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 purification of rabbit immunoglobulin molecules expressing kappa (κ) light chains, utilizing the allotypic specificity b4, has been achieved in stages involving isolation of specific antibody, preparation of a solid phase immunoadsorbent of coupled antibody, and subsequent isolation of b4 (κ) IgG. Cellulose trans-2.3-carbonate is shown to be an effective matrix enabling chemical coupling of antibodies and antigens to the support at neutral pH thus preservng immunological activity. The trans-2,3-carbonate derived from microcrystalline cellulose is more effective as a matrix than the trans-2,3-carbonate derived from macroporous cellulose for the chemical coupling of rabbit a1a3/b4 IgG antigen and binding of specific anti-b4 antibody. The microcrystalline celulose carbonate is also more efficient for the coupling of rabbit anti-b4 antibody and the subsequent binding and elution of rabbit b4 (κ) IgG, thus separating immunoglobulin, expressing kappa light chain, from that expressing lambda light chain. The purification technique has potential application in other allotypic systems and antibody- antigen populations.  相似文献   

2.
Summary 1-Alkyn-3-ols were resolved through the lipase-catalyzed enantioselective hydrolysis of the phenyl carbonates. Resolution of 1-heptyn-3-yl phenyl carbonate gave (R)-carbonate with an optical purity of 98%ee. While, 1-octyn-3-yl phenyl carbonate gave optically pure (S)-carbonate.  相似文献   

3.
Laser-Raman spectra of L-α-dimyristoylphosphatidylcholine (DMPC) liposomes in the spectral range 1000–1200 cm?1 were obtained as a function of temperature from ?80 to +50°C. The triplet found in this spectral region was resolved into Lorentzian components by means of an iterative computer program. The peak intensities, band widths, and band areas of the resolved 1062 cm?1 and 1130 cm?1 bands, assigned to CC stretching vibrations of trans segments, were evaluated as a function of temperature. While the peak intensities of the bands decrease substantially with temperature, the band widths show a considerable increase. The change in band areas is therefore smaller than the change in peak heights. Experiments with all trans carboxylic acids showed that in these compounds the area of the Raman bands at 1062 cm?1 and 1130 cm?1 is proportional to the number of trans bonds. The variation with temperature of the number of trans and gauche bonds in the studied phospholipid is reflected by the change of the area of the 1130 cm?1 Raman band.  相似文献   

4.
The i.r. spectra for aqueous solutions of sulfated glycosaminoglycans and model compounds in the transmittance “window” region of the solvent (1400-950 cm?1) are dominated by the strong and complex absorption centered at ~1230 cm?1 and associated with the antisymmetric stretching vibrations of the SO groups. Primary and secondary O-sulfate groups absorb at somewhat higher frequencies (1260-1200 cm?1) than N-sulfates (~1185 cm?1). Each sulfate band lends itself to quantitative applications, especially within a given class of sulfated polysaccharide. Laser-Raman spectra of heparin and model compounds have been obtained in aqueous solution and in the solid state. The most-prominent Raman peak (at ~1060 cm?1) is attributable to the symmetrical vibration of the SO groups, with N-sulfates emitting at somewhat lower frequencies (~1040 cm?1) than O-sulfates. The Raman pattern in the 950-800 cm?1 region (currently used in the i.r. for distinguishing between types of sulfate groups) also involves vibrations that are not localized only in the COS bonds.  相似文献   

5.
The conversion of cis-[14C]phytofluene to trans-[14C]phytofluene and the conversion of the latter to trans-ζ-[14C]carotene by a soluble enzyme system obtained from plastids of tangerine tomato fruits is reported. Each of these compounds is also converted to cis-ζ-carotene, proneurosporene, prolycopene, neurosporene, lycopene, and γ- and β-carotenes. [14C]Prolycopene was also incubated with the above enzyme system. No conversion of this compound to trans-lycopene or cyclic carotenes was observed. Proof for the formation of the above carotenes from each of the substrates mentioned above was obtained by cochromatography with authentic samples on an alumina column. A close correspondence between radioactivity and light absorbance of each carotene was observed. Further proof for the formation of acyclic and cyclic carotenes from the above radioactive substrates was obtained by gas-liquid chromatography of the hydrogenated products. Coincidence between mass and radioactivity was observed in each case.  相似文献   

