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 62 毫秒
1.
Apolipoprotein A-I can spontaneously associate with phosphatidylcholine and cholesterol in 2-chloroethanol-water mixture. It was demonstrated, using a spin label technique, that dissolved molecules participate in complex formation. The apolipoprotein A-I/phosphatidylcholine/cholesterol complexes were isolated by gel chromatography. Complexes of three types were prepared and characterized: type A, large heterogeneous aggregates with molecular weight 600 000, sedimentation coefficient 10 S and the following molar composition - protein/phosphatidylcholine/cholesterol, 1:(70-100):(10-12); types B and C, with weight average molecular weights 140 000 and 110 000, average sedimentation coefficients 3.6 S and 1.7 S, respectively. Both types have the same molar composition - protein/phosphatidylcholine/cholesterol, 1:25:8. The dissimilar sedimentation coefficients between complexes B and C may be explained by the difference in the monomer/tetramer ratio (monomer molecular weight 50 000). The spin label sn-1-O-stearoyl-2-O-9'-spiro(4',4'-dimethyloxazolidine-3'-oxyl) heptadecanoylglycero-3-phosphocholine introduced into the complexes A and B showed different thermal properties of these complexes, which may be due to differences in the lipid-protein interactions.  相似文献   

2.
Reference scanners are used in dental medicine to verify a lot of procedures. The main interest is to verify impression methods as they serve as a base for dental restorations. The current limitation of many reference scanners is the lack of accuracy scanning large objects like full dental arches, or the limited possibility to assess detailed tooth surfaces. A new reference scanner, based on focus variation scanning technique, was evaluated with regards to highest local and general accuracy. A specific scanning protocol was tested to scan original tooth surface from dental impressions. Also, different model materials were verified. The results showed a high scanning accuracy of the reference scanner with a mean deviation of 5.3 ± 1.1 µm for trueness and 1.6 ± 0.6 µm for precision in case of full arch scans. Current dental impression methods showed much higher deviations (trueness: 20.4 ± 2.2 µm, precision: 12.5 ± 2.5 µm) than the internal scanning accuracy of the reference scanner. Smaller objects like single tooth surface can be scanned with an even higher accuracy, enabling the system to assess erosive and abrasive tooth surface loss. The reference scanner can be used to measure differences for a lot of dental research fields. The different magnification levels combined with a high local and general accuracy can be used to assess changes of single teeth or restorations up to full arch changes.  相似文献   

3.
Sedimentation velocity is a classical method for measuring the hydrodynamic, translational friction coefficient of biological macromolecules. In a recent study comparing various analytical ultracentrifuges, we showed that external calibration of the scan time, radial magnification, and temperature is critically important for accurate measurements (Anal. Biochem. 440 (2013) 81–95). To achieve accurate temperature calibration, we introduced the use of an autonomous miniature temperature logging integrated circuit (Maxim Thermochron iButton) that can be inserted into an ultracentrifugation cell assembly and spun at low rotor speeds. In the current work, we developed an improved holder for the temperature sensor located in the rotor handle. This has the advantage of not reducing the rotor capacity and allowing for a direct temperature measurement of the spinning rotor during high-speed sedimentation velocity experiments up to 60,000 rpm. We demonstrated the sensitivity of this approach by monitoring the adiabatic cooling due to rotor stretching during rotor acceleration and the reverse process on rotor deceleration. Based on this, we developed a procedure to approximate isothermal rotor acceleration for better temperature control.  相似文献   

