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.
Dynamic light scattering measurements were performed on dilute aqueous solutions of native ovalbumin (OA) and on those of linear OA aggregates induced by thermal denaturation at low ionic strength and neutral pH. The weight-average molecular weight Mw of four aggregates tested ranged from 1,700,000 to 5,500,000. The translational diffusion coefficient D0 of native OA at infinite dilution was estimated as 8.70 × 10 ?7 cm2/s, which gave 56.0 Å as the diameter of the rigid spherical particle. The intensity autocorrelation function of linear OA polymers was analyzed with the cumulant method to obtain the first cumulant Λe. The dependence of Λe on the scattering vector q at very low polymer concentration was found intermediate between those of a flexible chain and a rigid rod. The translational diffusion coefficient Dtr [≡ (Te/q2)q → 0] was in proportion to M, and the magnitude was in good agreement with a value calculated from the wormlike cylinder model with values of three parameters determined in an earlier study, ML = 1600 Å?1, d = 120 Å, and Q = 230 Å, where ML, d, and Q are the molecular weight per unit length, diameter, and persistence length, respectively. Based on these results, a new model, to be called as the dimer model, was proposed to interpret the formation mechanism of linear OA polymers induced by thermal denaturation. © 1993 John Wiley & Sons, Inc.  相似文献   

2.
Summary In our preliminary subcellular localization experiment we demonstrated that annexin II co-localized with submembranous actin in subpopulations of both cultured fibroblasts and keratinocytes. To investigate the physical interaction between annexin II and actin at the cell periphery, in vitro reconstitution experiments were carried out with keratins used as a control. Annexin II, isolated by immunoaffinity column chromatography, was found to exist as globular structures measuring 10 to 25 nm in diameter by rotary shadowing, similar to a previous report. We believe that these structures represent its polymeric forms. By negative staining, monomeric annexin II was detectable as tapered rods, measuring 6 nm in length and 1 to 2 nm in diameter. When annexin II was mixed with actin in 3 mM piperazine-N, N-bis-2-ethanesulfonic acid (PIPES) buffer with 10 mM NaCl2, 2 mM MgCl2 and 0.1 mM CaCl2, thick twisting actin bundles formed, confirming previous reports. This bundling was much reduced when calcium was removed. In the presence of 5 mM ethylenediamine tetra-acetic acid (EDTA) in 5 mM tris, pH 7.2, keratins were found to form a network of filaments, which began to disassemble when the chelator was removed and became fragmented when 0.1 mM CaCl2 was added. Keratins under the same conditions did not fragment when annexin II was present. These results suggest that annexin II, in conjunction with Ca2+, may be involved in a flexible system accommodating changes in the membrane cytoskeletal framework at the cell periphery in keratinocytes.  相似文献   

3.
The physical dimensions of several (1 → 6) branched (1 → 3) -β-D -glucan samples obtained from different organisms and their derivatives have been studied by electron microscopy, light scattering measurements, viscometry, and gel permeation chromatography. The electron micrographs indicate that in most samples these biopolymers are adequately described as linear worm-like coils. A sample reconstituted from alkaline media appeared as a blend of the linear, circular, and aggregated polymer morphologies. The average mass per unit length, ML = Mw/Lw for the macroscopically linear samples, was estimated to be 2100 ± 200 g mol?1 nm?1. The parameter mL was determined from the contour lengths obtained by electron microscopy and the molecular weight by light scattering measurements. The observed ML was consistent with the triple-helical structure reported from x-ray diffraction studies and observed degree of side-chain substitution. From the molecular snapshots shown in the electron micrographs, the persistence lengths of these β-D -glucans were determined to be 140 ± 30 nm. The experimentally determined intrinsic viscosities were consistent with these estimates of ML and persistence length. Comparison of the molecular weight distributions obtained from gel permeation chromatography and those deduced from the electron micrographs indicates that number and weight average contour lengths are more reliable than z and z + 1 averages. © 1993 John Wiley & Sons, Inc.  相似文献   