6.
Abstract

The nucleoside derivative 1-(3-azido-2,3-dideoxy-beta-D-ribo-hexofuranosyl)thymine has been synthesized from 3-0-benzyl-1,2-0-isopropylidene-alpha-D-glucofuranose-5,6-carbonate in an overall yield of 16%. The key step in the synthesis involves the selective deacetylation of a nucleoside derivative having a cyclic carbonate moiety.  相似文献   

7.
Insertion of nickel ions into the empty catalytic site of horse liver alcohol dehydrogenase yields an active enzyme with 65% metal substitution and about 12% intrinsic activity. The electronic absorption spectrum is characterized by bands at 357 nm (2900 M?1 cm?1, 407 nm (3500 M?1 cm?1), 505 nm (300 M?1 cm?1), 570 nm (?130 M?1 cm?1), and 680 nm (?80 M?1 cm?1). The absorption and CD spectra are similar to those of nickel(II) azurin and nickel(II) aspartate transcarbamoylase and prove coordination of the nickel(II) ions to sulfur in a distorted tetrahedral coordination geometry. Changes of the spectra upon ligand binding at the metal or conformation changes of the protein induced by coenzyme, or both, indicate alterations of the metal geometry.The chromophoric substrate trans-4-(N, N-dimethylamino)-cinnamaldehyde forms a ternary complex with Ni(II) liver alcohol dehydrogenase and the coenzyme analogue 1,4,5,6-tetrahydronicotinamide-adenine-dinucleotide, stable between pH 6 and 10. The corresponding ternary complex with NADH is only stable at pH > 9.0. The spectral redshifts induced in the substrate are 11 nm larger than those found in the zinc enzyme. We suggest direct coordination of the substrate to the catalytic metal ion which acts as a Lewis acid in both substrate coordination and catalysis.  相似文献   

8.
Much is known about the bacterial precipitation of carbonate rocks, but comparatively little is known about the involvement of microbes in the formation of secondary mineral structures in caves. We hypothesized that bacteria isolated from calcareous stalactites, which are able to mediate CaCO3 precipitation in vitro, play a role in the formation of carbonate speleothems. We collected numerous cultivable calcifying bacteria from calcareous speleothems from Cervo cave, implying that their presence was not occasional. The relative abundance of calcifying bacteria among total cultivable microflora was found to be related to the calcifying activity in the stalactites. We also determined the δ 13C and δ 18 O values of the Cervo cave speleothems from which bacteria were isolated and of the carbonates obtained in vitro to determine whether bacteria were indeed involved in the formation of secondary mineral structures. We identified three groups of biological carbonates produced in vitro at 11°C on the basis of their carbon isotopic composition: carbonates with δ 13C values (a) slightly more positive, (b) more negative, and (c) much more negative than those of the stalactite carbonates. The carbonates belonging to the first group, characterized by the most similar δ 13C values to stalactites, were produced by the most abundant strains. Most of calcifying isolates belonged to the genus Kocuria. Scanning electron microscopy showed that dominant morphologies of the bioliths were sherulithic with fibrous radiated interiors. We suggest a mechanism of carbonate crystal formation by bacteria.  相似文献   

