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.
5-Thio-l-fucopyranose tetraacetate was synthesized in 11 steps from or d-arabinose diethyl dithioacetal by one-carbon elongation at C-5. Highly diastereo-selective addition of MeLi in ether to a derivative was achieved to give the corresponding 6-deoxy-β-d-altrofuranose isomer in good yield. A sulfur atom was introduced at C-5 of 6-deoxy-d-altrofuranose derivatives via substitution of a 5-tosylate with KSAc in HMPA with inversion of configuration, giving 5-thio-l-fucopyranose. A derivative was also prepared from 6-deoxy-β-d-altrofuranose derivatives. 5-Thio-d-arabinopyranose tetraacetate, the 5-demethyl analog of 5-thio-l-fucose, was also synthesized from in 5 steps. 5-Thio-d-arabinose showed weak inhibitory activity against α-l-fucosidase from bovine kidney (Ki = 0.77 mM).  相似文献   

2.
Using broadband dielectric spectroscopy, we have studied different types of relaxation processes, namely, primary (α), secondary (β), and another sub-Tg process called γ-process, in the supercooled state of d-lyxose, over a wide frequency (10-2–) and temperature range (120–340 K). In addition, the same sample was analyzed by differential scanning calorimeter. The temperature dependence of the relaxation times as well as the dielectric strength of different processes has been critically examined. It has been observed that the slower secondary relaxation (designated as β-) process shifts to lower frequencies with increasing applied pressure, but not the faster one. This pressure dependence indicates that the observed slower secondary relaxation (β-) is Johari–Goldstein relaxation process and faster one (γ-process) is probably the rotation of hydroxymethyl (–CH2OH) side group attached to the sugar ring, that is, of intramolecular origin.  相似文献   

3.
Three routes were investigated for the conversion of d-glucose into the title compound. In the first approach, reduction of the 5,6-thürane ring of 5,6-dideoxy-5,6-epithio-1,2-O-isopropylidene α-d-glucofuranose (17) as well as that of its 3-O-allyl derivative (13) with lithium aluminium hydride was investigated; 17 afforded the corresponding 6-deoxy derivative besides di-, tri-, and poly-mers, whereas only polymers were formed from 13. In the second approach, the oxirane ring of was reduced by sodium borohydride and the resulting 6-deoxy derivative was converted into the 5-thiobenzoate; the corresponding hex-4-enofuranose was formed as a byproduct. In the third approach partial mesylation of methyl 5-thio-α-d-glucopyranoside was attempted, but the 6-mesylate 27 could be isolated only in modest yield (28%) together with rearranged 2,5-thioanhydromannofuranoside derivatives. The mechanism of this rearrangement is discussed in detail. The 6-mesylate 27 was converted via the 6-iodo derivative into the title compound.  相似文献   

4.
The catalytic activity of a membrane-bound enzyme, d-fructose dehydrogenase (FDH), at the polarized oil/water (O/W) interface was studied. Multisweep cyclic voltammetry and ac voltammetry were carried out to show the irreversible adsorption of FDH at the interface. Using the thusly prepared FDH-adsorbed O/W interface, clear steady-state catalytic current was observed in amperometry and cyclic voltammetry, where 1,1′-dimethylferrocenium ion (DiMFc+, electron acceptor) and d-fructose (substrate) were added to the O and W phases, respectively. The observed catalytic current was then analyzed by using two mechanisms. In mechanism (A), the heme c site of FDH, where DiMFc+ is reduced, was assumed to be located in the O-phase side of the interface. The intramolecular electron transfer in FDH should be affected by the Galvani potential difference of the interface (). However, the theoretical equations derived for the catalytic current could not reproduce the experimental data. In mechanism (B), the heme c site was assumed to be in the W-phase side. In this case, should affect the interfacial distribution of DiMFc+. This mechanism could reproduce well the observed potential dependence of the catalytic current.  相似文献   

