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 531 毫秒
1.
Phosphorescence from the 9-adenylyl group in the fom of microcrystalline powders of adenosine films of poly(riboadenylic acid) (poly(rA)) in hyaluronic acid has been studied at 77 K. For adenosine, clearly resolved vibronic structure consists of two progressions, A and B, with A ‡ 1363 cm−1 and B ‡ 1575 cm −1, correlated with in-plane C5-N7 and in-plane C4-C5 stretch, respectively. The relative strength of the progressions varies with excitation wavelength and this, together with the absence of a common origin, indicates the existence of two independent emitting states with 0-0' levels separated by either 300 or 1000 cm−1. Two different excitation spectra are observed lying below the normal (ππ*) adsorption and one is assigned as a previously undetected 1(nπ*) transition. For poly(rA) films the emission band envelope is identical with that of adenosine but the vibronic structure is lost. Only one excitation peak is observed at 32.9×103 cm−1, identical with one of the adenosine spectra. The second adenosine excitation spectrum probably represents an intermolecular charge transfer transition. Comparison is made with the predictions of six semi-empirical MO calculations.  相似文献   

2.
[NBun4]2[W(C3Se5)3] (C3Se52− = 1,3-diselenole-2-selone-4,5- diselenolate(2−)) was prepared by the reaction of Na2[C3Se5] with WCl6 in ethanol, followed by addition of [NBun4]Br. The cyclic voltammogram in dichloromethane exhibits two oxidation peaks at −0.04 and +0.03 V (versus SCE). The complex reacted with [Fe(C5Me5)2][BF4], iodine or [TTF]3[BF4]2 (TTF·+ = the tetrathiafulvalenium radical cation) in acetonitrile to afford the oxidized complexes [Fe(C5Me5)2]0.5[W(C3Se5)3], [NBun4]0.1[W(C3Se5)3] and [TTF]0.5[W(C3Se5)3], respectively. Current-controlled electrochemical oxidation of the complex in acetonitrile gave [NBun4]0.6[W(C3Se5)3]. The oxidized complexes exhibit electrical conductivities of 4.7×10 −5−1.5×10−3 S cm−1 at room temperature measured for compacted pellets. Electronic absorption, IR and ESR spectra of these complexes are discussed.  相似文献   

3.
Rates of fixation of mutations during the evolution of the foot-and-mouth disease virus (FMDV) C1 in nature have been estimated by hybridization of viral RNA to cloned cDNAs representing defined FMDV genome segments, and comparison of the selected RNAs by T1 RNase oligonucleotide fingerprinting. Values ranged from <0.04 × 10−2 to 4.5 × 10−2 substitutions per nucleotide per year (s/nt/yr), depending on the time period and the genomic segment considered. Rates for viral structural protein genes were up to sixfold higher than for nonstructural protein genes. Values in excess of 10−2 s/nt/yr have been measured for the RNA region that encodes VP1–VP3. The nucleotide sequences of the major immunogenic region of capsid protein VP1 have been determined for six new FMDV C1 isolates, and they are compared with the two previously known sequences of FMDV C1 (C-S8 and C1-O). Both oligonucleotide fingerprinting of selected RNA fragments and direct nucleotide sequencing demonstrate that genetic heterogeneity exists among three viruses isolated on the same day, introducing a significant indetermination in the evaluation of fixation rates of mutations. During the FMDV C1 outbreak, amino acid substitutions did occur that are known to affect the immunological properties of the virus. The proportion of mutations between two viral RNAs does not increase significantly with the time elapsed between the two isolations, suggesting a cocirculation of multiple, related, nonidentical FMDVs (‘evolving quasispecies’) as the mode of evolution of this agent.  相似文献   