9.
This paper reports the conversion of cis-[14C]phytofluene to trans-[14C|phytofluene and the conversion of the latter compound to trans-ζ-[14C]carotene by a soluble enzyme system obtained from the plastids of red tomato fruits. Each of these radioactive compounds was also converted to labeled neurosporene, lycopenc, α-carotene, and β-carotene by the same enzyme system. The incorporation of each substrate into more unsaturated carotenes was carried out under nitrogen at pH 7.5–8.2 (borate buffer), at 25 °C in the dark.Proof of the formation of the above carotenes from each of the three radioactive substrates was demonstrated by cochromatography with authentic nonradioactive carotenes on an alumina chromatographic column. A close correspondence between radioactivity and light absorbance for each carotene was observed. Confirmation of these conversions was achieved by cochromatography with authentic samples on thinlayer plates. Final proof for the formation of the acyclic and cyclic carotenes from the above radioactive substrates was obtained by gas-liquid chromatography of the hydrogenated products. Coincidence between mass and radioactivity was observed.Maximum conversion of cis- and trans-phytofluenes to more unsaturated carotenes by the red tomato fruit enzyme system appears to be dependent upon the presence of NADP+, FAD, and Tween 80. The formation of the carotenes is also increased in the presence of Mg2+ or Mn2+ ions.  相似文献   

10.
The effects of pressure on the sol-gel transition of κ- and ι-carrageenans were studied in KCl solutions under high pressures up to 3000 kg/cm2. The carrageenan gels were destabilized by pressure: the pressure depression of melting temperature, (dT/dP)m, was ?5.7 × 10?3 and ?4.0 × 10?3 K cm2/kg independent of KCl concentration for κ- and ι-carrageenans, respectively. The enthalpy, entropy and volume changes accompanying the gel formation were calculated from the Eldridge-Ferry's plots and the Clausius-Clapeyron equation. The volume change per unit cross-link (two disaccharide residues) was estimated to be (2.5 ~ 4.9) and (1.7 ~ 3.4) ml/mol for κ- and ι-carrageenans, respectively. The compressibility of both carrageenan molecules appeared to be larger by (1.6 ~ 2.6) × 10?12 (κ-form) and by (0.8 ~ 1.3) × 10?12cm2/dyn (i-form) in gel state as compared with in sol state These increases in volume and compressibility on gelation were attributed to a reduction of water of hydration from the carrageenan molecules, which is mainly due to a replacement of the polymer-water hydrogen bond by the polymer-polymer hydrogen bond. These results seemed not inconsistent with the idea that a double helix structure of carrageenan gels may persist in solution as well as in the solid state.  相似文献   

11.
All-trans retinoyl fluoride was prepared by treating all-trans retinoic acid with diethylaminosulfurtrifluoride. The crystalline product, which was characterized by melting point, infrared, 1H-NMR, 19F-NMR and elementary analysis, showed λmax at 382 nm in hexane (ε = 4.98·104 M?1·cm?1) and at 392 nm in methanol (ε = 4.60·104 M?1·cm?1). Its biological activity in the rat growth assay, relative to all-trans retinyl acetate, was 22% ± 10%. Upon oral administration for 5 days to vitamin A-depleted rats, retinoyl fluoride (1020 μg) was rapidly metabolized to a polar metabolite fraction and, in the intestine, to an unstable retinol-like metabolite, purpotedly 15-fluororetinol. Upon administering intraperitoneally smaller doses (47–94 μg) of [11-3H]retinoyl fluoride, which was synthesized from [11-3H] retinoic acid, radioactive retinoic acid was noted in the liver and plasma but not in the intestine. As expected, a radioactive polar fraction appeared in the intestine and liver, but radioactive retinol, retinyl ester and some common oxidation products were not detected. Of the administered radioactivity, 72% was excreted in the urine, and only 4% was found in the feces over a 7-day period. Hydrolysis of the urine gave a radioactive fraction with a polarity similar to that of retinoic acid. Retinoyl fluoride also reacts readily with glycine to yield N-retinoyl glycine. Thus, the biological activity of retinoyl fluoride can be attributed to the formation of retinoic acid, probably by way of N-retinoyl derivatives. A possible pathway for its metabolism is presented.  相似文献   