5.
An immobilized d-hydantoinase was characterized and employed to produce n-carbamoyl-d-p-hydroxyphenylglycine (CpHPG) in a repeated batch process. The Vmax and Km of the immobilized d-hydantoinase at 50°C were 6.28 mm min−1 g−1 biocatalyst and 71.6 mm, respectively. The product CpHPG did not inhibit the activity of d-hydantoinase. Optimal reaction temperature was 60°C. A decrease in activity of immobilized d-hydantoinase due to thermal inactivation could be described as first-order decay; the deactivation energy was 23.97Kcal mol−1. Under process conditions (50°C, 10% w/v substrate, and pH 8.5), the half-life of the immobilized d-hydantoinase was eight batches. The attrition of immobilized d-hydantoinase particles with a large amount of insoluble substrate particles during stirring resulted in fine biocatalyst particles. In addition to the thermal inactivation, the loss of fine biocatalyst particles during the recovery step contributed to the low operational stability.  相似文献   

6.
A new sensitive high-performance liquid chromatographic procedure for the determination of l-carnitine (LC), acetyl-l-carnitine (ALC) and propionyl-l-carnitine (PLC) in human plasma has been developed. Precolumn derivatization with 1-aminoanthracene (1AA), performed in phosphate buffer in the presence of 1-(3-dimethylaminopropyl)-3-ethylcarbodiimide hydrochloride (EDC) as catalyst, is involved. The fluorescent derivatives were isocratically separated on a reversed-phase column (C18). The eluate was monitored with a fluorimetric detector set at 248 nm (excitation wavelength) and 418 nm (emission wavelength). Because of the presence of endogenous carnitines, the validation was performed using dialyzed plasma. The identity of the derivatized compounds was assessed by mass spectrometry and the purity of the chromatographic peaks was confirmed by HPLC-tandem mass spectrometry. The limits of quantitation were 5 nmol/ml for LC, 1 nmol/ml for ALC and 0.25 nmol/ml for PLC. The recovery of the extraction procedure was in the range 82.6%–95.4% for all 3 compounds. Good linearity (R≈0.99) was observed within the calibration ranges studied: 5–160 nmol/ml for LC, 1–32 nmol/ml for ALC and 0.25–8 nmol/ml for PLC. Precision was in the range 0.3–16.8% and accuracy was always lower than 10.6%.  相似文献   

7.
The GSH dependence of the metabolic pathways involved in the conversion of cysteine to sulfate in intact cells has been investigated. It was found that hepatocyte-catalysed sulfate formation from added

-cysteine did not occur if hepatocyte GSH was depleted beforehand, but was restored when GSH levels recovered. Furthermore, sulfate formation did not recover in GSH-depleted hepatocytes if GSH synthesis was prevented with buthionine sulfoximine. Thiosulfate formation was, however, markedly enhanced in GSH-depleted hepatocytes. These results suggest that thiosulfate is an intermediate in the formation of inorganic sulfate from

-cysteine and that GSH was required for the conversion of thiosulfate to inorganic sulfate. Much less sulfate was formed if the cysteine was replaced with cysteinesulfinate. Furthermore, sulfate formation from

-cysteine was markedly inhibited by the addition of the transaminase inhibitor

-cycloserine or the γ-cystathionase inhibitor

-propargylglycine. The major routes of sulfate formation from

-cysteine therefore seems to involve pathways that do not involve

-cysteinesulfinate. Similar amounts of sulfate were formed from

-cysteine as

-cysteine. Thiosulfate instead of sulfate was also formed in GSH-depleted hepatocytes. However, sulfate formation from

-cysteine differed from

-cysteine in that it was inhibited by the

-aminoacid oxidase inhibitor sodium benzoate and was not affected by transaminase or γ-cystathionase inhibitors. These results suggest that thiosulfate is an intermediate in sulfate formation from

-cysteine and involves the oxidation of

-cysteine by

-amino acid oxidase to form β-mercaptopyruvate.  相似文献   

8.
Four isomeric N-dimethylmaleoyl 4,6-O-benzylidene-protected d-hexosamine acceptors (2, 3, 4, and 5) with all possible configurations at C-1 and C-3 (e.g., derived from d-glucosamine and d-allosamine) were prepared, and the assessment of their O-3 relative reactivity through competition experiments using the known per-O-acetylated d-galactopyranosyl trichloroacetimidate donor (15) was then carried out. The reactivities are in the order 4 ? 2 > 5 > 3. The analysis of the NMR spectra of 2–5 at different temperature and modeling experiments carried out on analogs of 25 (DFT) and on the acceptors themselves (MM) are coincident, and have helped to establish the stability of the different hydrogen bonds, and of the conformers which carry them. The whole results suggest that the electronic effects (hydrogen bonds) are required to explain the observed trend, in spite of the axial conformation of the most reactive hydroxyl group. The steric effects appear only when hydrogen bonds are weak.  相似文献   