4.
The porcine pancrease lipase was immobilized by entrapment in the beads of K-carrageenan and cured by treatment with polyethyleneimine (PEI) in the phosphate buffer. The retention of hydrolytic activity of lipase and compressive strength of the beads were examined. The activity of free and immobilized lipase was assessed by using olive oil as the substrate. The immobilized enzyme exhibited a little shift towards acidic pH for its optimal activity and retained 50% of its activity after 5 cycles. When the enzyme concentration was kept constant and substrate concentration was varied the Km and Vmax were observed to be 0.18 × 10−2 and 0.10, and 0.10 × 10−2 and 0.09 respectively, for free and for entrapped enzymes. When the substrate concentration was kept constant and enzyme concentration was varied, the values of Km and Vmax were observed to be 0.19 × 10−7 and 0.41, and 0.18 × 10−7 and 0.41 for free and entrapped enzymes. Though this indicates that there is no conformational change during immobilization, it also shows that the reaction velocity depends on the concentration. Immobilized enzyme showed improved thermal and storage stability. Hydrolysis of olive oil in organic–aqueous two-phase system using fixed bed reactor was carried out and conditions were optimized. The enzyme in reactor retained 30% of its initial activity after 480 min (12 cycles).  相似文献   

5.
The thermolysis of trans-IrL2(CO)Cl(H)(C6H5) (1abd; L=P(i-Pr)3; H trans to CO) produces benzene and the Vaska-type complex IrL2(CO)Cl. A mechanistic study of the reaction has shown that 1a reversibly loses CO at 120 °C (as evidenced by the incorporation of 13CO) and isomerizes to the previously unreported 1b (H trans to Cl). It was found that 1b is the complex primarily responsible for the formation of benzene upon thermolysis under CO atmosphere; direct loss of benzene from 1a was determined to be, at most, a minor pathway. Benzaldehyde was also formed as a product of thermolysis of 1a under CO atmosphere. The first-order rate constant for benzene elimination in the absence of CO was found to be 8.5 × 10−5 s−1. The presence of only 5 Torr CO results in a decrease to 2.0 × 10−5 s−1, but little further inhibition is observed above 5 Torr CO. Added dihydrogen (100 Torr) was found to effect a novel catalysis of benzene elimination from 1a in the absence of CO atmosphere; it is suggested that trace amounts of dihydrogen, present in solutions of 1a, are responsible for the enhanced rate of elimination in the absence of CO. The thermolysis of 1-d6 in toluene was found to proceed without any toluene incorporation, implying that arene loss is irreversible.  相似文献   

6.
Nitrogen dioxide radical (NO·2) is known as a toxic agent produced in the metabolism of nitrates and nitrites. By the use of the pulse radiolysis technique, the mechanism of the reaction of NO·2 radical with hydroxycinnamic acid derivatives (HCA) was studied and the rate constants have been measured. The rate constants were found to be 7.4 × 108, 7.2 × 108, 8.6 × 108 dm3 mol-1s-1 for ferulic acid, sinapic acid and caffeic acid, respectively. The reactions produce the corresponding phenoxyl radical.  相似文献   

7.
Biological properties of amino-terminal PTHrP analogues modified in the region 11–13 were examined using ROS 17/2.8 cells. [Leu11,D-Trp12,Arg13,Tyr36]PTHrP(1–36)amide had a 17-fold lower binding affinity for the receptor (apparent Kd: 5 × 10−8 M) than [Tyr36]PTHrP(1–36)amide or [Arg11,13,Tyr36]PTHrP(1–36)amide (apparent Kd for both: 2 × 10−9 M). Moreover, it is only a weak partial agonist despite completely inhibiting radioligand binding. [Leu11,D-Trp12,Arg13,Tyr36,Cys38]PTHrP(7–38) and PTHrP(7–34)amide had similar receptor affinities (apparent Kds: 5 × 10−8 M and 8 × 10−8 M), while that of [Nle8,18,Tyr34]bPTH(7–34)amide was more than 10-fold lower (apparent Kd: 2 × 10−6 M). These changes in biological properties suggest that high affinity receptor binding requires both amino- and carboxyl-terminal domains of the PTHrP(1–36) sequence and/or intramolecular interactions which are impaired by the D-Trp substitution for Gly12.  相似文献   

