首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Whole cells of alkaliphilic Bacillus pseudofirmus AR-199, induced for β-galactosidase activity, were used for the synthesis of 1-hexyl-β- -galactoside and 1-octyl-β- -galactoside, respectively, by transglycosylation reaction between lactose and the corresponding alcohol acceptor. The product yield was strongly influenced by the initial water content in the reaction mixture. Water content of 10% (v/v) was optimal providing 3.6–36 mM hexyl galactoside from 10 to 150 mM lactose, and no secondary product hydrolysis. Product yield could be enhanced by supplementing the reaction mixture with more cells or partly replacing the product with fresh substrate, but was decreased with time to the initial equilibrium level. Cell permeabilisation or disruption resulted in increased reaction rate and higher product yield but was followed by product hydrolysis. Octyl galactoside synthesis using whole cells was optimal at water content of 2% (v/v) with a yield of 26%. The cells were immobilised in cryogels of polyvinyl alcohol for use in continuous process, where hexyl galactoside was produced with a constant yield of 50% from 50 mM lactose for at least a week.  相似文献   

2.
The oxidation of melatonin (MEL) using the Cu(II) + H2O2 + HO (the Fenton-like reaction) system was investigated by chemiluminescence (CL), fluorescence, spectrophotometric, and EPR spin trapping techniques. The reaction exhibits CL in the 400–730 nm region. The light emission from the Fenton-like reaction was greatly enhanced in the presence of MEL and was strongly dependent on its concentration. The spectrum measured with cut-off filters revealed maxima at around 460, 500, 580–590, 640–650, and 690–700 nm. The band at 460 nm may be due to the excited cleavage product, N1-acetyl-N2-formyl-5-methoxykynuramine, whereas the bands at 500, 580–590, 640–650, and 700 nm were similar to those observed for singlet molecular oxygen (1O2). The effect of reactive oxygen species (ROS) scavengers on the light emission was studied. The CL was strongly inhibited by the 1O2 scavengers in a dose-dependent manner; at concentration 1 mM the potency of 1O2 scavenging was 5,5-dimethylcyclohexandione-1,3 > methionine > histidine > hydroquinone. The potency of HO scavenging by thiourea, tryptophan, cysteine at concentration 5 mM was 79–94%, by 1 mM glutathione and trolox 75 and 94%, respectively, and by 10 mM cimetidine 18%. Specific acceptors of O2 such as p-nitroblue tetrazolium chloride and 4,5-dihydroxy-1,3-benzene disulfonic acid (tiron) at concentration 5 mM decreased the CL by 51 and 95%, respectively, whereas superoxide dismutase (SOD) does not reduce the emission at concentration 2.8 U/ml. At higher concentration SOD substantially enhanced the light emission. Addition of 1360 U/ml catalase and 100 μM desferrioxamine strongly inhibited CL (96 and 90%, respectively). The increased generation of 1O2 from the Cu/H2O2 system in the presence of MEL was confirmed using the spectrophotometric method based on the bleaching of p-nitrosodimethylaniline and by trapping experiments with 2,2,6,6-tetramethylpiperidine (TEMP) and subsequent electron paramagnetic (EPR) spectroscopy. These findings suggest the increased production of reactive oxygen species (O2, HO, 1O2) from the Fenton-like reaction in the presence of MEL. This means that the hormone is not able to act as classical chain-breaking antioxidant even at low concentration, and may show clear prooxidant activity at higher concentrations. In addition, long-lived carbonyl product of the MEL transformation in the triplet state can also be toxic by transferring its energy to organelles and causing a photochemical process.  相似文献   

3.
Aims:  The aim of this work was to investigate the germination and inactivation of spores of Bacillus species in buffer and milk subjected to high pressure (HP) and nisin.
Methods and Results:  Spores of Bacillus subtilis and Bacillus cereus suspended in milk or buffer were treated at 100 or 500 MPa at 40°C with or without 500 IU ml−1 of nisin. Treatment at 500 MPa resulted in high levels of germination (4 log units) of B. subtilis spores in both milk and buffer; this increased to >6 logs by applying a second cycle of pressure. Viability of B. subtilis spores in milk and buffer was reduced by 2·5 logs by cycled HP, while the addition of nisin (500 IU ml−1) prior to HP treatment resulted in log reductions of 5·7 and 5·9 in phosphate buffered saline and milk, respectively. Physical damage of spores of B. subtilis following HP was apparent using scanning electron microscopy. Treating four strains of B. cereus at 500 MPa for 5 min twice at 40°C in the presence of 500 IU ml−1 nisin proved less effective at inactivating the spores of these isolates compared with B. subtilis and some strain-to-strain variability was observed.
Conclusions:  Although high levels of germination of Bacillus spores could be achieved by combining HP and nisin, complete inactivation was not achieved using the aforementioned treatments.
Significance and Impact of the Study:  Combinations of HP treatment and nisin may be an appealing alternative to heat pasteurization of milk.  相似文献   