12.
A Bacillus sp. strain named BRC1 is capable of producing 2,3-butanediol (2,3-BD) using hydrolysates of the Jerusalem artichoke tuber (JAT), a rich source of the fructose polymer inulin. To enhance 2,3-BD production, we undertook an extensive analysis of the Bacillus sp. BRC1 genome, identifying a putative gene (sacC) encoding a fructan hydrolysis enzyme and characterizing the activity of the resulting recombinant protein expressed in and purified from Escherichia coli. Introduction of the sacC gene into Bacillus sp. BRC1 using an expression vector increased enzymatic activity more than twofold. Consistent with this increased enzyme expression, 2,3-BD production from JAT was also increased from 3.98 to 8.10 g L?1. Fed-batch fermentation of the recombinant strain produced a maximal level of 2,3-BD production of 28.6 g L?1, showing a high theoretical yield of 92.3%.  相似文献   

13.
The structure and reactivity of the complex [Ru(2,3-Medpp)2Cl2](PF6)2 (2,3-Medpp+=2-[2-(1-methylpyridiniumyl)]-3-(2-pyridyl)pyrazine) was investigated by X-ray diffraction (XRD), 1H NMR, redox, and UV-Vis absorption measurements. X-ray analysis shows that crystals obtained from an acetonitrile-toluene solution contain the trans-Cl2, trans-pyrazine isomeric form, while 1H NMR and redox measurements on the main product of the synthetic workup indicate the presence of the trans-Cl2, cis-pyrazine isomer. In the dark at 70 °C, the complex [Ru(2,3-Medpp)2Cl2]2+ reacts slowly in acetonitrile isomerizing to the cis-[Ru(2,3-Medpp)2(CH3CN)Cl]3+ species. Under ambient light in the presence of excess AgNO3 the cis-[Ru(2,3-Medpp)2(CH3CN)2]4+ species is obtained.  相似文献   

14.
This paper reports the first Raman spectroscopic study of the potassium complex of the cation-specific antibiotic valinomycin. Complete Raman spectra (140 to 3600 cm?1) of crystalline valinomycin-KSCN and its CCl4, CHCl3 and C2H5OH solutions are presented and used to probe the structure of the complex in these environments. In all cases a single, narrow peak is observed in the ester CO stretch region (1750 to 1775 cm?1) which contrasts strongly with the broad bands observed in solutions of uncomplexed valinomycin. This is consistent with the presence of a single conformation in which all six ester CO groups co-ordinate an enclosed potassium ion. We find that although the ester CO stretch frequencies of the complex are similar in the solid state and in non-polar solution (~1770 cm?1) they are considerably different in the presence of polar solvents (~1756 cm?1); this may indicate that the complexed potassium ion is still free to interact with nearby solvent ions (and possibly its counterion) through gaps in the hydrophobic “shield” provided by the hydrocarbon residues of valinomycin. In contrast the amide CO frequencies of the complex (~1650 cm?1) are solvent-independent. These groups are apparently strongly hydrogen-bonded to provide a rather rigid, compact framework for the complex conformation.  相似文献   

15.
The use of 4,4′-bis-dimethylaminodiphenylcarbinol (BDC-OH) as an analytical reagent for sulfhydryl residues and as a specific chemical modification reagent for proteins is dependent upon the unique properties of the BDC+ cation present in aqueous buffers below a pH of 6.5. In the presence of aqueous buffers, pH 5.1, BDC+ exhibits a λmax of 606 nm with an apparent molar absorption coefficient of 10,000 m?1 cm?1. Upon the addition of 4m guanidine hydrochloride this apparent coefficient is enhanced to 70,800 m?1 cm?1. The true molar extinction coefficient for BDC+ was determined to be 128,000 m?1 cm?1. The reaction of BDC+ with sulfhydryl residues of proteins or simple thiols is rapid and leads to a complex devoid of visible color. In the pH range 3.0–7.0, a complex equilibrium is established among the three species BDC-OH, BDC+, BDCH++. The formation of this equilibrium is proton mediated, and is discussed in terms of the equilibrium, rate, and acid dissociation constants.  相似文献   