4.
Quasielastic light scattering and electrophoretic light scattering experiments were performed on chicken erythrocyte polynucleosome solutions at various temperatures and ionic strengths. The apparent diffusion coefficient, Dapp, was found to depend on the scattering vector K. In general, Dapp can be described as a damped oscillatory function of K in the ionic strength range of 10 to 60 mM and over the temperature range of 10 to 40°C. Electrophoretic light scattering studies on total digest chromatin samples indicate the apparent charge on the polynucleosomes increases as the ionic strength is lowered from 10 to 1 mM. These data are interpreted in terms of fluctuations in the surface charge distribution of the polyion and subsequent inducement of an asymmetric distribution of small ions about the polyion. These fluctuation components lead to the formation of “clusters” of polyions.  相似文献   

5.
An NADP-specific glutamate dehydrogenase [L-glutamate: NADP+ oxidoreductase (deaminating), EC 1.4.1.4] from alkaliphilic Bacillus sp. KSM-635 was purified 5840-fold to homogeneity by a several-step procedure involving Red-Toyopearl affinity chromatography. The native protein, with an isoelectric point of pH 4.87, had a molecular mass of approximately 315 kDa consisting of six identical summits each with a molecular mass of 52 kDa. The pH optima for the aminating and deaminating reactions were 7.5 and 8.5, respectively. The optimum temperature was around 60°C for both. The purified enzyme had a specific activity of 416units/mg protein for the aminating reaction, being over 20-fold greater than that for deaminating reaction, at the respective pH optima and at 30°C. The enzyme was specific for NADPH (Km 44 μM), 2-oxoglutarate (Km 3.13 mM), NADP+ (Km 29 μM), and L-glutamate (Km 6.06 mM). The Km for NH4Cl was 5.96 mM. The enzyme could be stored without appreciable loss of enzyme activity at 5°C for half a year in phosphate buffer (pH 7.0) containing 2 mM 2-mercaptoethanol, although the enzyme activity was abolished within 20 h by freezing at ?20°C.  相似文献   

6.
G Cohen  H Eisenberg 《Biopolymers》1966,4(4):429-440
Solutions of calf thymus NaDNA in 0.2M NaCl and CsDNA in 0.2M CsCl (fragmented by sonic irradiation to a molecular weight of about 5 × 105) were examined by electron microscopy, light scattering, and viscosity measurements. Electron microscopy showed that the molecular fragments are rodlike in shape and that the distribution of their lengths is sharp (Lw/Ln ~ 1.06). The weight-average number of nucleotides Zw per DNA molecule derived from light scattering was found to be in very good agreement with the value Lw derived from electron microscopy. The z-average length Lz derived from light scattering at 25°C. was found to be about 89% of the corresponding value derived by electron microscopy. Similarly, the value of Lw derived from intrinsic viscosity at 25°C. on the basis of a rodlike model was found to be about 87% of the corresponding value derived by electron microscopy. It is concluded that the molecules are slightly flexible and do not, assume their full contour length in solution because of the disorienting effect of Brownian motion. The intrinsie viscosity was found to increase linearly with decreasing temperature toward a limiting value corresponding to the fully stretched length of the macromolecules at 0°K.: a reasonable value for the modulus of elasticity could be calculated. It was also found that, no differences in conformation could be delected in t he two systems CsDNA-0.2M CsCl and NaDNA-0.2M NaCl.  相似文献   