8.
The kinetics of several processes involving the potential antioxidant role of urate in physiological systems have been investigated by pulse radiolysis. While the monoanionic urate radical, ·UH-, can be produced directly by oxidation with ·Br-2 or ·OH, it can also be generated by oxidation with the neutral tryptophan radical, ·Trp, with a rate constant of 2 × 107 M-1s-1. This radical, ·UH-, reacts with ·O-2 with a rate constant of 8 × 108 M-1s-1. Also, ·UH- is reduced by flavonoids, quercetin and rutin in CTAB micelles at rate constants of 6 × 106 M-1s-1 and 1 × 106 M-1s-1, respectively. These results can be of value by providing reference data useful in further investigation of the antioxidant character of urate in more complex biological systems.  相似文献   

9.
Both prostaglandins (PGs) and nitric oxide (NO) have cytoprotective and hyperemic effects in the stomach. However, the effect of NO on PG synthesis in gastric mucosal cells is unclear. We examined whether sodium nitroprusside (SNP), a releaser of NO, stimulates PG synthesis in cultured rabbit gastric mucus-producing cells. These cells did not release NO themselves. Co-incubation with SNP (2 × 10−4, 5 × 10−4, 10−3 M) increased PGE2 synthesis, and SNP (10−3 M) increased PGI2 synthesis in these cells. Hemoglobin, a scavenger of NO, (10−5 M) eliminated the increase in PGE2 synthesis by SNP, but methylene blue, an inhibitor of soluble guanylate cyclase, (5 × 10−5 M) did not affect the increase in PGE2 synthesis by SNP. 8-bromo guanosine 3′ : 5′-cyclic monophosphate (8-bromo cGMP), a cGMP analogue, (10−6, 10−5, 10−4, 10−3 M) did not affect PGE2 synthesis. These findings suggest that NO increased PGE2 and PGI2 synthesis via a cGMP-independent pathway in cultured rabbit gastric cells.  相似文献   

10.
Aqueous solutions of fractions of an extracellular linear mannan formed by Rhodotorula rubra yeast have been investigated by hydrodynamic methods (high-speed sedimentation, translation isothermic diffusion and viscometry). The molecular weight was determined according to Svedberg ( ) and the polydispersity parameters of the initial sample were also determined (Mw/Mn = 1·20 and Mz/Mw = 1·21). Relationships between the molecular weight (M) and so, Do and [η] in the range were: [η] = 2·33 × 10−2 M0.75, Do = 1·65 × 10−4 M0·58, so = 2·24 × 10−15 M0·43. The equilibrium rigidity and hydrodynamic diameter of chains representing mannan molecules were evaluated.  相似文献   

11.
Microbial characterization during composting of municipal solid waste   总被引:29,自引:0,他引:29  
This study investigates the prevailing physico-chemical conditions and microbial community; mesophilic bacteria, yeasts and filamentous fungi, bacterial spores, Salmonella and Shigella as well as faecal indicator bacteria: total coliforms, faecal coliforms and faecal Streptococci, present in a compost of municipal solid waste. Investigations were conducted in a semi-industrial pilot plant using a moderate aeration during the composting process. Our results showed that: (i) auto-sterilization induced by relatively high temperatures (60–55°C) caused a significant change in bacterial communities. For instance, Escherichia coli and faecal Streptococci populations decreased, respectively, from 2×107 to 3.1×103 and 107 to 1.5×103 cells/g waste dry weight (WDW); yeasts and filamentous fungi decreased from 4.5×106 to 2.6×103 cells/g WDW and mesophilic bacteria were reduced from 5.8×109 to 1.8×107 bacteria/g WDW. On the other hand, the number of bacterial spores increased at the beginning of the composting process, but after the third week their number decreased notably; (ii) Salmonella disappeared completely from compost by the 25th day as soon as the temperature reached 60°C; and (iii) the bacterial population increased gradually during the cooling phase. While Staphylococci seemed to be the dominant bacteria during the mesophilic phase and at the beginning of the thermophilic phase, bacilli predominated during the remainder of the composting cycle. The appearance of gram-negative rods (opportunistic pathogens) during the cooling phase may represent a serious risk for the sanitary quality of the finished product intended for agronomic reuse. Compost sonication for about 3 min induced the inactivation of delicate bacteria, in particular gram-negatives. By contrast, gram-positive bacteria, especially micrococcus, spores of bacilli, and fungal propagules survived, and reached high concentrations in the compost.  相似文献   

