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.
X-ray structure analysis, IR spectrometry, differential thermal analysis, and viscosimetry have been used to investigate the properties of novel five-component polyhydroxyalkanoates formed by short- and medium-chain-length monomers synthesized by the bacterium Wautersia eutropha B5786. As the molar fraction of hydroxyhexanoate contained in polyhydroxyalkanoates samples increased from 2.5 to 18.0 mol%, their degree of crystallinity decreased from 72 to 57%. The melting temperature of multicomponent polyhydroxyalkanoates (Tm) and their temperature for the onset of decomposition (Td) are lower than those of polyhydroxybutyrate, whose Tm is 168-170 degrees C and Td 260-265 degrees C. In multicomponent polymers (PHA(SC+MC)), both parameters decrease as the molar fraction of hydroxyhexanoate grows to 156 and 252 degrees C, respectively, in the range of hydroxyhexanoate content studied. Hydroxyhexanoate influences the physicochemical properties of polyhydroxyalkanoates similarly to hydroxyvalerate; as the fraction of either of these medium-chain-length monomers in polyhydroxyalkanoates increases, the crystallinity of the polymer decreases, but its thermostability remains unchanged.  相似文献   

2.
We developed a detailed mathematical model describing the coupling between the molecular weight distribution dynamics of poly(3-hydroxybutyrate-co-3hydroxyvalerate) (PHBV) copolymer chains with those of hydroxybutyrate (HB) and hydroxyvalerate (HV) monomer formation. Sensitivity analysis of the model revealed that both the monomer composition and the molecular weight distribution of the copolymer chains are strongly affected by the ratio between the rates at which the two-monomer units are incorporated into the chains. This ratio depends on the relative HB and HV availability, which in turn is a function of the expression levels of genes encoding enzymes that catalyze monomer formation. Regulation of gene expression was accomplished through the aid of an artificial genetic network, the patterns of expression of which can be controlled by appropriately tuning the concentration of an extracellular inducer. Extensive simulations were used to study the effects of operating conditions and parameter uncertainties on the range of achievable copolymer compositions. Since the predicted conditions fell in the range of feasible bioprocessing manipulations, it is expected that such strategy could be successfully employed. Thus, the presented model constitutes a powerful tool for designing genetic networks that can drive the formation of PHBV copolymer structures with desirable characteristics.  相似文献   

3.
This article describes a new process for the production of biopolymers (polyhydroxyalkanoates, PHAs) based on the aerobic enrichment of activated sludge to obtain mixed cultures able to store PHAs at high rates and yields. Enrichment was obtained through the selective pressure established by feeding the carbon source in a periodic mode (feast and famine regime) in a sequencing batch reactor. A concentrated mixture of acetic, lactic, and propionic acids (overall concentration of 8.5 gCOD L(-1)) was fed every 2 h at 1 day(-1) overall dilution rate. Even at such high organic load (8.5 gCOD L(-1) day(-1)), the selective pressure due to periodic feeding was effective in obtaining a biomass with a storage ability much higher than activated sludges. The immediate biomass response to substrate excess (as determined thorough short-term batch tests) was characterized by a storage rate and yield of 649 mgPHA (as COD) g biomass (as COD)(-1) h(-1) and 0.45 mgPHA (as COD) mg removed substrates (as COD(-1)), respectively. When the substrate excess was present for more than 2 h (long-term batch tests), the storage rate and yield decreased, whereas growth rate and yield significantly increased due to biomass adaptation. A maximum polymer fraction in the biomass was therefore obtained at about 50% (on COD basis). As for the PHA composition, the copolymer poly(beta-hydroxybutyrate/beta-hydroxyvalerate) with 31% of hydroxyvalerate monomer was produced from the substrate mixture. Comparison of the tests with individual and mixed substrates seemed to indicate that, on removing the substrate mixture for copolymer production, propionic acid was fully utilized to produce propionylCoA, whereas the acetylCoA was fully provided by acetic and lactic acid.  相似文献   

4.
The copolyester of 3-hydroxybutyrate and 3- hydroxyvalerate was synthesized from the combined carbon sources of glucose and sodium propionate by a filamentaion-defective mutant of Sphaerotilus natans, which is a typical filamentous bacterium often found in activated sludge. The 3-hydroxyvalerate content in the produced polymer increased with increasing concentrations of propionate. Cell growth and polyester synthesis were observed even when 0.6% sodium propionate was added to the medium, when the 3-hydroxyvalerate content in the polymer produced was about 60 mol%. The monomer composition of the copolymer was also varied by aeration conditions, time of propionate feeding, and cultivation time. This strain flocculated in accordance with cell growth, allowing rapid and convenient separation of the biomass from the culture fluid.  相似文献   