7.
Abstract: Electron probe x-ray microanalysis (EPMA) was used to measure water content (percent water) and dry weight elemental concentrations (in millimoles per kilogram) of Na, K, Cl, and Ca in axoplasm and mitochondria of rat optic and tibial nerve myelinated axons. Myelin and cytoplasm of glial cells were also analyzed. Each anatomical compartment exhibited characteristic water contents and distributions of dry weight elements, which were used to calculate respective ionized concentrations. Free axoplasmic [K+] ranged from ≈155 mM in large PNS and CNS axons to ≈120–130 mM in smaller fibers. Free [Na+] was ≈15–17 mM in larger fibers compared with 20–25 mM in smaller axons, whereas free [Cl?] was found to be 30–55 mM in all axons. Because intracellular Ca is largely bound, ionized concentrations were not estimated. However, calculations of total (free plus bound) aqueous concentrations of this element showed that axoplasm of large CNS and PNS axons contained ≈0.7 mM Ca, whereas small fibers contained 0.1–0.2 mM. Calculated ionic equilibrium potentials were as follows (in mV): in large CNS and PNS axons, EK = ?105, ENa = 60, and ECl = ?28; in Schwann cells, EK = ?107, ENa = 33, and ECl = ?33; and in CNS glia, EK = ?99, ENa = 36, and ECl = ?44. Calculated resting membrane potentials were as follows (in mV, including the contribution of the Na+,K+-ATPase): large axons, about ?80; small axons, about ?72 to ?78; and CNS glia, ?91. ECl is more positive than resting membrane potential in PNS and CNS axons and glia, indicating active accumulation. Direct EPMA measurement of elemental concentrations and subsequent calculation of ionized fractions in axons and glia offer fundamental neurophysiological information that has been previously unattainable.  相似文献   

8.
A sensitive HPLC assay has been developed to determine the concentration of 17-(allylamino)-17-demethoxygeldanamycin (AAG) in human plasma over the concentration range of 12.5 to 2500 nM (7.33 to 1465 ng/mL). After the addition of 1000 nM geldanamycin as the internal standard, 1 mL samples of human plasma were subjected to solid-phase extraction, via Bond-Elut C18 cartridges, followed by analysis using an isocratic reversed-phase HPLC assay with UV detection. A Phenomenex Kingsorb, 3 micron, C18, 150×4.60 mm column and a Phenomenex Security Guard pre-column, C18 (ODS, Octadecyl), were used to achieve separation. AAG and GM were monitored at 334 and 308 nm, respectively, on a Hewlett-Packard 1050 Diode-Array Detector. The mobile phase, run at a flow-rate of 1 mL/min, was composed of 50% (v/v) 25 mM sodium phosphate (pH 3.00) with 10 mM triethylamine and 50% acetonitrile. HPLC effectively resolved AAG with retention times of 14.60 0.54 min and the internal standard geldanamycin at 10.72±0.38 min (n=15). This assay was able to measure plasma concentrations of AAG, the lower limit of quantitation being 12.5 nM, at a starting dose of 10 mg/m2 infused intravenously over 1 h in a Phase I clinical trial in adult patients with solid tumors.  相似文献   

9.
Abstract: In the present study, we investigated the existence of a binding site for l -carnitine in the rat brain. In crude synaptic membranes, l -[3H]carnitine bound with relatively high affinity (KD = 281 nM) and in a saturable manner to a finite number (apparent Bmax value = 7.3 pmol/mg of protein) of binding sites. Binding was reversible and dependent on protein concentration, pH, ionic strength, and temperature. Kinetic studies revealed a Koff of 0.018 min?1 and a Kon of 0.187 × 10?3 min?1 nM?1. Binding was highest in spinal cord, followed by medulla oblongata-pons ≥ corpus striatum ≥ cerebellum = cerebral cortex = hippocampus = hypothalamus = olfactory bulb. l -[3H]Carnitine binding was stereoselective for the l -isomers of carnitine, propionylcarnitine, and acetylcarnitine. The most potent inhibitor of l -[3H]carnitine binding was l -carnitine followed by propionyl-l -carnitine. Acetyl-l -carnitine and isobutyryl-l -carnitine showed an affinity ~500-fold lower than that obtained for l -carnitine. The precursor γ-butyrobetaine had negligible activity at 0.1 mM. l -Carnitine binding to rat crude synaptic membrane preparation was not inhibited by neurotransmitters (GABA, glycine, glutamate, aspartate, acetylcholine, dopamine, norepinephrine, epinephrine, 5-hydroxytryptamine, histamine) at a final concentration of 0.1 mM. In addition, the binding of these neuroactive compounds to their receptors was not influenced by the presence of 0.1 mMl -carnitine. Finally, a subcellular fractionation study showed that synaptic vesicles contained the highest density of l -carnitine membrane binding sites whereas l -carnitine palmitoyltransferase activity was undetectable, thus excluding the possibility of the presence of an active site for carnitine palmitoyltransferase. This finding indicated that the localization of the l -[3H]carnitine binding site should be essentially presynaptic.  相似文献   