12.
It is well recognized that estradiol (E2) is one of the most important hormones supporting the growth and evolution of breast cancer. Consequently, to block this hormone before it enters the cancer cell or in the cell itself, has been one of the main targets in recent years. In the present study we explored the effect of the progestin, nomegestrol acetate, on the estrone sulfatase and 17β-hydroxy-steroid dehydrogenase (17β-HSD) activities of MCF-7 and T-47D human breast cancer cells. Using physiological doses of estrone sulfate (E1S: 5 × 10−9 M), nomegestrol acetate blocked very significantly the conversion of E1S to E2. In the MCF-7 cells, using concentrations of 5 × 10−6 M and 5 × 10−5 M of nomegestrol acetate, the decrease of E1S to E2 was, respectively, −43% and −77%. The values were, respectively, −60% and −71% for the T-47D cells. Using E1S at 2 × 10−6 M and nomegestrol acetate at 10−5 M, a direct inhibitory effect on the enzyme of −36% and −18% was obtained with the cell homogenate of the MCF-7 and T-47D cells, respectively. In another series of studies, it was observed that after 24 h incubation of a physiological concentration of estrone (E1: 5 × 10−9 M) this estrogen is converted in a great proportion to E2. Nomegestrol acetate inhibits this transformation by −35% and −85% at 5 × 10−7 M and 5 × 10−5 M, respectively in T-47D cells; whereas in the MCF-7 cells the inhibitory effect is only significant, −48%, at 5 × 10−5 M concentration of nomegestrol acetate. It is concluded that nomegestrol acetate in the hormone-dependent MCF-7 and T-47D breast cancer cells significantly inhibits the estrone sulfatase and 17β-HSD activities which converts E1S to the biologically active estrogen estradiol. This inhibition provoked by this progestin on the enzymes involved in the biosynthesis of E2 can open new clinical possibilities in breast cancer therapy.  相似文献   

13.
We have designed and synthesized a series of small peptides containing a perfluoroalkyl ketone group at the C-terminal position of the angiotensin I sequence as inhibitors of human renin. From this series of compounds, 8 and 10 showed strong inhibition of human renin (IC50 = 3 × 10−9, 7 × 10−9 M, respectively). Compound 10 did not inhibit pepsin and cathepsin D at 10−4 M. Comparison of the IC50 of compound 8 and compound 11 (8.7 × 10−7 M) demonstrated the marked effect of the perfluoropropyl group on the potency of inhibition on renin, presumably due to the strong electron-withdrawing effect causing the ketone in 8 to exist predominantly as the hydrate — thus mimicking the tetrahedral transition state during hydrolysis of the scissile Leu10—Val11 amide bond.  相似文献   

14.
The binding of herbicides to the phylloquinone-(primary electron acceptor A1)-binding site in green plant photosystem (PS) I reaction centers is shown. Dissociation constants (Kd) of various herbicides to the phylloquinone-binding site were estimated by analyzing their competitive inhibition of the reconstitution of the phylloquinone analogue, menadione (vitamin K3), to the phylloquinone-extracted spinach PS I particles. The phylloquinone-binding site was found to bind o-phenanthroline (Kd = 1.2 × 10−4 M), but only weak binding was observed with atrazine (Kd > 10−2 M), although both are known to bind specifically to the quinone-(QB)-binding site in reaction centers of purple photosynthetic bacteria or PS II. The inhibitors of the cytochrome b/c1(ƒ) complex, myxothiazol (Kd=9.5 × 10−6 M) or antimycin A (Kd = 2.8 × 10−6 M), also strongly bound to the phylloquinone site. This is the first report showing that the PS I reaction center complex also has a herbicide-binding site, although the site is probably not sensitive in vivo to these herbicides due to its higher affinity for phylloquinone than herbicides. The inhibitor specificity of the PS I phylloquinone site is different from that of the other quinone-functioning sites in the photosynthetic or respiratory electron-transfer chain, suggesting it to have a unique structure.  相似文献   