9.
A modification of the resorcinol method of Kulka1 for the determination of ketoses is described. Though being a stopped enzyme test, it is much more sensitive than the carbazol method and by applying microtiter plates and measuring with an ELISA reader, a large number of tests can be performed within a short time, thereby facilitating initial velocity studies.

The test is linear up to a concentration of 2.5 m -xylulose even in the presence of 10 m -xylose and 2 m -fructose in the presence of 10 m -glucose. The sensitivity is 25 μ for xylulose and 38 μ for fructose. The test method is insensitive to perturbations of substances frequently used in isolation procedures such as ammonium sulfate, Triton X-100, PEG 6000, sodium dodecyl sulfate, and ethanol in moderate concentrations.  相似文献   


10.
We have determined the content of free l-amino acids and d-aspartate in the nervous tissue of three representative cephalopods: Sepia officinalis, Octopus vulgaris, and Loligo vulgaris, and the optic lobes of adult and embryo Sepia officinalis. Taurine is the most abundant amino acid in the cephalopod nervous tissue. Its content amounts to more than 50% of the total free amino acids. The other most concentrated amino acids are Glu, Ala, Asp, and GABA. High concentrations of d-aspartate were found in the nervous tissue of all cephalopods examined (7–12 μmol/g wet tissue) which represents 50–80% of the total aspartate (d + l), depending on the animal. Among the various regions of the brain of Octopus vulgaris, d-aspartate was found to be evenly distributed in the various regions of the brain. In nerve tissue of Sepia officinalis, there is no significant difference in the pattern of free l-amino acids, in particular of the d-aspartate concentration, between adults and embryos, except for GABA, Gly, His and Thr. This suggests that d-aspartate in nerve tissue of the Cephalopoda is of endogenous origin and not a product of accumulation from exogenous sources. From a comparative study of the content of d-aspartate in the nervous tissue of different animals, we found that protostomia contain a significantly higher amount than deuterostomia. Thus, d-aspartate could be a criterion to distinguish the protostomia phyla from the deuterostomia phyla.  相似文献   

11.
Optical sensing of F, Cl, Br, I, OAc, , , and by cis-dithiocyanatobis(2,2′-bipyridyl-4,4′-dicarboxylic acid)ruthenium(II) (N3) and bis(tetrabutylammonium) cis-dithiocyanatobis(2,2′-bipyridine-4-COOH,4′-COO)ruthenium(II) (N719) have been investigated in dimethyl sulfoxide (DMSO), by means of UV-Vis absorption and emission spectrophotometric titrations. Additions of F, OAc, and in DMSO solution caused obvious UV-Vis spectral changes with appearance of several isosbestic points, and remarkable emission enhancements along with large blue shifts in emission bands. The values of F-induced emission intensity enhancement factor (emission quantum yield enhancement factor), I/I0 (φ/φ0), were found to be 40 (86) and 38 (58) for N3 and N719, respectively. No obvious spectral changes were observed upon addition of Cl, Br, I, and in DMSO solutions. Luminescent F sensing in DMSO/H2O (4:1, v/v) has also been demonstrated to be operative with a luminescence enhancement factor of 12, indicating that N3 is very potential for practical application as fluorescent anion sensor in aqueous solution. An interaction mechanism of anion-induced deprotonation of N3 and N719 was confirmed, and the deprotonation reaction equilibrium constants of N3 and N719 were derived as well.  相似文献   

12.
Ru/MCM-41 catalyst prepared by an impregnation–formaldehyde reduction method showed higher catalytic activity and sorbitol selectivity than other catalysts in the hydrogenation of glucose. SEM and XRD indicated the partial surface properties of Ru/MCM-41. Moreover, Ru dispersion and Ru surface area of Ru/MCM-41 were determined by pulse chemisorption, and the result further proved that Ru/MCM-41 had higher catalytic activity. A catalyst recycling experiment demonstrated that Ru/MCM-41 was a better catalyst and it could be reused three or four times. A speculated mechanism was proposed to illustrate the detailed process of d-glucose hydrogenation to produce sorbitol.  相似文献   

