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.
To verify the previous theoretical prediction that the disturbed flow distal to a stenosis enhances lipid accumulation at the blood/arterial wall interface, we designed a canine carotid arterial stenosis model and measured ex vitro the luminal surface concentration of bovine serum albumin (as a tracer macromolecule) by directly taking liquid samples from the luminal surface of the artery. The experimental results showed that due to the presence of a filtration flow, the luminal surface albumin concentration cw was higher than the bulk concentration co as predicted by our theory. The measurement revealed that the luminal surface concentration of macromolecules was indeed enhanced significantly in regions of the disturbed flow. At Re = 50, the relative luminal surface concentration cw/co was 1.66±0.10 in the vortex region, while the cw/co was 1.37±0.06 in the laminar flow region. When Re increased to 100,the cw/co in the vortex flow region and the laminar flow region reduced to 1.39±0.07 and 1.24±0.04,respectively. The effect of the filtration rate, vw, on the luminal surface concentration of albumin was remarkably apparent. At Re=50 and 100, when vw=8.9±1.7×10-6 cm/s, cw in the vortex region was 77% and 52% higher than co respectively, meanwhile when vw = 4.8±0.6×10-6 cm/s, cw in the vortex region was only 66% and 39% higher than co respectively. In summary, the present study has provided further experimental evidence that concentration polarization can occur in the arterial system and fluid layer with highly concentrated lipids in the area of flow separation point may be responsible for the formation and development of atherosclerosis.  相似文献   

2.
In ruminants, the uptake of inorganic phosphate (Pi) across the intestinal mucosa epithelium by Na-dependent and Na-independent mechanisms is a main regulatory factor in P homeostasis. The aim of the study was to elucidate to which extent Na-independent mechanisms, including pH effects or composition of mucosal brush-border membranes, could be involved in positive stimulation of Pi absorptive processes seen under the P deficient condition. Therefore, luminal, surface and intracellular pH of the jejunal epithelial cells in control and P depleted goats were compared and biochemical analyses of membrane phospholipids in the apical membrane of the jejunal epithelium were performed. Dietary P depletion resulted in decreased plasma Pi levels. While pH in jejunal ingesta was not significantly changed, P depletion resulted in a significantly lower surface pH in the crypt region compared to control animals (7.62 ± 0.02 vs. 7.77 ± 0.04, n = 4, P < 0.01). Inhibition of apical Na+/H+-exchange resulted in an increase of the jejunal surface pH in P depleted animals by 0.07 ± 0.01 (n = 6, P < 0.01) and 0.05 ± 0.01 (n = 6, P < 0.01) for the villus and the crypt region, respectively. This increase were inversely correlated with the initial surface pH prior to inhibition. In contrast to surface pH, intracellular pH of the jejunal epithelium and the phospholipid composition of the apical jejunal membrane were not affected by P depletion. Although the data suggest the existence of a Na+/H+-exchange mechanism at the luminal surface of goat jejunum they do not support the hypothesis that adaptational processes of active Pi absorption from goat jejunum in response to low dietary P could be based on “non Pi transporter events”.  相似文献   

3.
The paper reports a study involving the use of Halomonas boliviensis, a moderate halophile, for co-production of compatible solute ectoine and biopolyester poly(3-hydroxybutyrate) (PHB) in a process comprising two fed-batch cultures. Initial investigations on the growth of the organism in a medium with varying NaCl concentrations showed the highest level of intracellular accumulation of ectoine (0.74 g L−1) at 10–15% (w/v) NaCl, while at 15% (w/v) NaCl, the presence of hydroxyectoine (50 mg L−1) was also noted. On the other hand, the maximum cell dry weight and PHB concentration of 10 and 5.8 g L−1, respectively, were obtained at 5–7.5% (w/v) NaCl. A process comprising two fed-batch cultivations was developed—the first culture aimed at obtaining high cell mass and the second for achieving high yields of ectoine and PHB. In the first fed-batch culture, H. boliviensis was grown in a medium with 4.5% (w/v) NaCl and sufficient levels of monosodium glutamate, NH4+, and PO43−. In the second fed-batch culture, the NaCl concentration was increased to 7.5% (w/v) to trigger ectoine synthesis, while nitrogen and phosphorus sources were fed only during the first 3 h and then stopped to favor PHB accumulation. The process resulted in PHB yield of 68.5 wt.% of cell dry weight and volumetric productivity of about 1 g L−1 h−1 and ectoine concentration, content, and volumetric productivity of 4.3 g L−1, 7.2 wt.%, and 2.8 g L−1 day−1, respectively. At salt concentration of 12.5% (w/v) during the second cultivation, the ectoine content was increased to 17 wt.% and productivity to 3.4 g L−1 day−1.  相似文献   