5.
Bacillus megaterium strain OU303A isolated from municipal sewage sludge was selected for the study of biosynthesis of polyhydroxybutyrate (PHB) and polyhydroxybutyrate-co-hydroxyvalerate P (HB-co-HV) copolymer. The strain yielded a maximum of 62.43% DCW polymer in the medium containing glycerol as carbon source, which was followed by 58.63% DCW polymer in glucose containing medium. We found that this strain was capable of producing 2.5% hydroxyvalerate copolymer from a single carbon substrate, glucose. The strain showed an increase in the amount of HV monomer content, when the precursor for the copolymer was included in the fermentation medium. The characterization of the biopolymers was carried out using FTIR, GC-MS, H1 NMR and DSC. This is the first report of B. megaterium strain producing HV copolymer, without the addition of any precursor in the fermentation medium.  相似文献   

6.
The thermal unfolding of Amy 34, a recombinant alpha-amylase from Bacillus halodurans, has been investigated using differential scanning calorimetry (DSC). The denaturation of Amy 34 involves irreversible processes with an apparent denaturation temperature (T(m)) of 70.8 degrees C at pH 9.0, with four transitions, as determined using multiple Gaussian curves. The T(m) increased by 5 degrees C in the presence of 100-fold molar excess of CaCl2 while the aggregation of Amy 34 was observed in the presence of 1000-fold molar excess of CaCl2. Increase in the calcium ion concentration from 1- to 5-fold molar excess resulted in an increase in calorimetric enthalpy (DeltaH(cal)), however, at higher concentrations of CaCl2 (up to 100-fold), DeltaH(cal) was found to decrease, accompanied by a decrease in entropy change (DeltaS), while the T(m) steadily increased. The presence of 100-fold excess of metal chelator, EDTA, resulted in a decrease in T(m) by 10.4 degrees C. T(m) was also decreased to 61.1 degrees C and 65.9 degrees C at pH 6.0 and pH 11.0, respectively.  相似文献   

7.
The oscillatory rheology of one rodlike and one semiflexible xanthan sample has been investigated as a function of temperature in the range of xanthan concentrations where the polymer forms a lyotropic liquid crystalline phase in aqueous NaCl solutions. Readily observed changes in the rheological observables at temperatures corresponding to phase boundaries permit construction of the biphasic chimney region of the temperature-composition phase diagram. The chimney region leans toward larger values of the polymer concentration with increasing temperature, presumably as a consequence of a reduction in the effective axial ratio of the helical polymer with increasing temperature. The results permit construction of plots of the rheological observables as a function of polymer concentration at temperatures T in the range 20 相似文献   

8.
Copolymers of (R)-3-hydroxybutyric acid (HB) and epsilon-caprolactone (CL) with a composition ranging from 28 to 81 mol % of HB were synthesized by transesterification of the corresponding homopolymers in solution in the presence of 4-toluenesulfonic acid. The copolyesters were characterized with regard to their molecular weights, thermal properties, molar compositions, and average block length of repeating units by gel permeation chromatography (GPC), differential scanning calorimetry, (1)H NMR, and (13)C NMR, respectively. Random and microblock copolymers could be obtained depending on experimental conditions, with weight-average molecular weights of up to 20,000. The glass transition temperature decreased from 2 to -42 degrees C as the CL content was increased from 0 to 72 mol %. The melting temperature (T(m)) of the PCL phase decreased from 70 to 46 degrees C as the HB content changed from 0 to 47 mol %, while the T(m) of the PHB phase decreased from 177 degrees C to 163 degrees C as the CL content changed from 0 to 72 mol %. Matrix-assisted laser desorption ionization time-of-flight mass spectra of GPC fractionated samples allowed us to ascertain that copolymers rich in HB units have mostly hydroxyl and carboxyl end groups, while copolymers rich in CL units have mostly tosyl and carboxyl end groups.  相似文献   