4.
Density functional theory (DFT) computations at the B3LYP/Lanl2DZ level were used to elucidate the oxygen atom transfer (OAT) and coupled electron proton transfer (CEPT) reaction steps involved in the biomimetic catalytic cycle performed by polymer-supported MoVIO2(NN′)2 complexes [NN′ = phenyl-(pyrrolato-2-ylmethylene)-amine] with water as oxygen source, trimethyl-phosphane as oxygen acceptor and one-electron oxidising agents. The DFT method employed has been validated against experimental data [X-ray crystal structures of a NN′ ligand and a MoVIO2(NN′)2 complex as well as kinetic data]. The rate-limiting step in the forward-OAT from [MoVIO2] to PMe3 is the attack of PMe3 at an oxo ligand with ΔG (298 K) = 64.6 kJ mol−1. Dissociation of the product OPMe3 is facile with ΔG (298 K) = 26.3 kJ mol−1 giving a mono-oxo [MoIVO] complex which fills its coordination sphere with a further PMe3 substrate with ΔG (298 K) = 39.2 kJ mol−1. One-electron oxidation to a Mo(V) phosphane complex precedes the coordination of water/hydroxide. Additionally, the comproportionation of [MoVIO2] and [MoIVO] to dinuclear oxo-bridged [OMoV–O–MoVO] species has been calculated as the thermodynamic sink in this system and the back-OAT from dmso to mono-oxo [MoIVO] to give [MoVIO2] has been shown to involve an equilibrium between stereoisomeric [MoVIO2] complexes with an activation barrier of ΔG (298 K) = 113.1 kJ mol−1.  相似文献   

5.
Small (150–250 μm in diameter) and large (251–400 μm in diameter) preantral follicles (PFs) in sheep were cultured for 6 days in four different concentrations of transforming growth factor-alpha (TGF-), epidermal growth factor (EGF), FSH and LH. Proportions of follicles exhibiting growth, antrum formation and increase in follicular and oocyte diameter were the initial indicators of development. The ability of the oocytes isolated from these cultured follicles to mature to metaphase II (MII), after 24 h culture in a known in vitro maturation medium was the final criterion of success. TGF- 2.5 ng ml−1, EGF 50 ng ml−1 and FSH 1 and 2 μg ml−1 supported good initial growth of the PFs. Thirty and seventeen percent of the oocytes from the large PFs cultured in TGF- 2.5 ng ml−1 and FSH 2 μg ml−1 respectively, matured to the MII stage. These proportions for oocytes from small PFs were 11 and 6%, respectively. Oocytes from follicles cultured in EGF did not mature to the MII stage. LH at all concentrations tested and TGF-, EGF and FSH above 5, 50 ng ml−1 and 2 μg ml−1, respectively, induced degeneration of the PFs. It was concluded that (i) TGF- 2.5 ng ml−1 supports development of large PFs in sheep to obtain meiotically competent oocytes, (ii) PFs > 250 μm in initial diameter develop better in vitro, and (iii) in vitro development of sheep PFs could be obtained independent of gonadotropin stimulation.  相似文献   