4.
5.
Biofilms of sulphate-reducing Desulfovibrio sp. EX265 were grown in square section glass capillary flow cells under a range of fluid flow velocities from 0.01 to 0.4 m/s (wall shear stress, τw, from 0.027 to 1.0 N/m2). In situ image analysis and confocal scanning laser microscopy revealed biofilm characteristics similar to those reported for aerobic biofilms. Biofilms in both flow cells were patchy and consisted of cell clusters separated by voids. Length-to-width ratio measurements (l c:w c) of biofilm clusters demonstrated the formation of more “streamlined” biofilm clusters (l c:w c=3.03) at high-flow velocity (Reynolds number, Re, 1200), whereas at low-flow velocity (Re 120), the l c:w c of the clusters was approximately 1 (l c:w c of 1 indicates no elongation in the flow direction). Cell clusters grown under high flow were more rigid and had a higher yield point (the point at which the biofilm began to flow like a fluid) than those established at low flow and some biofilm cell aggregates were able to relocate within a cluster, by travelling in the direction of flow, before attaching more firmly downstream. Received 01 February 2002/ Accepted in revised form 16 July 2002  相似文献   

6.
The use of rotating flow in an annulus is investigated as a means of enhancing the yield of glucose and xylose in the acid hydrolysis of cellulosic slurries. A one-dimensional model of such a cyclone reactor is developed for flow cases, co-current and counter-current flow. For the case of 250°C, 1% w/w acid, the one-dimensional model indicates an increase in the maximum glucose yield from 48.1% in a plug flow reactor to 69.3% in a co-current cyclone reactor, and up to 81.0% in a countercurrent cyclone reactor. The corresponding xylose yields are 91.6% for co-current operation and 97.7% for countercurrent operation. In the co-current case the maximum glucose and xylose yields do not occur at the same location in the reactor; however, in the countercurrent case they do. Although product yields are dramatically improved over those obtained in a plug flow reactor, the product concentrations are lower than would typically be obtained in a plug flow reactor.List of Symbols A cm2 cross sectional area perpendicular to radial flow - A c cm2 cross sectional area of slurry inlet - A c cm2 cross sectional area of steam inlet - A w cm2 cross sectional area of water inlet - C c concentration of cellulose as potential glucose (grams of potential glucose/cm3 of total stream) - C c * grams cellulose/cm3 of solids concentration of cellulose as potential glucose - C ginitial * grams glulose/cm3 of solids concentration of cellulose entering reactor - C g grams glucose/cm3 of total stream concentration of glucose - C g * grams glucose/cm3 of liquid stream concentration of glucose - C cinitial * grams cellulose/cm3 of liquid concentration of glucose entering reactor - C xn concentration of xylan as potential xylose (grams of potential xylose/cm3 of total stream) - C xs grams xyclose/cm3 of total stream concentration of nylose - d f dilution factor - dr cm radial increment - g cm/s2 gravitational acceleration - g * centrifugal acceleration proportionality constant - h cm height of cyclone reactor - j cm/s flux - K constant in general equation for vortex flow, Eq. (4.9) - k 1 1/s kinetic rate constant of cellulose hydrolysis - k a 1/s kinetic rate constant of xylan hydrolysis - k 2 1/s kinetic rate constant of glucose decomposition - k 2a 1/s kinetic rate constant of xylose decomposition - m vortex exponent - M steam g/s mass rate of steam addition at outer radius - M water g/s mass rate of cold water addition at outer radius - n cm3/s empirically determined settling parameter - Q cm3/s net volumetric flow in outward radial direction - Q tot cm3/s total volumetric flow through reactor - q c cm3/s volumetric flow of slurry feed - q s cm3/s volumetric flow of stream feed - q water cm3/s volumetric flow of cold water feed - r cm radial position - r c 1/s rate of cellulose hydrolysis - r g 1/s rate of glucose decomposition - r i cm inner radius - r o cm outer radius - r xn 1/s rate of xylan hydrolysis - r xs 1/s rate of xylose decomposition - s mom cm g/s2 inlet steam momentum - T bulk s bulk residence time in reactor - T °C reactor temperature - v c cm3/g specific volume of slurry feed - v s cm3/g specific volume of steam - v w cm3/g specific volume of water - V f cm/s velocity of liquid as a function of radius - V i cm/s inlet velocity - V s cm/s velocity of solids as a function of radius - V steam cm/s inlet steam velocity to cyclone - V cm/s terminal settling velocity - V q cm/s tangential velocity - w mom cm g/s2 water inlet momentum - Y grams product out/grams reactant in yield of product - solids volumetric fraction - f solids volumetric fraction in slurry feed - i initial solids volumetric fraction of slurry - Pi  相似文献   