15.
Experimental evidence is provided that selenomethionine oxide (MetSeO) is more readily reducible than its sulfur analogue, methionine sulfoxide (MetSO). Pulse radiolysis experiments reveal an efficient reaction of MetSeO with one-electron reductants, such as e-aq (k = 1.2 × 1010M-1s-1), CO·-2 (k = 5.9 × 108 M-1s-1) and (CH3)2) C·OH (k = 3.5 × 107M-1s-1), forming an intermediate selenium-nitrogen coupled zwitterionic radical with the positive charge at an intramolecularly formed Se N 2σ/1σ* three-electron bond, which is characterized by an optical absorption with λmax at 375 nm, and a half-life of about 70 μs. The same transient is generated upon HO· radical-induced one-electron oxidation of selenomethionine (MetSe). This radical thus constitutes the redox intermediate between the two oxidation states, MetSeO and MetSe. Time-resolved optical data further indicate sulfur-selenium interactions between the Se N transient and GSH. The Se N transient appears to play a key role in the reduction of selenomethionine oxide by glutathione.  相似文献   

16.
The kinetics of O·-2 reaction with semi-oxidized tryptophan radicals in lysozyme, Trp·(Lyz) have been investigated at various pHs and conformational states by pulse radiolysis. The Trp·(Lyz) radicals were formed by Br·-2 oxidation of the 3-4 exposed Trp residues in the protein. At pH lower than 6.2, the apparent bimolecular rate is about 2 × 108M-1s-1; but drops to 8 × 107M-1s-1 or less above pH 6.3 and in CTAC micelles. Similarly, the apparent bimolecular rate constant for the intermolecular Trp·(Lyz) + Trp·(Lyz) recombination reaction is about (4-7 × 106M-1s-1) at/or below pH 6.2 then drops to 1.3-1.6 × 106M-1s-1 at higher pH or in micelles. This behavior suggests important conformational and/or microenvironmental rearrangement with pH, leading to less accessible semioxidized Trp· residues upon Br·-2 reaction. The kinetics of Trp·(Lyz) with ascorbate, a reducing species rather larger than O·-2 have been measured for comparison. The well-established long range intramolecular electron transfer from Tyr residues to Trp radicals-leading to the repair of the semi-oxidized Trp·(Lyz) and formation of the tyrosyl phenoxyl radical is inhibited by the Trp·(Lyz)+O·-2 reaction, as is most of the Trp·(Lyz)+Trp·(Lyz) reaction. However, the kinetic behavior of Trp·(Lyz) suggests that not all oxidized Trp residues are involved in the intermolecular recombination or reaction with O·-2. As the kinetics are found to be quite pH sensitive, this study demonstrates the effect of the protein conformation on O·-2 reactivity. To our knowledge, this is the first report on the kinetics of a protein-O·-2 reaction not involving the detection of change in the redox state of a prosthetic group to probe the reactivity of the superoxide anion.  相似文献   

17.
Low-molecular-weight chitosan were prepared using 85% phosphoric acid at different reaction temperatures and reaction time. At room temperature, the viscosity average-molecular weights (Mv) of chitosan decreased to 7.1×104 from 21.4×104 after 35 days treatment. The degradation rate decreased with increasing hydrolysis time. The yields of chitosan also continuously decreased from 68.4 to 40.2% after 35 days. At 40, 60 and 80 °C, the molecular weight decreased to 3.70×104, 3.50×104 and 2.00×104 on 8 h hydrolysis, respectively. The yields of chitosan remain at a high level compared with that at room temperature and were 86.5, 71.4 and 61.3% at 40, 60 and 80 °C treatment, respectively. The different reaction time gave chitosan with different molecular weights. At 60 °C, the molecular weight of products decreased to 7.40×104 from 21.4×104 within 4 h, then decreased slowly to 1.90×104 in 15 h. It was also found that the water-solubility of chitosan increased as the molecular weight decreased. Results show the changes in yields and molecular weight of chitooligomers were strongly dependent on the reaction temperature and reaction time.  相似文献   