6.
Knowledge of the mechanical behaviour of immature tracheae is crucial in order to understand the effects exerted on central airways by ventilatory treatments, particularly of Total Liquid Ventilation. In this study, a combined experimental and computational approach was adopted to investigate the compliance and particularly collapsibility of preterm lamb tracheae in the range of pressure likely applied during Total Liquid Ventilation (−30 to 30 cmH2O). Tracheal samples of preterm lambs (n=5; gestational age 120–130 days) were tested by altering transmural pressure from −30 to 30 cmH2O. Inflation (Si) and collapsing (Sc) compliance values were calculated in the ranges 0 to 10 cmH2O and –10 to 0 cmH2O, respectively. During the tests, an asymmetric behaviour of the ΔV/V0 vs. P curves at positive and negative pressure was observed, with mean Si=0.013 cmH2O−1 and Sc=0.053 cmH2O−1. A different deformed configuration of the sample regions was observed, depending on the posterior shape of cartilaginous ring. A three-dimensional finite-element structural model of a single tracheal ring, based on histology measurements of the tested samples was developed. The model was parameterised in order to represent rings belonging to three different tracheal regions (craniad, median, caudal) and numerical analyses replicating the collapse test conditions were performed to evaluate the ring collapsibility at pressures between 0 and −30 cmH2O. Simulation results were compared to experimental data to verify the model's reliability. The best model predictions occurred at pressures −30 to −10 cmH2O. In this range, a model composed of median rings best interpreted the experimental data, with a maximum error of 2.7%; a model composed of an equal combination of all rings yielded an error of 12.6%.  相似文献   

7.
The oxalate catalyzed iron(III) transfer from a trihydroxamate siderophore ferrioxamine B, [Fe(Hdfb)+], to ethylenediaminetetraacetic acid (H4edta) has been studied spectro-photometrically in weakly acidic aqueous solutions at 298 K and a constant 2.0 M ionic strength maintained by NaClO4. The results reveal that oxalate is a more efficient catalyst than the so far studied synthetic monohydroxamic acids. Any role of reduction of Fe(Hdfb)+ by oxalate in the catalysis has been rejected by the experimentally observed preservation of the oxalate concentration during the reaction time. Therefore, catalysis has been proposed to be a substitution based process. Under our experimental conditions Fe(Hdfb)+ is hexacoordinated and addition of oxalate results in the formation of Fe(H2dfb)(C2O4), Fe(H3dfb)(C2O4)2 and Fe(C2O4)3−3. Therefore, catalysis was proposed to be accomplished by the intermediate formation of the ternary and tris(oxalato) complexes. All three complexes react with H2edta2− to form thermodynamically stable Fe(edta) as a final reaction product. Whereas the formation of the ternary complexes is fast enough to feature a pre-equilibrium process to the iron exchange reaction, the formation of Fe(C2O4)3−3 is slow and is directly involved in the rate determining step of the Fe(edta) formation. Nonlinear dependencies of the rate constant on the oxalate and the proton concentrations have been observed and a four parallel path mechanism is proposed for the exchange reaction. The rate and equilibrium constants for the various reaction paths were determined from the kinetic and equilibrium study involving the desferrioxamine B- (H4dfb+), oxalate- and proton-concentration variations. The observed proton catalysis was attributed to the fast monoprotonation of ferrioxamine B as well as of the oxalate ligand. The observed catalysis of iron dissociation from the siderophore has been discussed in view of its significance with respect to in vivo microbial iron transport.  相似文献   

8.
The detailed engineering characterisation of shaken microtitre-plate bioreactors will enhance our understanding of microbial and mammalian cell culture in these geometries and will provide guidance on the scale-up of microwell results to laboratory and pilot scale stirred bioreactors. In this work computational fluid dynamics (CFD) is employed to provide a detailed characterisation of fluid mixing, energy dissipation rate and mass transfer in single well bioreactors from deep square 24-well and 96-well microtitre plates. The numerical predictions are generally found to be in good agreement with experimental observation of the fluid motion and measured values of the key engineering parameters. The CFD simulations have shown that liquid mixing is more intensive in 96-well than in 24-well bioreactors due to a significant axial component to the fluid velocity. Liquid motion is strongly dependent on the orbital shaking amplitude which generally has a greater impact than the shaking frequency. Average power consumptions of 70–100 W m−3 and 500–1000 W m−3, and overall mass transfer coefficient, kLa, values of 0.005–0.028 s−1 and 0.056–0.10 s−1 were obtained for 24-well and 96-well bioreactors respectively at an orbital shaking amplitude of 3 mm and shaking frequencies ranging from 500 rpm to 1500 rpm. The distribution of energy dissipation rates within each bioreactor showed these to be greatest at the walls of the well for both geometries. Batch culture kinetics of E. coli DH5 showed similar maximum specific growth rates and final biomass yields in shaken 24-well and shake flask bioreactors and in stirred miniature and 20 L bioreactors at matched kLa values. The CFD simulations thus give new insights into the local and overall engineering properties of microwell bioreactor geometries and further support their use as high throughput tools for the study and optimisation of microbial and mammalian cell culture kinetics at this scale.  相似文献   