16.
Summary

Wheat bran contains several ester-linked dehydrodimers of ferulic acid, which were detected and quantified after sequential alkaline hydrolysis. The major dimers released were: trans-5-[(E)-2-carboxyvinyl]-2-(4-hydroxy-3-methoxy-phenyl)-7-methoxy-2,3-dihydrobenzofuran-3-carboxylic acid (5–8-BendiFA), (Z)-β-(4-[(E)-2-carboxyvinyl]-2-methoxy-phenoxy)-4-hydroxy-3-methoxycinnamic acid (8-O-4-diFA) and (E,E)-4,4′-dihydroxy-5,5′-dimethoxy-3,3′-bicinnamic acid (5–5-diFA). trans-7-hydroxy-1-(4-hydroxy-3methoxyphenyl)-6-methoxy-1,2-dihydro-naphthalene-2,3-dicarboxylic acid (8–8-diFA cyclic form) and 4,4′-dihydroxy-3,3′-dimethoxy-β,β'-bicinnamic acid (8–8-diFA non cyclic form) were not detected. One of the most abundant dimers, 8-O-4-diFA, was purified from de-starched wheat bran after alkaline hydrolysis and preparative HPLC. The resultant product was identical to the chemically synthesised 8-O-4-dimer by TLC and HPLC as confirmed by 1H-NMR and mass spectrometry. The absorption maxima and absorption coefficients for the synthetic compound in ethanol were: λmax: 323 nm, λmin: 258 nm, ελmax (M?1cm?1): 24800 ± 2100 and ε280 (M?1cm?1): 19700 ± 1100. The antioxidant properties of 8-O-4-diFA were assessed using: (a) inhibition of ascorbate/iron-induced peroxidation of phosphatidylcholine liposomes and; (b) scavenging of the radical cation of 2,2′-azinobis (3-ethyl-benzothiazoline-6-sulphonate) (ABTS) relative to the water-soluble vitamin E analogue, Trolox C. The 8-O-4-diFA was a better antioxidant than ferulic acid in both lipid and aqueous phases. This is the first report of the antioxidant activity of a natural diferulate obtained from a plant.  相似文献   

17.
This study focuses on the analysis of polysaccharide residues from the cell walls of fruits and vegetables: tomato, potato, pumpkin, carrot and celery root. An alcohol-insoluble residue was prepared from plant material by extraction using the hot ethyl alcohol method and then cell wall fractions soluble in trans-1,2-diaminocyclohexane-N,N,N′,N′-tetraacetate, sodium carbonate and alkaline solution were sequentially extracted. Infrared spectroscopy combined with Fourier transform (FT-IR) was used to evaluate differences among cell wall residues and among species after each step of sequential extraction of pectins and hemicelluloses. Additionally, pectic substances were identified using an Automated Wet Chemistry Analyser. Principal component analysis (PCA) was applied to FT-IR spectra in two regions: 1,800–1,200 cm?1 and 1,200–800 cm?1 in order to distinguish different components of cell wall polysaccharides. This method also allowed us the possibility of highlighting the most important wavenumbers for each type of polysaccharide: 1,740, 1,610 and 1,240 cm?1 denoting pectins or 1,370 and 1,317 cm?1 denoting hemicelluloses and cellulose, respectively.  相似文献   