7.
In this study, we have formulated chitosan-coated sodium alginate microparticles containing meloxicam (MLX) and aimed to investigate the correlation between in vitro release and in vivo absorbed percentages of meloxicam. The microparticle formulations were prepared by orifice ionic gelation method with two different sodium alginate concentrations, as 1% and 2% (w/v), in order to provide different release rates. Additionally, an oral solution containing 15 mg of meloxicam was administered as the reference solution for evaluation of in vitro/in vivo correlation (ivivc). Following in vitro characterization, plasma levels of MLX and pharmacokinetic parameters [elimination half-life (t 1/2), maximum plasma concentration (C max), time for C max (t max)] after oral administration to New Zealand rabbits were determined. Area under plasma concentration–time curve (AUC0–∞) was calculated by using trapezoidal method. A linear regression was investigated between released% (in vitro) and absorbed% (in vivo) with a model-independent deconvolution approach. As a result, increase in sodium alginate content lengthened in vitro release time and in vivo t max value. In addition, for ivivc, linear regression equations with r 2 values of 0.8563 and 0.9402 were obtained for microparticles containing 1% and 2% (w/v) sodium alginate, respectively. Lower prediction error for 2% sodium alginate formulations (7.419 ± 4.068) compared to 1% sodium alginate formulations (9.458 ± 5.106) indicated a more precise ivivc for 2% sodium alginate formulation.  相似文献   

8.
Guy C. Fletcher 《Biopolymers》1976,15(11):2201-2217
Solutions of native collagen extracted from rat tail tendons in neutral salt solution have been studied by dynamic light scattering. The spectra obtained are consistent with the presence in solution of both single rod-shaped collagen molecules and aggregates of molecules. No contribution to the spectrum has been detected at any scattering angle from rotational diffusion of single molecules, although a measurable broadening effect is expected at high angles. The translational diffusion coefficient D of single molecules, calculated from the broader spectral component, shows an anomalous dependence on collagen concentration with a maximum value of D20,w = 8.6 ± 0.2 × 10?12 m2/sec near the concentration 0.04% by weight. Above 0.05% D falls linearly with increasing concentration and takes the value D 20,w = 8.1 ± 0.2 × 10?12 m2/sec at 0.064% collagen.  相似文献   