9.
Nitrogen excreted as ammonium, urea, and dissolved primary amines (DPA), and nitrogen ingested by the planktonic calanoid copepod, Acartia tonsa, were measured while fed 4 foods with different N/C ratios in high (500 μg C l− 1) and low (50 μg C l− 1) concentrations. Adult copepods were fed the ciliate, Uronema marinum (N/C = 0.26), the diatom, Thalassiosira weissflogii, in log-phase growth (N/C = 0.20), and in senescent-phase growth (N/C = 0.12), and detritus derived from the saltmarsh grass, Spartina alterniflora, (N/C = 0.04). Total nitrogen excreted ranged from 0.06 to 0.18 μg N copepod− 1 d− 1 whereas nitrogen ingested exhibited considerably more variation (0.01 to 0.39 μg N copepod − 1d − 1). Ammonium was the dominant form of nitrogen excreted and was influenced by both food concentration and N/C ratio. Copepods fed foods with N/C ratios resembling their own body composition (log-phase diatoms and ciliates) excreted more ammonium when fed higher concentrations of food. In contrast, copepods fed foods with lower N/C ratios than their own body composition excreted more ammonium when fed lower concentrations of food, suggesting that they were catabolizing body protein for survival. Excretion of urea varied with food N/C ratio, with more urea excreted when the copepods were fed higher N/C foods. The excretion of DPA did not vary with either food concentration or food N/C ratio. Homeostasis serves to conserve the N/C ratio of copepods. Thus nitrogen excretion by healthy copepods should be expected to increase with ingestion only when copepods have high quantities of nitrogen-rich foods relative to the body composition of the copepods.  相似文献   

10.
The effects of temperature and concentration on the viscosity of orange peel pectin solutions were examined at five different temperatures between 20 and 60°C and five concentration levels between 2.5–20 kg/m3. The effects of temperature was described by an Arrhenius-type equation. The activation energy for viscous flow was in the range 19.53–27.16 kJ/mol, depending on the concentration. The effect of concentration was described by two types of equation, power-law and exponential. Equations were derived which describes the combined effects of temperature and concentration on the viscosity for two different models in the range of temperatures and concentrations studied. Orange peel pectin was extracted by using HCl (pH 2.5, 90°C, 90 min) ammonium oxalate (0.25%, pH 3.5, 75°C, 90 min) and EDTA (0.5%, 90°C, 90 min) extraction procedures. The best result was obtained with ammonium oxalate extraction in which the pectin content of the final product was 30.12%, although the efficiency among the procedures varied.The average molecular weight was measured by light scattering technique. Magnitudes of intrinsic viscosity and molecular weight of pectins obtained by extraction with HCl, ammonium oxalate and EDTA were 0.262, 0.281, 0.309 m3/kg and 84 500, 91 400, 102 800 kg/kgmol, respectively. The molecular weight dependence of the intrinsic viscosity of the orange peel pectin solutions was expressed by Mark–Houwink–Sakurada equation. The data were fitted to equation as ηi=2.34×10−5(Mw,ave)0.8224 which helps to evaluate the average molecular weight of pectin solutions from orange peel with a knowledge of their intrinsic viscosity.  相似文献   