10.
Six samples of Na xanthan in 0.01M aqueous NaCl containing 0.01 N HCl (pH = 2) were studied by light scattering and viscosity. This study was motivated by the finding that the intrinsic viscosity [η] fairly sharply decreased when the pH of the solvent was lowered from about 6 to 2 by adding HCl to 0.01M aqueous NaCl in which Na xanthan dissolves as rigid dimers having a double-helical structure. The data for weight-average molecular weight, radius of gyration, and [η] showed that Na xanthan at pH = 2 remains a dimer behaving as a semiflexible chain. Data analysis in terms of known theories for unperturbed wormlike chains yielded 0.47 ± 0.02, 2.0 ± 0.6, and 68 ± 7 nm for the contour length h per main-chain residue, diameter d, and persistence length q of the dimer, respectively. these h and d values agreed with the pitch per main-chain residue and the diameter of the double helix of Na xanthan in 0.01 or 0.1M aqueous NaCl. However, the q value, which was close to the intrinsic persistence length q0 ( = q in the absence of electrostatic interaction) of Na xanthan at pH = 2, was much smaller than the q0 (106 nm) of this helix. We concluded that the xanthan dimer at pH = 2 assumes a double-helical structure, which is geometrically the same as, but is more flexible than, that at neutral pH.  相似文献   

11.
There are conflicting opinions on the relative importance of photosynthetic limitations under salinity. Quantitative limitation analysis of photosynthesis provides insight into the contributions of different photosynthetic limitations, but it has only been applied under saturating light conditions. Using experimental data and modelling approaches, we examined the influence of light intensity on photosynthetic limitations and quantified the osmotic and ionic effects of salinity on stomatal (LS), mesophyll (LM), biochemical (LB) and light (LL) limitations in cucumber (Cucumis sativus L.) under different light intensities. Non‐linear dependencies of LS, LM and LL to light intensity were found. Osmotic effects on LS and LM increased with the salt concentration in the nutrient solution (Ss) and the magnitude of LM depended on light intensity. LS increased with the Na+ concentration in the leaf water (Sl) and its magnitude depended on Ss. Biochemical capacity declined linearly with Sl but, surprisingly, the relationship between LB and Sl was influenced by Ss. Our results suggest that (1) improvement of stomatal regulation under ionic stress would be the most effective way to alleviate salinity stress in cucumber and (2) osmotic stress may alleviate the ionic effects on LB but aggravate the ionic effects on LS.  相似文献   

12.
We extended the application of a sensitive high-performance liquid chromatography assay of amoxicillin developed in this laboratory for human plasma and middle ear fluid (MEF) to other sample matrices including chinchilla plasma or MEF and human and chinchilla whole blood with minor modification and validated the limit of quantitation at 0.25 μg/ml with a 50-μl sample size for human and chinchilla plasmas or MEFs. Amoxicillin and cefadroxil, the internal standard, were extracted from 50 μl of the samples with Bond Elut C18 cartridges. The extract was analyzed on a Keystone MOS Hypersil-1 (C8) column with UV detection at 210 nm. The mobile phase was 6% acetonitrile in 5 mM phosphate buffer, pH 6.5 and 5 mM tetrabutylammonium. The within-day coefficients of variation were 2.7–9.9 (n=4) and 1.7–7.2% (n=3) for chinchilla plasma and MEF samples, respectively; 2.8–8.1% (n=3) and 2.9–4.7% (n=3) for human and chinchilla whole blood, respectively. An alternative mobile phase composition for chinchilla plasma and MEF samples reduced the analysis time significantly.  相似文献   