18.
Fourier transform-infrared (FT-IR) spectroscopy has been used to rapidly and nondestructively analyze bacteria, bacteria-polymer mixtures, digester samples and microbial biofilms. Diffuse reflectance FT-IR (DRIFT) analysis for freeze-dried, powdered samples offered a means of obtaining structural information. The bacteria examined were divided into two groups. The first group was characterized by a dominant amide I band and the second group of organisms displayed an additional strong carbonyl stretch at ~ 1740 cm?1. The differences illustrated by the subtraction spectra obtained for microbes of the two groups suggests that FT-IR spectroscopy can be utilized to recognize differences in microbial community structure. Calculation of specific band ratios has enabled to composition of bacteria and extracellular or intracellular storage product polymer mixtures to be determined for bacteria-gum (amide I/carbohydrate C-O-~ 1150 cm?1) and bacteria-poly-β-hydroxybutyrate (amide I/carbonyl ~ 1740 cm?1). The key band ratios correlate with the compositions of the material and provide useful information for the application of FT-IR sepectroscopy to environmental biofilm samples and for distinguishing bacteria grown under differing nutrient conditions. DRIFT spectra have been obtained for biofilms produced by Vibrio natriegens on stainless steel disks. Between 48 and 144 h, an increase in bands at ~ 1740 cm?1 was seen in FT-IR spectra of V. natriegens biofilm. DRIFT spectra of mixed culture effluents of anaerobic digesters show differences induced by shifts in input feedstocks. The use of flow-through attenuated total reflectance has permitted in situ real-time changes in biofilm formation to be monitored and provides a powerful tool for understanding the interactioni within adherent microbial consortia.  相似文献   

19.
In this study, a propionate CoA-transferase (H16_A2718; EC 2.8.3.1) from Ralstonia eutropha H16 (Pct Re ) was characterized in detail. Glu342 was identified as catalytically active amino acid residue via site-directed mutagenesis. Activity of Pct Re was irreversibly lost after the treatment with NaBH4 in the presence of acetyl-CoA as it is shown for all CoA-transferases from class I, thereby confirming the formation of the covalent enzyme-CoA intermediate by Pct Re . In addition to already known CoA acceptors for Pct Re such as 3-hydroxypropionate, 3-hydroxybutyrate, acrylate, succinate, lactate, butyrate, crotonate and 4-hydroxybutyrate, it was found that glycolate, chloropropionate, acetoacetate, valerate, trans-2,3-pentenoate, isovalerate, hexanoate, octanoate and trans-2,3-octenoate formed also corresponding CoA-thioesters after incubation with acetyl-CoA and Pct Re . Isobutyrate was found to be preferentially used as CoA acceptor amongst other carboxylates tested in this study. In contrast, no products were detected with acetyl-CoA and formiate, bromopropionate, glycine, pyruvate, 2-hydroxybutyrate, malonate, fumarate, itaconate, β-alanine, γ-aminobutyrate, levulate, glutarate or adipate as potential CoA acceptor. Amongst CoA donors, butyryl-CoA, crotonyl-CoA, 3-hydroxybutyryl-CoA, isobutyryl-CoA, succinyl-CoA and valeryl-CoA apart from already known propionyl-CoA and acetyl-CoA could also donate CoA to acetate. The highest rate of the reaction was observed with 3-hydroxybutyryl-CoA (2.5 μmol mg?1 min?1). K m values for propionyl-CoA, acetyl-CoA, acetate and 3-hydroxybutyrate were 0.3, 0.6, 4.5 and 4.3 mM, respectively. The rather broad substrate range might be a good starting point for enzyme engineering approaches and for the application of Pct Re in biotechnological polyester production.  相似文献   

20.
The energetically most stable head group conformations of a racemic mixture of diacyl-glycero-phosphorylethanolamine in a planar quasihexagonal lattice were calculated using empirical 1-6-12 atom-atom potential functions for intra- and intermolecular interactions. The results demonstrate that the conformation of phospholipid head groups in bilayer systems is determined by intramolecular interactions as well as by intermolecular interactions with neighbouring phospholipid molecules and with solvent molecules. The most stable conformers are that with a φ2 = guache? conformation of the phosphodiester group. All conformers with a φ2 = gauche(+) or trans conformation have energies more than 15 kcal ☆ mol?1 above that of the global minimum. The calculated torsional angles ?1 and φ1 are in very good agreement with the results of the X-ray diffraction analysis of 1,2-dilauroyl-DL-phosphatidylethanolamine (DLPE) acetic acid single crystals.  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