11.
High-pressure liquid-chromatography and microcalorimetry have been used to determine equilibrium constants and enthalpies of reaction for the disproportionation reaction of adenosine 5′-diphosphate (ADP) to adenosine 5′-triphosphate (ATP) andadenosine 5′-monophosphate (AMP). Adenylate kinase was used to catalyze this reaction. The measurements were carried out over the temperature range 286 to 311 K, at ionic strengths varying from 0.06 to 0.33 mol kg−1, over the pH range 6.04 to 8.87, and over the pMg range 2.22 to 7.16, where pMg = -log a(Mg2+). The equilibrium model developed by Goldberg and Tewari (see the previous paper in this issue) was used for the analysis of the measurements. Thus, for the reference reaction: 2 ADp3− (ao) AMp2− (ao)+ ATp (ao), K° = 0.225 ± 0.010, ΔG° = 3.70 +- 0.11 kJ mol −1, ΔH° = −1.5 ± 1. 5 kJ mol −1, °S ° = −17 ± 5 J mol−1 K−1, and ACPp°≈ = −46 J mo1l−1 K−1 at 298.15 K and 0.1 MPa. These results and the thermodynamic parameters for the auxiliary equilibria in solution have been used to model the thermodynamics of the disproportionation reaction over a wide range of temperature, pH, ionic strength, and magnesium ion morality. Under approximately physiological conditions (311.15 K, pH 6.94, [Mg2+] = 1.35 × 10−3 mol kg−1, and I = 0.23 mol kg−1) the apparent equilibrium constant (KA′ = m(ΣAMP)m(ΣATP)/[ m(ΣADP)]2) for the overall disproportionation reaction is equal to 0.93 ± 0.02. Thermodynamic data on the disproportionation reaction and literature values for this apparent equilibrium constant in human red blood cells are used to calculate a morality of 1.94 × 10−4 mol kg−1 for free magnesium ion in human red blood cells. The results are also discussed in relation to thermochemical cycles and compared with data on the hydrolysis of the guanosine phosphates.  相似文献   

12.
The reaction of peroxynitrous acid with monohydroascorbate, over the concentration range of 250 μM to 50 mM of monohydroascorbate at pH 5.8 and at 25°C, was reinvestigated and the rate constant of the reaction found to be much higher than reported earlier (Bartlett, D.; Church, D. F.; Bounds, P. L.; Koppenol, W. H. The kinetics of oxidation of L-ascorbic acid by peroxynitrite. Free Radic. Biol. Med. 18:85–92; 1995; Squadrito, G. L.; Jin, X.; Pryor, W. A. Stopped-flow kinetics of the reaction of ascorbic acid with peroxynitrite. Arch. Biochem. Biophys. 322:53–59; 1995). The new rate constants at pH 5.8 are k1 = 1 × 106 M−1 s−1 and k−1 = 500 s−1 for 25°C and k1 = 1.5 × 106 M−1 s−1 and k−1 = 1 × 103 s−1 for 37°C. These values indicate that even at low monohydroascorbate concentrations most of peroxynitrous acid forms an adduct with this antioxidant. The mechanism of the reaction involves formation of an intermediate, which decays to a second intermediate with an absorption maximum at 345 nm. At low monohydroascorbate concentrations, the second intermediate decays to nitrate and monohydroascorbate, while at monohydroascorbate concentrations greater than 4 mM, this second intermediate reacts with a second monohydroascorbate to form nitrite, dehydroascorbate, and monohydroascorbate. EPR experiments indicate that the yield of the ascorbyl radical is 0.24% relative to the initial peroxynitrous acid concentration, and that this small amount of ascorbyl radicals is formed concomitantly with the decrease of the absorption at 345 nm. Thus, the ascorbyl radical is not a primary reaction product. Under the conditions of these experiments, no homolysis of peroxynitrous acid to nitrogen dioxide and hydroxyl radical was observed. Aside from monohydroascorbate's ability to “repair” oxidatively modified biomolecules, it may play a role as scavenger of peroxynitrous acid.  相似文献   

13.
Acylation reactions of naringin with palmitic acid were performed by a lipase after formation of highly concentrated homogeneous solutions. Their initial naringin concentration was 840–950 mM, which is 20–60 times greater than that in organic solvent media. The overall productivity of highly concentrated solutions was more than 15 times greater than those of organic phase media. The addition of DMSO (20–40%, w/w) to substrate mixtures lowered the melting temperature of a naringin–palmitic acid mixture (1:1 molar ratio) to about 40 °C. Reactions at 80 °C apparently followed Michaelis–Menten kinetics despite extremely high substrate concentrations. As the temperature increased from 60 °C to 80 °C, the apparent viscosity of the highly concentrated solution decreased remarkably from 4.31 Pa s to 0.063 Pa s. An activation energy of 7.65 kcal/mol obtained in a range of 60–75 °C suggests a diffusion-control. On the other hand, an activation energy of 17.09 kcal/mol in a range of 75–90 °C indicates a reaction-control. The highest product conversion yield of 33% (mol/mol) was obtained in a 10 h reaction at 80 °C. Addition of activated molecular sieves to the highly concentrated solution increased the product conversion yield by 7% (mol/mol), suggesting that the original equilibrium was disrupted by removing water and then a new equilibrium was reached.  相似文献   