13.
The stability of RNA hairpin loops containing A-U-G: An-U-G-Um   总被引:5,自引:0,他引:5  
E Wickstrom  I Tinoco 《Biopolymers》1974,13(11):2367-2383
RNA oligomers with the sequence An-U-G-Um, n = 7–9, m = 5–10, have been synthesized and found to form hairpin loops in 21 mM or 101 mM Na+. The hairpin loops displayed melting temperatures 13°–29°C greater than that of the hairpin loop A6-C8-U6 in the same solvent. The increased stability of these hairpin loops is attributed to the presence of the trinucleotide A-U-G in the loop. Circular dichroism (CD) spectra were taken of the hairpin loops A7-U-G-U6,7,8, A8-U-G-U6,7,8, and A9-U-G-U6 in 21 mM Na+, and compared with circular dichroism spectra of A6-U6 in 1 M Na+. Difference spectra were calculated between each An-U-G-Um and 11.5 mM (nucleotide) A6-U6 at similar temperatures and identical singlestrand fractions to give the “experimental” CD spectrum of the unbase-paired nucleotides in the loop, assuming, five, six, or seven base pairs. CD spectra were calculated for each of the assumed unbase-paired sequences using the measured CD spectra of ApA, A-U-G-, A, and U, and compared with the experimental spectra. The best agreement was found for hairpin-loop models containing five base pairs and five to eight unbase-paired nucleotides in the loop.  相似文献   

14.
Y Tsunashima  K Moro  B Chu  T Y Liu 《Biopolymers》1978,17(2):251-265
Group-specific polysaccharides isolated by means of a cetavlon procedure are immunogenic in man and induce protective immunity against meningococcal meningitis. Minute quantities of the polymers in solution can act as vaccines. We now report the first characterization of a fractionated (C-1) group C polysaccharide in 0.4KM KCl and 0.05M sodium acetate by means of light-scattering spectroscopy. Independent measurements of refractive index increments, absolute scattered intensities, angular scattering intensities and line widths as a function of scattering angles and delay times at different concentrations using incident wavelengths of 632.8 nm from a He–Ne laser and of 488 nm from an argon–ion laser yield information on aggregation properties, molecular weight (Mr), radius of gyration 〈r0g1/2z, translational diffusion coefficient 〈D〉0z, and second virial coefficients A2 and B2 of C-1 polysaccharide. At relatively high ionic strength (0.04M KCl + 0.05M sodium acetate), we obtain for the C-1 polysaccharide in solution Mr = 5.15 × 105, 〈r2g1/2z = 345 Å, A2 = 1.25 × 10?4 ml/g, 〈D〉 = 1.16 × 10?7 cm2/sec with a corresponding Stokes radius of 240 Å and B2 = 4.4 ml/g. A2 and B2 are the second virial coefficients from intensity- and diffusion-coefficient measurements. The C-1 polysaccharide aggregates in solution and behaves hydrodynamically like random coils. Viscosity and sedimentation studies further confirm our conclusions that the fractioned C-1 polysaccharide aggregates in solution and EDTA can partially break up those aggregates. However, the system remains polydisperse even after adding an excess amount of EDTA. The weight-average molecular weight of the C-1 polysaccharide in solution depends upon ionic strength and exhibits a minimum at ~0.2M KCl. Finally, viscosity, light-scattering, and sedimentation results all show that the aggregated macromolecular system behaves like random-coiled polymers with no measurable shape factors.  相似文献   