9.
Lutein is widely used as diet supplement for prevention of age-related macular degeneration. However, the application and efficacy of lutein in food and nutritional products has been hampered due to its poor solubility and low oral bioavailability. This study aimed to develop and evaluate the formulation of oral fast-dissolving film (OFDF) containing lutein nanocrystals for enhanced bioavailability and compliance. Lutein nanocrystals were prepared by anti-solvent precipitation method and then encapsulated into the films by solvent casting method. The formulation of OFDF was optimized by Box-Behnken Design (BBD) as follows: HPMC 2.05% (w/v), PEG 400 1.03% (w/v), Cremophor EL 0.43% (w/v). The obtained films exhibited uniform thickness of 35.64 ± 1.64 μm and drug content of 0.230 ± 0.003 mg/cm2 and disintegrated rapidly in 29 ± 8 s. The nanocrystal-loaded films with reconstituted particle size of 377.9 nm showed better folding endurance and faster release rate in vitro than the conventional OFDFs with raw lutein. The microscope images, thermograms, and diffractograms indicated that lutein nanocrystals were highly dispersed into the films. After administrated to SD rats, t max was decreased from 3 h for oral solution formulation to less than 0.8 h for OFDF formulations, and C max increased from 150 ng/mL for solution to 350 ng/mL for conventional OFDF or 830 ng/mL for nanocrystal OFDF. The AUC 0-24h of conventional or nanocrystal OFDF was 1.37 or 2.08-fold higher than that of the oral solution, respectively. These results suggested that drug nanocrystal-loaded OFDF can be applied as a promising approach for enhanced bioavailability of poor soluble drugs like lutein.  相似文献   

10.
At 7 days after first feeding (DAFF), the peptide hormone cholecystokinin (CCK) content (fmol individual?1) and the tryptic activity [μmol arginine‐methyl‐coumarinyl‐7‐amide (MCA) min?1 individual?1] per individual gut of Atlantic halibut Hippoglossus hippoglossus larvae were low: 0·2 ± 0·1 and 0·14 ± 0·10, respectively. Thereafter, both parameters increased with the increase in gut mass and reached 19·67 ± 5·58 and 2·71 ± 0·64 at 26 DAFF, respectively. Due to the small sample size, the dry mass (MG, mg) of the individual gut could not be determined accurately at 7 DAFF. At 13 DAFF MG represented 5·5% of whole body dry mass (Mw, mg) while at 26 DAFF it had increased to 23%. The mass specific tryptic activity [μmol MCA min?1 per mg dry mass (M)] in the gut increased from 2·74 ± 1 ± 98 at 13 DAFF to 5·00 ± 0·78 at 26 DAFF. There was more individual variation in the mass specific CCK content (fmol M?1) but no significant differences were found, although the data indicated an increase (from 23·38 ± 11·26 at 13 DAFF to 36·27 ± 8·96 fmol M?1 at 26 DAFF). At 7 DAFF the CCK content of the gut represented c. 2% of the whole body CCK content while it increased to c. 62% of the whole body CCK content at 26 DAFF. This demonstrates that it is necessary to separate neural and gastrointestinal sources of CCK in order to determine its alimentary role in fish larvae. Trypsin activity was only found in the gut compartment. In larvae aged 45 DAFF dietary proteins delivery into the gut by tube‐feeding appeared to stimulate post‐prandial secretion of CCK from the gut as well as stimulate pancreatic trypsin secretion, suggesting that both factors contribute to protein digestion.  相似文献   

11.
We have constructed an apparatus for the simultaneous measurement of electrophoretic mobility, μ, and diffusion coefficient, D, of macromolecules and cells. It combines band electrophoresis in a vertical, sucrose-gradient stabilized column, with quasielastic laser light-scattering determination of the diffusion coefficient of the species within the band. The entire electrophoresis cell is scanned through the laser beam of the quasielastic laser light-scattering apparatus by a vertical translation stage. Total intensity light-scattering measurement at each point in the cell gives the macromolecular concentration at that point. Solvent viscosity and electrical potential are measured at each point in the cell. Application of this apparatus to resealed red blood cell ghosts and to bovine hemoglobin indicates that measurements of field, viscosity, and migration distance are reliable, and that electroosmosis is insignificant. Application to T4D bacteriophage gives μ20,w = (?1.05 ± 0.05) × 10?4 cm2/V sec and D20,w = (3.35 ± 0.10) × 10?8 cm2/sec for fiberless particles, and μ20,w = ?(0.59 ± 0.03) × 10?4 cm2/V sec and D20,w = (2.86 ± 0.09) × 10?8 cm2/sec for whole phage with 6 fibers. Approximate analysis of these results with the Henry electrophoresis theory for spheres in dicates that each fiber contributes about 193 positive charges to the phage particle, compared with 327 from amino-acid analysis. The advantages and disadvantages of this apparatus, relative to conventional electrophoresis and to electrophoretic light scattering, are discussed.  相似文献   