14.
The bioconversion of propionitrile to propionamide was catalysed by nitrile hydratase (NHase) using resting cells of Microbacterium imperiale CBS 498-74 (formerly, Brevibacterium imperiale). This microorganism, cultivated in a shake flask, at 28 °C, presented a specific NHase activity of 34.4 U mgDCW−1 (dry cell weight). The kinetic parameters, Km and Vmax, tested in 50 mM sodium phosphate buffer, pH 7.0, in the propionitrile bioconversion was evaluated in batch reactor at 10 °C and resulted 21.6 mM and 11.04 μmol min−1 mgDCW−1, respectively. The measured apparent activation energy, 25.54 kJ mol−1, indicated a partial control by mass transport, more likely through the cell wall.

UF-membrane reactors were used for kinetic characterisation of the NHase catalysed reaction. The time dependence of enzyme deactivation on reaction temperature (from 5 to 25 °C), on substrate concentrations (from 100 to 800 mM), and on resting cell loading (from 1.5 to 200 μg  ml−1) indicated: lower diffusional control (Ea=37.73 kJ mol−1); and NHase irreversible damage caused by high substrate concentration. Finally, it is noteworthy that in an integral reactor continuously operating for 30 h, at 10 °C, 100% conversion of propionitrile (200 mM) was attained using 200 μg  ml−1 of resting cells, with a maximum volumetric productivity of 0.5 g l−1 h−1.  相似文献   


15.
Oleoyl ester of -ascorbic acid was synthesized by using immobilized lipases from Candida sp. A series of solvents, such as ethanol, tetrahydrofuran, pyridine, butanol, tertiary amyl alcohol (t-amyl alcohol), hexanol, octanol and hexane (log P from −0.24 to 3.5) were investigated for the reaction, and t-amyl alcohol was found to be the most suitable from the standpoint of the substrate concentration and the enzyme activity. And the equilibrium of the reaction was affected by the addition of the molecular sieves and the temperature. Reaction carried out at 55 °C and with 50 g/l of 4 Å molecular sieves is good for the enzyme to keep its activity and for making the equilibrium go to the product. The kinetic model was studied and the result showed that the reaction can be described by Ping-Pong mechanism. Parameters value of Vm and Km′ were obtained. Last, the pure products of the reaction were attained and determined by IR spectra, mass spectrometry and 1H NMR spectra.  相似文献   

16.
This paper investigates the behavior of crosslinked chitosans and metal-complexed crosslinked chitosans under similar hydrolytic conditions. Crosslinked chitosans with trimellitic anhydride, diisocyanatohexane, and dibromodecane as crosslinking agents under heterogenous reaction conditions were used as metal complexing agents by equilibrating them with metal salts such as ZnCl2, MnSO4, CuSO4, CdSO4, Pb(NO3)2, and HgCl2. Crosslinked chitosan without metal complexation had the same hydrolytic behavior as uncrosslinked chitosan. However, when the crosslinked chitosans were complexed with metals, their rates of hydrolysis and extent of hydrolysis were significantly reduced. Thus, while for chitosan about 840 μg/ml reducing sugar was produced in 4 h time, and 780 μg/ml was produced for diisocyanatohexane crosslinked chitosan, only 400 μg/ml and 320 μg/ml reducing sugars were produced for cadmium sulfate with crosslinked chitosan and diisocyanatohexane crosslinked chitosan, respectively. Similar results are obtained for other crosslinking agents. Studies on preincubation of the metal with the enzyme show that of the metals studied, Mn has no effect on preincubatioin with the enzyme, Hg, Cd, Pb, and Cu completely deactivates the enzyme, while Zn reduces the enzyme activity by about 43.3%. Preincubation of the metal salts with the chitosan shows that Hg and Cu completely deactivate the molecule from enzyme hydrolysis, Cd and Zn inactivate it to the extent of 56.8% and 43.3%, respectively, while Mn has no effect. Availability of the amino functions seems to be a key feature for the chitosanase to hydrolyze the chitosan polymer. This was also proved by the significant increase in the extent of hydrolysis for chitosan samples with 88% (final value 1120 μg/ml reducing sugar) and 85% deacetylation (final value 840 μg/ml reducing sugar). HPIC studies of the products show that a variety of oligomers are produced in the chitosanase enzyme hydrolytic reaction.  相似文献   