15.
A sensitive reversed-phase high-performance liquid chromatographic method for the determination of atevirdine and its primary metabolite in human saliva or cerebrospinal fluid using solid-phase extraction is described. Samples mixed with internal standard and sodium phosphate buffer were applied to an activated C18 solid-phase extraction column. The reconstituted eluate was injected onto a Zorbax RX C8 column utilizing a mobile phase of 100 mM ammonium acetate (pH 4.0)–isopropyl alcohol–acetonitrile (55:20:25, v/v/v). Fluorescence detection was employed with excitation at 295 nm and emission at 456 nm. Quantitation was achieved using peak-height ratios. The detection response curve was linear from 2 to 850 nM for atevirdine in both human saliva and cerebrospinal fluid and from 2 to 250 nM for the metabolite in human saliva. The method was utilized to analyze cerebrospinal fluid and saliva samples from clinical studies.  相似文献   

16.
The secondary structure of xanthan in solutions of relatively low salt concentration and at room temperature has been investigated using static light scattering experiments. Additional evidence has been found for a dimeric structure at 25°C in 0.01M NaCl. From the experimental z-average mean square (ms) radius of gyration, a value for the persistence length p has been estimated, taking explicitly into account the polydispersity of the three samples used, which has been established by gel permeation chromatography (GPC) measurements. The experimental particle scattering functions of the three samples are consistent with theoretical estimates for polydisperse systems with the same value of p = 65 ± 10 nm and the molar mass per unit length for a dimeric structure. This secondary structure remains unaffected by the ionic strength in the 0.005–0.0lM range. Partial aggregation seems to occur at higher NaCl concentrations. Light scattering and GPC data show that heating the xanthan 0.01M NaCl solutions to about 70°C considerably reduces the Mw of the low molar mass sample (2.3 × 105-g·mol?1), contrary to what is observed for the high molar mass sample (1.8 × 106-g·mol?1). These experimental findings can be accounted for by a partial temperature-induced dissociation of the xanthan dimers according to an all-or-none mechanism. © 1994 John Wiley & Sons, Inc.  相似文献   

17.
A water soluble acidic heteropolysaccharide named WAF was isolated from Auricularia auricula‐judae by extracting with 0.9% NaCl solution. By using gas chromatography, gas chromatography‐mass spectrometry, and NMR, its chemical structure was determined to be composed of a backbone of α‐(1→3)‐linked D ‐mannopyranose residues with pendant side groups of β‐D ‐xylose, β‐D ‐glucose, or β‐D ‐glucuronic acid at position O6 or O2. Six fractions prepared from WAF with a weight‐average molecular mass (Mw) between 5.9 × 104 and 64.7 × 104 g/mol were characterized with laser light scattering and viscometry in 0.1M NaCl at 25°C. The dependence of intrinsic viscosity ([η]) and radius of gyration (Rg) on Mw for this polysaccharide were found to be [η] = 1.79 × 10?3Mw0.96 cm3 g?1 and Rg = 6.99 × 10?2 Mw0.54 nm. The molar mass per unit contour length (ML) and the persistence length (Lp) were estimated to be 1124 nm?1 and 11 nm, respectively. The WAF exhibited a semirigid character typical of linear polysaccharides. Molecular modeling was then used to predict the ordered and disordered states of WAF; the simulated ML and Lp were however much smaller than the experimental values. Taken altogether, the results suggested that WAF formed a duplex in solution. © 2010 Wiley Periodicals, Inc. Biopolymers 95: 217–227, 2011.  相似文献   

18.
The reproductive biology of the guitarfish Rhinobatos percellens was studied from 751 specimens caught by bottom pair trawlers off the coast of São Paulo, Brazil, between c. 24° 00′ S; 45° 15′ W and c. 25° 10′ S; 47° 52′ W, from September 2007 to August 2009. The total length (LT) and total mass (MT) relationship for males and females combined was MT = 1·29E‐06 LT3·15 (r = 0·99, n = 751). The mean LT of sexually mature specimens was 548 mm for males and 583 mm for females. Clasper growth was allometric and showed three distinct phases. Most claspers were calcified in specimens of c. 550 mm LT. The mean diameter of the largest oocyte was 29·8 mm, the mean ovarian fecundity was seven oocytes and ovulation occurred between August and November. Uterine fecundity ranged from two to 13 embryos (mean of five embryos). Larger females had higher litter sizes and larger embryos; the size‐at‐birth was c. 200 mm LT. The hepato‐somatic index oscillated seasonally for males and females; the gonado‐somatic index had little variation in males, but varied seasonally in females. The presence of many non‐pregnant adult females and of encapsulated eggs during two consecutive seasons suggests a resting period between gestations and the possibility of diapause.  相似文献   