12.
T. Raj  W. H. Flygare 《Biopolymers》1977,16(3):545-549
The translational diffusion coefficient of a pure sample of α-chymotrypsinogen A is measured by laser light scattering to give a value of D20,w0 = (8.40 ± 0.15) × 10?7 cm2/sec.  相似文献   

13.
A fish respirometer-metabolism chamber was used to obtain in vivo respiratory-cardiovascular and chloroethane gill flux data on transected channel catfish (Ictalurus punctatus). Methods used for spinal transection, attachment of an oral membrane (respiratory mast), placement and attachment of blood cannulas and urine catheters are described. Respiratory physiology, cardiac output and chemical extraction efficiencies for 1,1,2,2-tetrachloroethane (TCE), pentachloroethane (PCE), and hexachloroethane (HCE) were determined on 419–990 g catfish. The overall mean values (± s.d.) for ventilation volume (Qv), effective respiratory volume (Qw), oxygen consumption (Vo2 and percentage utilization of oxygen (U) were 17-3 ±4–71 h?1 kg?1, 9·8±l·71 h?1 kg?1, 71·6±12·5mg h?1 kg?1, and 49± 10%, respectively, while cardiac output calculated via the Fick Method was 2·4±0·61 h?1 kg?1. Additional measurements were made on ventilation rate (Vr), total plasma protein, haematocrit (Hct), and urine volume; while both arterial and venous blood were analysed for pH, oxygen partial pressure (P02), carbon dioxide partial pressure (Pco2), total oxygen (To2), total carbon dioxide (Tco2) and total ammonia (TAMM). Physiological measurements taken at 24 h were not significantly different from those taken at 48 h and indicated no deterioration of the in vivo preparation. All of these values agreed well with literature values on UTitransected channel catfish, except for Hct which was lower for cannulated animals used in this study. Overall, these data provide strong support for the use of transected channel catfish for in vivo collection of physiological and chemical gill flux data. The mean initial chemical extraction efficiencies for TCE, PCE and HCE were 41, 61 and 73%, respectively. Chemical clearances (ClX) for these same three chemicals were 5·9, 9·3 and 10·8 1 h?1 kg?1, respectively. The approximate 1: 1 relationship between effective respiratory volume (Qw) and chemical clearance (Clx) indicated that branchial uptake of PCE and HCE was water flow-limited. Chemical gill flux observed for channel catfish and chloroethanes was similar to that observed for rainbow trout in previous studies and provided further support for the flow-limited model of chemical flux across fish gills.  相似文献   