13.
A high-performance liquid chromatographic procedure has been developed for the determination of [d-Ala2, d-Leu5]enkephalin (DADLE) and the fragments containing d-leucine in rat blood. The procedure was applied to the determination of blood levels of [3H-d-Leu5]DADLE and the C-terminal fragments after intravenous administration of [3H-d-Leu5]DADLE to a rat. Unlabelled DADLE and the C-terminal fragments were spiked as carriers to rat blood samples and the blood samples were extracted with 1% trifluoroacetic acid in methanol. The recoveries from rat blood were quantitative for all compounds. DADLE and the C-terminal four fragments were well separated on a reversed-phase column with gradient elution using a mobile phase composed of 0.14% HClO4 and acetonitrile.  相似文献   

14.

-Tyrosyl-

-arginine (kyotorphin) is known as an endogenous analgesic neuropeptide. We examined whether kyotorphin and other arginine-containing neuropeptides were endogenous substrates for neuronal nitric oxide synthase (NOS) in the rat brain. Cytosol fractions of the rat cerebellum contained higher concentrations of neuronal NOS (nNOS) than endothelial NOS. In rat cerebellar cytosol, the binding activity of [3H]NG-nitro-

-arginine (NNA) was inhibited equally by

-arginine (

-Arg), kyotorphin, and

-leucyl-

-Arg (a kyotorphin receptor antagonist). Binding activities were inhibited to lesser degrees by fibronectin active fragments, bradykinin, and dynorphin A, but were not inhibited by

-tyrosyl-

-Arg or substance P. Interestingly, the inhibition of [3H]NNA binding by kyotorphin was attenuated by inhibitors of kyotorphin-hydrolyzing peptidases (KTPases) such as bestatin and arphamenine B. These results suggest that kyotorphin is degraded to

-Arg by KTPases, which in turn may act as substrate for nNOS.  相似文献   

15.
The quest for more specific drugs in antitumor chemotherapy led us to the design of anthraquinone–peptide conjugates capable of selective recognition of the nucleic acid. We present here the DNA binding characteristics, sequence specificity and geometry of interaction of a pair of enantiomers containing the lysine–glycine dipeptide in the side chains. The d enantiomer binds right handed double stranded DNA more efficiently than the l form under all conditions tested. The source of higher binding affinity is not electrostatic in nature and rests in the more favorable hydrophobic contacts of the d-lysyl side chains in the drug-DNA complex. Both derivatives exhibit preference for alternating GC base sequences and intercalate into DNA in a threading mode as suggested by chiroptical and theoretical studies. The d enantiomer, being a peptidyl derivative that contains a non-natural amino acid, has the considerable advantage of being less susceptible to enzymatic hydrolysis and could therefore represent a lead compound for further development.  相似文献   

16.
Protein (d-aspartyl/l-isoaspartyl) carboxyl methyltransferase (PCM, E.C. 2.1.1.77) was previously shown to be enzymatically methyl esterified in an autocatalytic manner at altered aspartyl residues; methyl esters are observed in a subpopulation of the enzyme termed thePCM fraction [Lindquist and McFadden (1994),J. Protein Chem. 13, 23–30]. The altered aspartyl sites serving as methyl acceptors inPCM have now been localized by using proteolytic enzymes and chemical cleavage techniques in combination with matrix-assisted laser desorption/ionization (MALDI) mass spectrometry to identify fragments of the [3H]automethylated enzyme that contain a [3H]methyl ester. Methylation was positively identified at positions Asn188 and Asp217 in the enzyme sequence, a consequence of the spontaneous alteration of these sites tol-isoaspartyl ord-aspartyl sites and their methylation by active PCM molecules. The identification of more than one site of automethylation shows thatPCM is not a homogeneous population of damaged PCM molecules, but rather a complex population of molecules with a variety of age-altered damage sites.Abbreviations PCM protein (d-aspartyl/l-isoaspartyl) carboxyl methyltransferase - EDTA disodium ethylenediaminetetraacetate - PMSF phenylmethylsulfonyl fluoride - TEA trifluoroacetic acid - HPLC high-pressure liquid chromatography  相似文献   