9.
Park DH  Kim BS 《New biotechnology》2011,28(6):719-724
High-yield production of polyhydroxyalkanoates (PHAs) by Ralstonia eutropha KCTC 2662 was investigated using soybean oil and γ-butyrolactone as carbon sources. In flask culture, it was shown that R. eutropha KCTC 2662 accumulated PHAs during the growth phase. The optimum carbon to nitrogen ratio (C/N ratio) giving the highest cell and PHA yield was 20 g-soybean oil/g-(NH(4))(2)SO(4). The 4-hydroxybutyrate (4HB) fraction in the copolymer was not strongly affected by the C/N ratio. In a 2.5-L fermentor, a homopolymer of poly(3-hydroxybutyrate) [P(3HB)] was produced from soybean oil as the sole carbon source by batch and fed-batch cultures of R. eutropha with dry cell weights of 15-32 g/L, PHA contents of 78-83 wt% and yields of 0.80-0.82 g-PHA/g-soybean oil used. By co-feeding soybean oil and γ-butyrolactone as carbon sources, a copolymer of poly(3-hydroxybutyrate-co-4-hydroxybutyrate) [P(3HB-co-4HB)] could be produced with dry cell weights of 10-21 g/L, yields of 0.45-0.56 g-PHA/g-soybean oil used (0.39-0.50g-PHA/g-carbon sources used) and 4HB fractions of 6-10 mol%. Higher supplementation of γ-butyrolactone increased the 4HB fraction in the copolymer, but decreased cell and PHA yield.  相似文献   

10.
Amphiphilic block copolymers were synthesized by transesterification of hydrophilic methoxy poly(ethylene glycol) (mPEG) and hydrophobic poly(propylene fumarate) (PPF) and characterized. Four block copolymers were synthesized with a 2:1 mPEG:PPF molar ratio and mPEGs of molecular weights 570, 800, 1960, and 5190 and PPF of molecular weight 1570 as determined by NMR. The copolymers synthesized with mPEG of molecular weights 570 and 800 had 1.9 and 1.8 mPEG blocks per copolymer, respectively, as measured by NMR, representing an ABA-type block copolymer. The number of mPEG blocks of the copolymer decreased with increasing mPEG block length to as low as 1.5 mPEG blocks for copolymer synthesized with mPEG of molecular weight 5190. At a concentration range of 5-25 wt % in phosphate-buffered saline, copolymers synthesized with mPEG molecular weights of 570 and 800 possessed lower critical solution temperatures (LCST) between 40 and 45 degrees C and between 55 and 60 degrees C, respectively. Aqueous solutions of copolymer synthesized with mPEG 570 and 800 also experienced thermoreversible gelation. The sol-gel transition temperature was dependent on the sodium chloride concentration as well as the mPEG block length. The copolymer synthesized from mPEG 570 had a transition temperature between 40 and 20 degrees C with salt concentrations between 1 and 10 wt %, while the sol-gel transition temperatures of the copolymer synthesized from mPEG molecular weight 800 were higher in the range 75-30 degrees C with salt concentrations between 1 and 15 wt %. These novel thermoreversible copolymers are the first biodegradable copolymers with unsaturated double bonds along their macromolecular chain that can undergo both physical and chemical gelation and hold great promise for drug delivery and tissue engineering applications.  相似文献   

11.
Jiang Z 《Biomacromolecules》2008,9(11):3246-3251
Candida antarctica lipase (CALB) has been successfully used as catalyst for copolymerization of dialkyl diester with diol and lactone to form aliphatic polyesters. The polymerization reactions were performed using a two stage process: first stage oligomerization under low vacuum followed by second stage polymerization under high vacuum. Use of the two-stage process is required to obtain products with high molecular weights at high yields for the following reasons: (i) the first stage reaction ensures that the monomer loss via evaporation is minimized to maintain 1:1 diester to diol stoichiometric ratio, and the monomers are converted to nonvolatile oligomers; (ii) use of high vacuum during the second stage accelerates equilibrium transesterification reactions to transform the oligomers to high molecular weight polymers. Thus, terpolymers of omega-pentadecalactone (PDL), diethyl succinate (DES), and 1,4-butanediol (BD) with a M w of whole product (nonfractionated) up to 77000 and M w/ M n between 1.7 and 4.0 were synthesized in high yields (e.g., 95% isolated yield). A desirable reaction temperature for the copolymerizations was found to be around 95 degrees C. At 1:1:1 PDL/DES/BD monomer molar ratio, the resultant terpolymers contained equal moles of PDL, succinate, and butylene repeat units in the polymer chains. (1)H and (13)C NMR analyses were used to determine the polyester microstructures. The synthesized PDL-DES-BD terpolymers possessed near random structures with all possible combinations of PDL, succinate, and butylene units via ester linkages in the polymer backbone. Furthermore, thermal stability and crystallinity of a pure PDL-DES-BD terpolymer with 1:1:1 PDL to succinate to butylene unit ratio and M w of 85400 were studied by thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC). The copolyester was found to be a semicrystalline material with a T g of -34 degrees C and a T m of 64 degrees C, which degrades in a single weight loss step centered at T max = 408 degrees C.  相似文献   