14.
Venous blood flow was measured for the first time in a cephalopod. Blood velocity was determined in the anterior vena cava (AVC) of cuttlefish S. officinalis with a Doppler, while simultaneously, ventilatory pressure oscillations were recorded in the mantle cavity. In addition, magnetic resonance imaging (MRI) was employed to investigate pulsatile flow in other major vessels. Blood pulses in the AVC are obligatorily coupled to ventilatory pressure pulses, both in frequency and phase. AVC peak blood velocity (vAVC) in animals of 232 (± 30 SD) g wet mass at 15°C was found to be 14.2 (± 7.1) cm s−1, AVC stroke volume (SVAVC) was 0.2 (± 0.1) ml stroke−1, AVC minute volume (MVAVC) amounted to 5.5 (± 2.8) ml min−1. Intense exercise bouts of 1–2 min resulted in 2.2-fold increases in MVAVC, enabled by 1.6-fold increments in both, AVC pulse frequency (f AVC) and vAVC. As increases in blood flow occurred delayed in time by 1.7 min with regard to exercise periods, we concluded that it is not direct mantle cavity pressure conveyance that drives venous return in this cephalopod blood vessel. However, during jetting at high pressure amplitude (> 1 kPa), AVC blood flow and mantle cavity pressure pulse shapes completely overlap, suggesting that under these conditions, blood transport must be driven passively by mantle cavity pressure. MRI measurements at 15°C also revealed that under resting conditions, f AVC and ventilation frequency (f V) match at 31.6 (± 2.1) strokes min−1. In addition, rates of pulsations in the cephalic artery and in afferent branchial vessels did not significantly differ from f AVC and f V. It is suggested that these adaptations are beneficial for high rates of oxygen extraction observed in S. officinalis and the energy conserving mode of life of the cuttlefish ecotype in general.  相似文献   

15.
A Malvern laser light-scattering instrument has been modified for use at scattering angles down to 5° and both total intensity and quasi-elastic scattering experiments. A sample of sheared, length-fractionated calf-thymus DNA was characterized by sedimentation, viscosity and electron microscopy. Quasi-elastic scattering and absolute intensity determinations were performed with the laser instrument and intensity determinations only with a Fica conventional light-scattering photometer. The total intensity experiments gave M?w = (3.75 ± 0.15) × 106 and 〈R21/2z = (206.9 ± 10.3) nm which yielded a value for the persistence length, allowing for polydispersity, of 66 ± 6nm. The quasi-elastic experiments at scattering angles below 20° gave D020, w = (2.23 ± 0.06) × 10?8 cm2/sec which combined with S020, w = 15.6 in the Svedberg equation gave M?w = (3.73 ± 0.18) × 106. In addition, from the higher angle data we extracted a value of the longest intramolecular relaxation time, τ1 of 17.5 msec. This is not in particularly good agreement with τ1 predicted by the Zimm–Rouse theory using our other experimental parameters. The disagreement may be due to the restricted applicability of the Zimm–Rouse spring-bead model as a quantitative representation of DNA molecules. Alternatively, it may be due to present difficulties in the unambiguous interpretation of molecular motions from the experimental autocorrelation functions.  相似文献   

16.
Electric birefringence measurements of suspensions of T3 and T7 bacteriophages in 10?2 M phosphate buffer, pH 6.9, show that there is a difference in their rotational diffusion coefficient. The values corrected to 25°C and water viscosity are D25,w = 4630 ± 130 sec?1 and D25,w = 5290 ± 260 sec?1 for T3 and T7, respectively. The value obtained from shell model calculations (according to Filson and Bloomfield) is D25,w = 4500 ± 600 sec?1. The apparent permanent dipole moments are 4.5 × 10?26 C·m and 1.7 × 10?26 C·m for T3 and T7, respectively. For both phage particles the intrinsic optical anisotropy is +7.2 × 10?3. It is shown that this anisotropy is mainly due to the DNA molecule inside the head of the phage. Its positive value means that there exists an excess orientation of the DNA helix perpendicular to the symmetry axis of the particle. For T7 an unexpectedly large increase of Δns and Ksp occurs at a glycerol concentration of about 30% (v/v). This increase is interpreted as being caused by a change of the shape of the particle and/or a change in the secondary structure of the DNA inside the head of the bacteriophage.  相似文献   