17.
The C-terminal heptapeptide-amide (C7-sorbin) is the minimal biologically active fragment of sorbin inducing an increase in intestinal hydroelectrolytic absorption. An analogue (D7-sorbin), characterized by the replacement of the ultimate C-terminal amino acid l-alanine-amide by d-alanine-amide, was synthetized. For pharmacokinetic studies, D7-sorbin and C7-sorbin were tritium labeled. After IV injection, clearances were 10.6 and 30.2 ml−1 for D7-sorbin and C7-sorbin, respectively, and MRT were 34 and 18 min. After SC administration, Cmax attained 0.41% and 0.12% of the dose/ml, respectively. The IP route showed a 45-min delay before Cmax and a 100% bioavailability for both peptides. D7-sorbin was principally excreted in urine, as shown by balance study, and in part in intact form, as controlled by mass spectrometry. D7-sorbin induced a significant decrease of the VIP-induced ileal secretion, previously observed with C7-sorbin. The change of l-Ala to d-Ala increased the stability of the synthetic C-terminal peptide of sorbin whereas its biological activity, bioavailability, and route of elimination were unchanged.  相似文献   

18.
2,6-Dideoxy- -lyxo-hexose was prepared efficiently from -galactose in eight steps (25% overall yield). The synthesis is amenable to multigram scale-up.  相似文献   

19.
Objective: In this clinical trial we studied whether oral supplementation with d-alpha-tocopherol (α-Toc), l-ascorbic acid (Asc), or α-Toc combined with Asc influenced the solar simulated radiation (SSR) induced skin inflammation in healthy volunteers. Methods: We investigated the following groups in a prospective, randomized and placebo controlled study: Group (1) α-Toc 2 g / day, group (2) Asc 3 g / day, group (3) α-Toc 2 g / day combined with Asc 3 g / day, and group (4) placebo. Before and 50 days after supplementation we analyzed α-Toc and Asc concentrations in keratinocytes. The dose response curve of UV erythema was determined by reflectance spectrophotometry and the minimal erythema dose (MED) by visual grading before and after supplementation. Results: 50 days after supplementation α-Toc keratinocyte levels were increased in groups (1) and (3), Asc concentrations were elevated in groups (2) and (3), and the a/γ-Toc ratio increased in groups (1) and (3). The dose response curve of UVR induced erythema showed a significant flattening and the MED increased from 103 ± 29 mJ/cm2 (before supplementation) to 183 ± 35 mJ/cm2 (after supplementation) in group (3), while there were no significant changes in groups (1) and (2) after vitamin supplementation. Conclusion: α-Toc and Asc act synergistically in suppression of the sunburn reaction.  相似文献   

20.
Methods with which to simply and rapidly assay l-aspartate (l-Asp) and d-aspartate (d-Asp) would be highly useful for physiological research and for nutritional and clinical analyses. Levels of l- and d-Asp in food and cell extracts are currently determined using high-performance liquid chromatography. However, this method is time-consuming and expensive. Here we describe a simple and specific method for using an l-aspartate dehydrogenase (l-AspDH) system to colorimetrically assay l-Asp and a system of three hyperthermophilic enzymes—aspartate racemase (AspR), l-AspDH, and l-aspartate oxidase (l-AO)—to assay d-Asp. In the former, the reaction rate of nicotinamide adenine dinucleotide (NAD+)-dependent l-AspDH was measured based on increases in the absorbance at 438 nm, reflecting formation of formazan from water-soluble tetrazolium-1 (WST-1), using 1-methoxy-5-methylphenazinum methyl sulfate (mPMS) as a redox mediator. In the latter, d-Asp was measured after first removing l-Asp in the sample solution with l-AO. The remaining d-Asp was then changed to l-Asp using racemase, and the newly formed l-Asp was assayed calorimetrically using NAD+-dependent aspartate dehydrogenase as described above. This method enables simple and rapid spectrophotometric determination of 1 to 100 μM l- and d-Asp in the assay systems. In addition, methods were applicable to the l- and d-Asp determinations in some living cells and foods.  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