12.
Responses of broiler chickens to a high ambient temperature (35 degrees C) were measured in two experiments. In one experiment temperatures were increased abruptly from 21 degrees C to a daily range of 21-35 degrees C whereas, in the other, temperatures were increased more gradually over 6 days. The high temperatures were maintained for 5 h/day. In both experiments, birds exposed to the high temperatures ate less food and gained less liveweight than birds maintained at 21 degrees C. Efficiency of food conversion to liveweight gain and body composition were not affected by high temperature but there was a tendency for thyroid weight to decrease. Overall, the plasma concentration of triiodothyronine (T3) decreased and the plasma concentration of thyroxine (T4) increased, resulting in a decreased T3/T4 molar ratio, during exposure to high temperature. The concentration of plasma growth hormone, but not plasma reverse T3, was increased by high temperature. The initial responses to increased temperature were variable, with birds exposed more gradually adjusting relatively well until the maximum temperature was increased to 35 degrees C. All heated birds readjusted quickly to the daily reduction in temperature to 21 degrees C.  相似文献   

13.
Poly(epsilon-caprolactone) polyurethane and its shape-memory property   总被引:5,自引:0,他引:5  
Ping P  Wang W  Chen X  Jing X 《Biomacromolecules》2005,6(2):587-592
A series of segmented poly(epsilon-caprolactone) polyurethanes (PCLUs) were prepared from poly(epsilon-caprolactone) (PCL) diol, 2,4-toluene diisocyanate and ethylene glycol. The molecular weight (M(n)) of PCL was 500-10,000, and the soft-to-hard molar ratio was 1:2 to 1:6. Their shape-memory behaviors were investigated as a function of PCL molecular weight, PCLU composition, and thermal/mechanical history. When a deformation temperature 15-20 degrees C below T(m) was chosen, the lowest recovery temperature (LRT) was 15-18 degrees C below T(m), and the recovery ratio was 94-100% for tensile deformation of 300% and for compression of 2.7-fold. The reasons for this deformation-recovery procedure and the mechanism for this shape recovery below T(m) were discussed. The shape recovery was associated with the premelting of the crystals formed during the deformation and fixation, and, thus, it could be accomplished in the solid state. Its driving force was the inner stress stored in the deformed specimen during deformation and crystallization. Therefore, the LRT was a more practical temperature for shape-memory PCLU than T(m). It could be conveniently measured by means of thermal mechanical analysis. By adjusting the molecular weight of the PCL diol and the hard-to-soft ratio, the LRT of PCLU could be adjusted to the range of 37-42 degrees C, and reasonable rigidity could be retained after shape recovery, fulfilling the essential requirements of medical implantations.  相似文献   