17.
The experiments and simulations reported in this paper show that, for stomata sensitive to both CO2 and water vapour concentrations, responses of stomatal conductance (gws) to boundary layer thickness have two components, one resulting from changes in intercellular CO2 concentration (χci) and another from changes in leaf surface water vapour saturation deficit (Dws). The experiments and simulations also show that the boundary layer conductance (gwb) can significantly alter the apparent response of gws to ambient air CO2 mole fraction (χca) and water vapour mole fraction (χwa). Because of the feedback loop involved the responses of gws for χca and χwa each include responses to both χci and Dws. The boundary layer alters the state of the variables sensed by the guard cells—i.e. χci and Dws—and so it is a source of feedback. Thus, when scaling up from responses of stomata to the response of gws for a whole leaf, the effect of the boundary layer must be considered. The results indicate that, for given responses of gws to χci and Dws, the apparent responses of gws to Dwa and χca depend on the size of the leaf and wind speed, showing that this effect of the boundary layer should be considered when comparing data measured under different conditions, or with different methods.  相似文献   

18.
19.
Strain ZZ-8T, a Gram-negative, aerobic, non-spore-forming, non-motile, yellow-pigmented, rod-shaped bacterium, was isolated from metolachlor-contaminated soil in China. The taxonomic position was investigated using a polyphasic approach. Phylogenetic analysis based on 16S rRNA gene sequences revealed that strain ZZ-8T is a member of the genus Flavobacterium and shows high sequence similarity to Flavobacterium humicola UCM-46T (97.2%) and Flavobacterium pedocola UCM-R36T (97.1%), and lower (<?97%) sequence similarity to other known Flavobacterium species. Chemotaxonomic analysis revealed that strain ZZ-8T possessed MK-6 as the major respiratory quinone; and iso-C15:0 (28.5%), summed feature 9 (iso-C17:1 w9c/C16:0 10-methyl, 22.9%), iso-C17:0 3-OH (17.0%), iso-C15:0 3-OH (8.9%), iso-C15:1 G (8.6%) and summed feature 3 (C16:1 w7c/C16:1 w6c, 5.7%) as the predominant fatty acids. The polar lipids of strain ZZ-8T were determined to be lipids, a glycolipid, aminolipids and phosphatidylethanolamine. Strain ZZ-8T showed low DNA–DNA relatedness with F. pedocola UCM-R36T (43.23?±?4.1%) and F. humicola UCM-46T (29.17?±?3.8%). The DNA G+C content was 43.3 mol%. Based on the phylogenetic and phenotypic characteristics, chemotaxonomic data and DNA–DNA hybridization, strain ZZ-8T is considered a novel species of the genus Flavobacterium, for which the name Flavobacterium zaozhuangense sp. nov. (type strain ZZ-8T?=?KCTC 62315 T?=?CCTCC AB 2017243T) is proposed.  相似文献   

20.
A water‐soluble α‐(1→4)‐D ‐glucan heteropolysaccharide with 37% degree of branch extracted by base from Rhizoma Panacis Japonici, coded as RPS3, was fractionated into six fractions by the method of nonsolvent addition. Their weight‐average molecular mass (Mw), polydispersity index (Mw/Mn), and radius of gyration (〈s2z1/2) were determined with laser light scattering (LLS) and size exclusion chromatography combined with LLS. The structure of the fraction was determined by methylation analyses and 13C NMR. The dependences of intrinsic viscosity ([η]) and 〈s2z1/2 on Mw were established as [η] = 0.71 Mw0.27 ± 0.01 (cm3/g) and 〈s2z1/2 = 1.53 Mw0.27 ± 0.02 (nm) in the Mw range from 5.62 × 104 to 3.05 × 106 (g/mol) for RPS3 in 0.15M NaCl aqueous solution at 25°C. On the basis of the current theory of the polymer solution, the fractal dimension (df), unperturbed chain dimension (A), and characteristic ratio (C) were calculated to be 3.0, 1.48 Å, and 15.1, respectively. The results revealed that the RPS3 chains existed as spherical conformation in the aqueous solution. Transmission electron microscope further provided the evidence of the sphere shape of the RPS3 and its fractionated molecules in water. In vitro cytotoxicity assay indicated that the fractions could inhibit the tumor cells and showed no harm to normal cells at low dose. The bioactivity was relative with molecular mass of the samples. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 383–390, 2010. This article was originally published online as an acceptedpreprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office atbiopolymers@wiley.com  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