-1). The binding of PEP to EI(H189A) is synergistic with that of Mg(2+). Thus, physiological concentrations of PEP and Mg(2+) increase, whereas pyruvate and Mg(2+) decrease the amount of dimeric, active, dephospho-enzyme I.  相似文献   

14.
Kinetics of the hydrolysis of a P(1)-(7-methylguanosinyl-5') P(3)-(guanosinyl-5') triphosphate (m(7)GpppG), P(1)-(7-methylguanosinyl-5') P(4)- (guanosinyl-5') tetraphosphate (m(7)GppppG), diadenosine-5',5'-P(1),P(3)-triphosphate (ApppA), and diadenosine-5',5'-P(1),P(4)-tetraphosphate (AppppA) promoted by Cu(2+) or Zn(2+) has been investigated. Time-dependent products distributions at various metal ion concentrations have been determined by CZE and HPLC-RP. The results show that in acidic conditions, in the presence of metal ion, the predominant hydrolytic reaction is the cleavage of 5',5'-oligophosphate bridge. The 5',5'-oligophosphate bridge of the dinucleotides studied is hydrolyzed by Cu(2+) more efficiently than by Zn(2+). At the catalyst concentration of 2 mM the cleavage of the 5',5'-triphosphate bridge of m(7)GpppG was ~3.6 times faster, and that of the tetraphosphate bridge of m(7)GppppG ~2.3-fold faster in the presence of Cu(2+) compared to the Zn(2+) ion, applied as catalysts. Dependence of the rates of hydrolysis on the catalyst concentration was in some instances not linear, interpreted as evidence for participation of more than one metal ion in the transition complex.  相似文献   

15.
Three new mycophenolic acid derivatives, penicacids A-C (1-3), together with two known analogues, mycophenolic acid (MPA, 4) and 4'-hydroxy-MPA (5), were isolated from a fungus Penicillium sp. SOF07 derived from a South China Sea marine sediment. The structures of compounds 1-3 were elucidated on the basis of MS and NMR ((1)H, (13)C, HSQC and HMBC) data analyses and comparisons with the known compounds. Structure-activity relationship studies of compounds 1-5 focused on inosine-monophosphate dehydrogenase inhibition revealed that hydroxylation at C-4', methylation at C-7-OH, dual hydroxylation at C-2'/C-3' double bond of MPA diminished bioactivity whereas glucosyl hydroxylation at C-4' correlated to bioactivity comparable to that observed for MPA.  相似文献   

16.
The small ribosome subunit of Escherichia coli contains 10 base-methylated sites distributed in important functional regions. At present, seven enzymes responsible for methylation of eight bases are known, but most of them have not been well characterized. One of these enzymes, RsmE, was recently identified and shown to specifically methylate U1498. Here we describe the enzymatic properties and substrate specificity of RsmE. The enzyme forms dimers in solution and is most active in the presence of 10-15 mM Mg(2+) and 100 mM NH(4)Cl at pH 7-9; however, in the presence of spermidine, Mg(2+) is not required for activity. While small ribosome subunits obtained from an RsmE deletion strain can be methylated by purified RsmE, neither 70S ribosomes nor 50S subunits are active. Likewise, 16S rRNA obtained from the mutant strain, synthetic 16S rRNA, and 3' minor domain RNA are all very poor or inactive as substrates. 30S particles partially depleted of proteins by treatment with high concentrations of LiCl or in vitro reconstituted intermediate particles also show little or no methyl acceptor activity. Based on these data, we conclude that RsmE requires a highly structured ribonucleoprotein particle as a substrate for methylation, and that methylation events in the 3' minor domain of 16S rRNA probably occur late during 30S ribosome assembly.  相似文献   

17.
The bacterial PEP:sugar phosphotransferase system couples the phosphorylation and translocation of specific sugars across the membrane. The activity of the first protein in this pathway, enzyme I (EI), is regulated by a monomer-dimer equilibrium where a Mg(2+)-dependent autophosphorylation by PEP requires the dimer. Dimerization constants for dephospho- and phospho-EI and inactive mutants EI(H189E) and EI(H189A) (in which Glu or Ala is substituted for the active site His189) have been measured under a variety of conditions by sedimentation equilibrium at pH 7.5 and 4 and 20 degrees C. Concurrently, thermal unfolding of these forms of EI has been monitored by differential scanning calorimetry and by changes in the intrinsic tryptophanyl residue fluorescence. Phosphorylated EI and EI(H189E) have 10-fold increased dimerization constants [ approximately 2 x 10(6) (M monomer)(-1)] compared to those of dephospho-EI and EI(H189A) at 20 degrees C. Dimerization is strongly promoted by 1 mM PEP with 2 mM MgCl(2) [K(A)' > or = 10(8) M(-1) at 4 or 20 degrees C], as demonstrated with EI(H189A) which cannot undergo autophosphorylation. Together, 1 mM PEP and 2 mM Mg(2+) also markedly stabilize and couple the unfolding of C- and N-terminal domains of EI(H189A), increasing the transition temperature (T(m)) for unfolding the C-terminal domain by approximately 18 degrees C and that for the N-terminal domain by approximately 9 degrees C to T(max) congruent with 63 degrees C, giving a value of K(D)' congruent with 3 microM PEP at 45 degrees C. PEP alone also promotes the dimerization of EI(H189A) but only increases T(m) approximately 5 degrees C for C-terminal domain unfolding without affecting N-terminal domain unfolding, giving an estimated value of K(D)' congruent with 0.2 mM for PEP dissociation in the absence of Mg(2+) at 45 degrees C. In contrast, the dimerization constant of phospho-EI at 20 degrees C is the same in the absence and presence of 5 mM PEP and 2 mM MgCl(2). Thus, the separation of substrate binding effects from those of phosphorylation by studies with the inactive EI(H189A) has shown that intracellular concentrations of PEP and Mg(2+) are important determinants of both the conformational stability and dimerization of dephospho-EI.  相似文献   