14.
Circular dichroism measurements were used to study the binding of fd gene 5 protein to fd DNA, to six polydeoxynucleotides (poly[d(A)], poly[d(T)], poly[d(I)], poly[d(C)], poly[d(A-T)], and the random copolymer poly[d(A,T)]), and to three oligodeoxynucleotides (d(pA)20, d(pA)7, and d(pT)7). Titrations of these DNAs with fd gene 5 protein were generally done in a low ionic strength buffer (5 mM Tris-HCl, pH 7.0 or 7.8) to insure tight binding, needed to obtain stoichiometric endpoints. By monitoring the CD of the nucleic acids above 250 nm, where the protein has no significant intrinsic optical activity, we found that there were two modes of binding, with the number of nucleotides covered by a gene 5 protein monomer (n) being close to either 4 or 3. These stoichiometries depended upon which polymer was titrated as well as upon the protein concentration. Single endpoints at nucleotide/protein molar ratios close to 3 were found during titrations of poly[d(T)] and fd DNA (giving n = 3.1 and 2.8 +/- 0.2, respectively), while CD changes with two apparent endpoints at nucleotide/protein molar ratios close to 4 and approximately 3 were found during titrations of poly[d(A)], poly[d(I)], poly[d(A-T)], and poly[d(A,T)] (with the first endpoints giving n = 4.1 4.0, 4.0, and 4.1 +/- 0.3, respectively). Calculations showed that the CD changes we observed during these latter titrations were consistent with a switch between two non-interacting binding modes of n = 4 and n = 3. We found no evidence for an n = 5 binding mode. One implication of our results is that the Brayer and McPherson model for the helical gene 5 protein-DNA complex, which has 5 nucleotides bound per protein monomer (G. Brayer and A. McPherson, J. Biomol. Struct. and Dyn. 2, 495-510, 1984), cannot be correct for the detailed solution structure of the complex. We interpreted the CD changes above 250 nm upon binding of the gene 5 protein to single-stranded DNAs to be the result of a slight unstacking of the bases, along with a significant alteration of the CD contributions of the individual nucleotides in the case of A-and/or T-containing DNAs. Interestingly, CD contributions attributed to nearest-neighbor interactions in free poly[d(A-T)], poly[d(A,T)], poly[d(A)], and poly[d(T)] were partially maintained in the CD spectra of the protein-saturated polymers, so that neighboring nucleotides, when bound to the protein at 20 degrees C, appeared to interact with one another in much the same manner as in the free polymers at 50 degrees C.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Degradable shape-memory polymer networks intended for biomedical applications were synthesized from oligo[(epsilon-hydroxycaproate)-co-glycolate]dimethacrylates with glycolate contents between 0 and 30 mol % using a photopolymerization process. In addition AB copolymer networks were prepared by adding 60 wt % n-butyl acrylate as comonomer. All synthesized polymer networks are semicrystalline at room temperature. A melting transition T(m) between 18 and 53 degrees C which can be used as switching transition for the shape-memory effect can be attributed to the crystalline poly(epsilon-hydroxycaproate) phase. At temperatures below T(m) the elastic properties are dominated by these physical cross-links. At temperatures higher than T(m) the E modulus of the amorphous polymer networks is lowered by up to 2 orders of magnitude, depending on the chemical cross-link density. Copolymer networks based on macrodimethacrylates with a M(n) of up to 13,500 g x mol(-1) and a maximum glycolate content of 21 mol % show quantitative strain recovery rates in stress-controlled cyclic thermomechanical experiments. Hydrolytic degradation experiments of polymer networks performed in phosphate buffer solution at 37 degrees C show that the degradation rate can be accelerated by increasing the glycolate content and decelerated by the incorporation of n-butyl acrylate.  相似文献   

16.
The initial reaction rate and the thermostability of the mesophilic alcohol dehydrogenase (ADH) from Lactobacillus brevis (LBADH), and the thermophilic ADH from Thermoanaerobacter sp. (ADH T) in gas-phase reaction were compared. The effects of water activity, cofactor-to-protein molar ratio, and reaction temperature on the reduction of acetophenone to 1-phenylethanol were studied. An optimal water activity of 0.55 in terms of productivity was found for both ADHs. The cofactor-to-protein molar ratio was chosen slightly higher than equimolar to increase both activity and thermostability. An excellent optimal productivity of 1,000 g x L(-1) x d(-1) for LBADH and 600 g x L(-1) x d(-1)for ADH T was found at 60 degrees C, while the highest total turnover numbers with respect to the enzyme were achieved at 30 degrees C and amounted to 4.2 million for LBADH and 1.7 million for ADH T, respectively. Interestingly, the ADH from the mesophilic L. brevisshowed the higher thermostability in the nonconventional medium gas phase.  相似文献   

17.
P Tauc  C R Mateo    J C Brochon 《Biophysical journal》1998,74(4):1864-1870
The effects of hydrostatic pressure and temperature on the phase behavior and physical properties of the binary mixture palmitoyloleoylphosphatidylcholine/cholesterol, over the 0-40 molar % range of cholesterol compositions, were determined from the changes in the fluorescence lifetime distribution and anisotropy decay parameters of the natural lipid trans-parinaric acid (t-PnA). Pressurized samples were excited with a Ti-sapphire subpicosecond laser, and fluorescence decays were analyzed by the quantified maximum entropy method. Above the transition temperature (T(T) = -5 degrees C), at atmospheric pressure, two liquid-crystalline phases, alpha and beta, are formed in this system. At each temperature and cholesterol concentration below the transition pressure, the fluorescence lifetime distribution pattern of t-PnA was clearly modulated by the pressure changes. Pressure increased the fraction of the liquid-ordered beta-phase and its order parameter, but it decreased the amount of cholesterol in this phase. Palmitoyloleoylphosphatidylcholine/cholesterol phase diagrams were also determined as a function of temperature and hydrostatic pressure.  相似文献   