17.
Enzymes from extremophiles operate at conditions that are different from their ‘normal’ counterparts, and are therefore a useful extension of the enzyme toolbox. In this paper, the direct glucosylation reaction mediated by a hyperthermophilic β-glucosidase from Pyrocuccus furiosus was investigated. Hexanol was successfully coupled to glucose with this enzyme. A preliminary study was conducted to improve the product yield. A maximum product concentration of 12.9 g.l−1 was attainable by increasing the glucose concentration to the maximum solubility of 2000 g.(kg buffer solution)−1 at the reaction temperature. The highest glucose based yield of 2.64% was achieved with a glucose concentration of 900 g.(kg buffer solution)−1 at a reaction temperature of 65°C and a pH of 6.0. Performing the reaction at higher pH and temperature led to lower product concentrations. This was caused by deactivation of the enzyme accompanied by browning of the reaction mixture. A pH of 4.4 did have a negative effect on both the storage and the operational stability of the enzyme.  相似文献   

18.
Anti-allergic substances from the rhizomes of Dioscorea membranacea   总被引:2,自引:0,他引:2  
Extracts of five species of Thai medicinal plants, locally known as Hua-Khao-Yen, were screened for anti-allergic activities using RBL-2H3 cells. Of the five species studied, the ethanolic extract of Dioscorea membranacea exhibited potent inhibitory activity against β-hexosaminidase release as a marker of degranulation in RBL-2H3 cells, with an IC50 value of 37.5 μg/mL. Eight compounds were isolated from this crude ethanolic extract, [two naphthofuranoxepins (1, 2), one phenanthraquinone (3), three steroids (4–6), and two steroidal saponins (7, 8)], and tested for their anti-allergic activities. The results showed that dioscorealide B (2) possessed the highest activity with an IC50 value of 5.7 μM, followed by dioscoreanone (3, IC50 = 7.7 μM), dioscorealide A (1, IC50 = 27.9 μM), and diosgenin (9, IC50 = 29.9 μM). Structure–activity relationship studies of naphthofuranoxepins on anti-allergic activity revealed that the hydroxylation at position 8 conferred higher activity than methoxylation. For diosgenin derivatives, the aglycone was found to possess higher activity than the diglucosylated molecule; whereas substitution with rhamnoglucosides apparently results in loss of activity. Furthermore, effects of dioscorealide A, dioscorealide B, and dioscoreanone on antigen-induced release of TNF- and IL-4 in the late phase reaction were also examined.  相似文献   

19.
A method for measuring 5′-adenylic acid is described which is based on its capacity to stimulate the activity of phosphorylase a in the presence of low levels of glycogen and inorganic P. The method is useful with very low concentrations of AMP (0.02 to 0.5 μm) and is unaffected by much higher levels of ADP and ATP. Other common mononucleotides do not interfere at concentrations likely to be encountered. Yields of final product (NADPH) as high as 500 mol/mol of AMP can be obtained. Protocols are given for measuring as much as 5 × 10−10 mol at one extreme and as little as 10−15 mol at the other.  相似文献   

20.
The aim of the study was to investigate the incorporation of the antibacterial agent, miconazole nitrate into cyclodextrin cavities covalently bonded onto cloth fibers. The cellulosic fabric was grafted with β-cyclodextrin molecules through reaction with monochlorotriaziny β-cyclodextrin (MCT-β-CD). The suitable bonded reaction conditions were found to be MCT-β-CD 60–100 g/L, catalyst Na2CO3 50–60 g/L, the reaction temperature 150–160 °C and the reaction time 5–8 min.

The modified and unmodified fabrics were characterized by UV spectrophotometry. The level of miconazole nitrate entrapped in the fabrics were determined by HPLC and was founded to be much higher (0.458% w/w) for the textile functionalized with MCT-β-CD compared to the unmodified fabric (0.056% w/w). The antibacterial abilities measured by shaker flask method showed that the antibacterial property was markedly enhanced by impregnation with miconazole nitrate of the MCT-β-CD grafted textile. The finished fabric kept the antibacterial abilities more than 70% even after washing 10 times, while the antibacterial activity of the unmodified textile was almost lost.  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