18.
Methyl 2,3-O-isopropylidene-alpha-D-mannofuranosidurononitrile [alternative name: methyl (5R)-5-C-cyano-2,3-O-isopropylidene-alpha-D-lyxofuranoside] (2), methyl 2,3-O-isopropylidene-alpha-D-mannofuranosiduronamide [methyl (5S)-5-C-carbamoyl-2,3-O-isopropylidene-alpha-D-lyxofuranoside; methyl (5S)-2,3-O-isopropylidene-alpha-D-lyxo-hexofuranosiduronamide] (3), methyl 2,3-O-isopropylidene-alpha-D-mannofuranosiduronic acid [methyl (5S)-2,3-O-isopropylidene-alpha-D-lyxo-hexofuranosiduronic acid] (4), methyl 5-deoxy-2,3-O-isopropylidene-5-ureido-beta-L-gulofuranosiduronamide [methyl (5R)-5-deoxy-2,3-O-isopropylidene-5-ureido-alpha-D-lyxo-hexofuranosiduronamide (5), and (4S,5S,6R)-5,6-dihydro-6-hydroxy-4,5-isopropylidenedioxy-4H-pyrido[2,1-e]imidazolidine-2',4'-dione [IUPAC name: (3aS,4R,8aS)-4-hydroxy-2,2-dimethyl-3a,8a-dihydro-4H-1,3-dioxa-4a,6-diaza-s-indacene-5,7-dione] (6), instead of the expected hydantoin derivative, were obtained from the Bucherer-Bergs reaction of methyl 2,3-O-isopropylidene-alpha-D-lyxo-pentodialdo-1,4-furanoside (1). The structure of 6 was deduced from NMR and mass spectral data and confirmed by X-ray crystallography. The configuration at C-5 in 2-5 was confirmed by establishing the 5S configuration of 3 by X-ray crystallography. Conformations of the six- and five-membered rings in 3 and 6 are also discussed.  相似文献   

19.
A series of carotenoids with a 5,6-dihydro-5,6-dihydroxy-beta-end group, named ipomoeaxanthins A (1), B (2), C1 (3) and C2 (4) were isolated from the flesh of yellow sweet potato "Benimasari", Ipomoea batatas Lam. Their structures were determined to be (5R,6S,3'R)-5,6-dihydro-beta,beta-carotene-5,6,3'-triol (1), (5R,6S,5'R,6'S)-5,6,5',6'-tetrahydro-beta,beta-carotene-5,6,5'6'-tetrol (2), (5R,6S,5'R,8'R)-5',8'-epoxy-5,6,5',8'-tetrahydro-beta,beta-carotene-5,6-diol (3), and (5R,6S,5'R,8'S)-5',8'-epoxy-5,6,5',8'-tetrahydro-beta,beta-carotene-5,6-diol (4) by UV-Vis, NMR, MS and CD data.  相似文献   

20.
Three 5-hydroxy-seco-carotenoids were isolated from seeds of Pittosporum tobira. These structures were determined to be (3S,3'S,5'?)-3,3'-di(tetradecanoyloxy)-5'-hydroxy-5,6,5',6'-diseco-beta,beta-carotene-5,6,6'-trione (1), (3S,5?,3'S,5'R,6'S,9'Z)-3-tetradecanoyloxy-5',6'-epoxy-5,3'-dihydroxy-5',6'-dihydro-5,6-seco-beta,beta-caroten-6-one (2), and (3S,5?,3'S,5'R,6'R)-3-tetradecanoyloxy-5,3',5'-trihydroxy-6',7'-didehydro-5',6'-dihydro-5,6-seco-beta,beta-caroten-6-one (3) based on analysis of UV-vis, IR, FAB MS, and NMR spectroscopic data.  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