19.
Complexation between lysozyme and sodium poly(2-acrylamido-2-methylpropanesulfonate) (PAMPS) was studied by quasielastic light scattering, electrophoretic light scattering, fluorescence, and turbidimetry in electrolyte solution. These techniques show that complexation occurs at pH 9.6 in an ionic strength buffer of 0.25M NaCl + 25 mM Na2B4O7. At constant lysozyme concentration (Cpro). The structure of the complex depends on the polymer concentration. At low polymer concentration (relative to Cpro), an intrapolymer complex is formed. This intrapolymer complex aggregates to an interpolymer species upon increase in polymer concentration. Complex formation was also studied by fluorescence using pyrene-labeled PAMPS (Py-PAMPS). Energy transfer from singlet-excited tryptophan residues in lysozyme to the pyrene label occurs when the complex is formed. Fluorescence and turbidity data indicate that lysozyme interacts with Py-PAMPS preferentially at pyrene sites, which leads to static quenching of tryptophan fluorescence via energy transfer to the pyrene label. © 1995 John Wiley & Sons, Inc.  相似文献   

20.
Laser light-scattering has been used to investigate the size of native proteoglycan aggregates (PGA-aA1) from day-8 chick limb-bud chondrocyte cultures isolated under associative extraction and purification conditions in 0.4M guanidinium chloride (GdnHCl) solution. Dynamic light-scattering measurements yielded a hydrodynamic radius, Rs, of 244 ± 10 nm for PGA-aA1 in 0.4M GdnHCl, and a weight-average molecular weight (M w) of 150 ± 50 × 106 was obtained from a Zimm plot. Disaggregation in 4.0M GdnHCl aqueous solution yielded proteoglycan subunits (PGS) with Rs = 39 ± 2 nm, M w = 1.6 ± 0.3 × 106, which reassembled in 0.4M GdnHCl to form “reconstituted native” aggregates (PGA-raA1) with Rs = 121 ± 6 nm, M w = 17 ± 3 × 106. A second specimen of PGA-aA1 had Rs = 192 ± 10 nm, M w = 100 ± 10 × 106. The latter value was estimated from an empirical relationship between M w and Rs. After dissociation, this specimen reassembled to form PGA-raA1 with Rs = 85 ± 5 nm, M w = 12 ± 1 × 106. These data are compared with those for a specimen of reconstituted aggregate (PGA-A1) that had been extracted under dissociative conditions and then reaggregated by dialysis to 0.4M GdnHCl aqueous solution, for which Rs = 138 ± 9 nm, M w = 45 ± 8 × 106. From these values, we have calculated the weight-average number of subunits per aggregate Nw: 111 for PGA-aA1 and 12 for raA1 (70 and 7 for the second PGA-aA1 and PGA-raA1 specimen, respectively) as compared to 32 for PGA-A1. The numbers of subunits per aggregate were also determined from electron micrographs of spread specimens. The latter results show the same trends as those obtained by light scattering, but lead in each case to lower numbers of subunits per aggregate. These data demonstrate conclusively that PGA samples exhibit a higher degree of aggregation in solution than visualized in typical electron microscopy (EM) preparations, probably due to disaggregation during EM specimen preparation. Since Nw determined both by light scattering (LS) and by EM are larger for native versus reconstituted aggregate samples, our data point to a more compact aggregation of subunits along the hyaluronic acid (HA) chains in the former.  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