18.
The microbiota of completely mixed soil slurry was acclimated with pentachlorophenol (PCP) or with a wood preservative mixture (WPM) containing several pollutants such as PCP and petroleum hydrocarbons. The impact of these compounds on the bacterial diversity was studied by using molecular tools. PCR amplifications of the 16S ribosomal RNA gene sequences (rDNA) were carried out with total DNA extracted from soil slurry samples taken at different time points during the enrichment process of the PCP and WPM reactors. The composition of these PCR products, reflecting the bacterial diversity, was monitored by the single-strand-conformation polymorphism (SSCP) method. Our results showed that the complexity of the SSCP profiles in the PCP reactor decreased significantly during the enrichment process, whereas they remained complex in the WPM reactor. PCR-amplified 16 rDNA libraries were generated from each reactor. The SSCP method was used to rapidly screen several clones of these libraries to find specific single-strand DNA migration profiles. In the PCP-activated soil, 96% of examined clones had the same SSCP profile, and sequences of representative clones were related to the genusSphingomonas, suggesting that the enrichment with PCP resulted in a selection of little phylogenetic diversity. Four different SSCP profiles were observed with the 68 examined clones from the WPM reactor. Representative clones of these profiles were related to Methylocystaceae or Rhizobiaceae, to sulfur-oxidizing symbionts, to the genusAcinetobacter, and to the genusSphingomonas. We also cloned and sequenced PCR-amplified DNA related to thepcpB gene, coding for theSphingomonas PCP-4-monooxygenase and detected in both reactors after two weeks of enrichment. Of the 16 examined clones, deduced amino acid sequences of 13 clones were highly related to theSphingomonas sp. strain UG30pcpB. The three remainingpcpB clones were not closely related to the three knownSphingomonas pcpB.  相似文献   

19.
Amperometric choline biosensors were fabricated by the covalent immobilization of an enzyme of choline oxidase (ChO) and a bi-enzyme of ChO/horseradish peroxidase (ChO/HRP) onto poly-5,2′:5′,2″-terthiophene-3′-carboxylic acid (poly-TTCA) modified electrodes (CPMEs). A sensor modified with ChO utilized the oxidation process of enzymatically generated H2O2 in a choline solution at +0.6 V. The other one modified with ChO/HRP utilized the reduction process of H2O2 in a choline solution at −0.2 V. Experimental parameters affecting the sensitivity of sensors, such as pH, applied potential, and temperature were optimized. A performance comparison of two sensors showed that one based on ChO/HRP/CPME had a linear range from 1.0×10−6 to 8.0×10−5 M and the other based on ChO/CPME from 1.0×10−6 to 5.0×10−5 M. The detection limits for choline employing ChO/HRP/CPME and ChO/CPME were determined to be about 1.0×10−7 and 4.0×10−7 M, respectively. The response time of sensors was less than 5 s. Sensors showed good selectivity to interfering species. The long-term storage stability of the sensor based on ChO/HRP/CPME was longer than that based on ChO/CPME.  相似文献   

20.
R.J.W. De Wit 《FEBS letters》1982,150(2):445-448
Folic acid is degraded too fast by Dictyostelium discoideum to study binding of this ligand to cell surface binding proteins. Folate deaminase activity was inhibited in the presence of 3.3 × 10−4 M 8-azaguanine. This inhibitor enabled us to detect two folate binding proteins. One type bound folic acid and deamino-folic acid with the same affinity (K0.5 = 3–6 × 10−7 M) and apparently negative cooperativity. Binding to only this type was observed if 8-azaguanine was omitted. The second type bound folic acid noncooperatively with Kd = 7 × 10−7 M. Deamino-folic acid did not compete even at a 1000-fold excess. This type may correspond to the chemotactic receptor.  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