18.
The serum fraction of latex from Hevea brasiliensis, the para rubber tree, is known to contain an endo-chitinolytic enzyme, hevamine. Herein the activity of the rubber serum towards beta-chitin is investigated. The serum contained 6 mg/mL of protein and a chitinolytic activity of 18 mU permg of protein. The optimum ratio of enzyme to chitin was 0.22 mU/mg, and the optimum substrate concentration was 60 mg/mL. The optimum pH range was pH2-4, and the optimum temperature was 45 degrees C. At these conditions both (GlcNAc)2 and GlcNAc were produced in a molar ratio of approximately 2:1. The hydrolysis of 300 mg of chitin with 64 mU of the rubber serum for 8 days under the optimum conditions gave 39 mg of GlcNAc and 108 mg of (GlcNAc)2 as determined by HPLC. Mixing the rubber serum preparation with an Aspergillus niger pectinase preparation containing beta-N-acetylhexosaminidase can be used to produce almost exclusively the GlcNAc monomer in about 50% yield.  相似文献   

19.
Eu(2+), Dy(3+) and Tb(3+) co-doped strontium aluminate phosphor with high brightness and long afterglow was synthesized by a combustion method, using urea as a reducer. The properties of SrAl(2)O(4):Eu(2+),Dy(3+),Tb(3+) phosphor with a series of initiating combustion temperatures, urea concentrations and boric acid molar fractions were investigated. The sample at initiating combustion temperature of 600 degrees C exhibited an intense emission peak at 513 nm, in which the phosphor existed as a single-phase monoclinic structure. The experimental results showed that the optimum ratio of urea is 2.0 times higher than theoretical quantities and that the suitable molar fraction of H(3)BO(3) is 0.08. The average particle size of the phosphor was 50-80 nm and its luminescence properties were studied systematically. Compared with SrAl(2)O(4):Eu(2+),Dy(3+) phosphor, the initial luminescence brightness improved from 2.50 candela (cd)/m(2) to 3.55 cd/m(2) and the long afterglow time was prolonged from 1290 s to 2743 s.  相似文献   

20.
The purpose of this experiment was to characterize the high density lipoproteins (HDL) as a function of hydrated density. HDL was subfractionated on the basis of hydrated density by CsCl density gradient centrifugation of whole serum or the d 1.063-1.25 g/ml HDL fraction isolated from three men and three women. Apolipoprotein A-I and A-II quantitation by radial immunodiffusion showed that the A-I/A-II ratio varied with the lipoprotein hydrated density. The A-I/A-II molar ratio of HDL lipoproteins banding between d 1.106 and 1.150 g/ml was nearly constant at 2.2 +/- 0.2. In the density range 1.151-1.25 g/ml the A-I/A-II ratio increased as the density increased. On the other hand, in the density range between 1.077 and 1.105 the A-I/A-II ratio increased as the density decreased, ranging from 2.8 +/- 0.5 for the d 1.093-1.105 g/ml fraction to 5.6 +/- 1.3 for the d 1.077-1.082 g/ml fraction. The d 1.063-1.076 g/ml fraction and the d 1.077-1.082 g/ml fractions had comparable A-I/A-II ratios. Serum and the d 1.063-1.25 g/ml HDL fraction exhibited similar trends. The cholesterol/(A-I + A-II) ratio decreased as the density increased in all 12 samples (six serum and six HDL) examined. Gradient gel electrophoresis of the density gradient fractions showed that as the density increased from 1.063 to 1.200 g/ml the apparent molecular weight decreased from 3.9 x 10(5) to 1.1 x 10(5). HDL subfractions with the same hydrated densities had comparable molecular weights and A-I/A-II and cholesterol/(A-I + A-II) ratios when isolated from men or women. HDL contains subpopulations that differ in the A-I/A-II molar ratio.-Cheung, M. C., and J. J. Albers. Distribution of cholesterol and apolipoprotein A-I and A-II in human high density lipoprotein subfractions separated by CsCl equilibrium gradient centrifugation: evidence for HDL subpopulations with differing A-I/A-II molar ratios.  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