4.
Sedimentation coefficients and apparent molecular masses of 5.8S rRNA from rat liver and yeast (Saccharomyces cerevisiae) depend considerably on the ionic strength and the kind of ions in solution. At 20°C the sedimentation coefficient of 5.8S rRNA in 10 mm sodium cacodylate, pH 7.0, amounts to 5.1 ± 0.2 S. By addition of NaCl up to 1.1 m the data increase reversibly to 6.1 ± 0.2 S (rat liver) or 5.4 ± 0.1 S (yeast) without significant changes of the molar mass (52 000 ± 2000) g/mol. Similar effects but with different extent were obtained using KCl or LiCl. These results can be explained by counterion effects on the conformation and changing of the water shell surrounding the RNA molecule. Short heat incubation (5 min at 65°C) and immediate cooling of rat liver 5.8S rRNA lead to dimer or oligomer formation. Its portions depend strongly on RNA concentration and are enhanced also with increasing NaCl concentration and incubation temperature as can be seen fro higher sedimentation coefficients and molecular masses as well as from additional bands in the electrophoretic pattern. At 20°C MgCl2 provokes, in concentrations up to 1.5 mm, a reversible increase of sedimentation coefficients of rat liver 5.8S rRNA to 6.65 ± 0.1 S whereas the molecular mass remains unchanged indicating strong Mg++ effects on conformation and/or water shell of the 5.8S rRNA. A further increase of sedimentation coefficients up to 8.2 ± 0.1 S combined with higher apparent molar masses up to 90 000 g/mol was observed in the presence of 30 to 50 mm MgCl2. In this concentration range of Mg++ the association constants of 5.8S rRNA dimerization increase from about 105to 3 × 107m?1. After removal of free Mg++ by addition of EDTA the 5.8S rRNA dimers dissociate if no incubation step at higher temperature in involved. The Mg++ induced 5.8S rRNA dimers differ in their stability from those formed by incubation at 65°C in the presence of higher concentrations of monovalent ions.  相似文献   

5.
The aggregation behavior of cobra venom (Naja naja naja) phospholipase A2 in the presence of lipids and Ca2+ was examined using ultracentrifugation and crosslinking techniques. Velocity sedimentation experiments were performed in sucrose gradients. The sedimentation coefficients of the cobra phospholipase A2 and various controls, including bovine serum albumin (BSA), malate dehydrogenase, carbonic anhydrase and pancreatic phospholipase A2, were calculated both in the presence and absence of ligands. The monomeric phospholipid, diheptanoylphosphatidylcholine, and the phospholipid analogue, dodecylphosphocholine (DPC), increased the sedimentation coefficient of the cobra phospholipase A2 from 2.2 S to 2.9 S, a value that is consistent with the formation of an enzyme dimer. The control proteins were unaffected by the presence of phospholipid, except for BSA, which apparently binds large amounts of DPC. Crosslinking experiments with glutaraldehyde showed that in the presence of diheptanoylphosphatidylcholine or DPC, the amount of crosslinked enzyme increased. Ca2+ had no effect on the aggregation state of the enzyme as measured by either technique. Both the ultracentrifugation data and crosslinking data are consistent with the hypothesis that the cobra venom phospholipase A2 exists as a dimer or higher-order aggregate in the presence of lipid substrate, although it is yet to be determined whether the functional subunit is a monomer, dimer or higher-order oligomer.  相似文献   

6.
Translation initiation factor 5B (IF5B) is required for initiation of protein synthesis. The solution structure of archaeal IF5B (aIF5B) was analysed by small-angle X-ray scattering (SAXS) and dynamic light scattering (DLS) and was indicated to be in both monomeric and dimeric form. Sedimentation equilibrium (SE) analytical ultracentrifugation (AUC) of aIF5B indicated that aIF5B forms irreversible dimers in solution but only to a maximum of 5.0-6.8% dimer. Sedimentation velocity (SV) AUC at higher speed also indicated the presence of two species, and the sedimentation coefficients s(20,w)(0) were determined to be 3.64 and 5.51±0.29 S for monomer and dimer, respectively. The atomic resolution (crystallographic) structure of aIF5B (Roll-Mecak et al. [6]) was used to model monomer and dimer, and theoretical sedimentation coefficients for these models were computed (3.89 and 5.63 S, respectively) in good agreement with the sedimentation coefficients obtained from SV analysis. Thus, the structure of aIF5B in solution must be very similar to the atomic resolution structure of aIF5B. SAXS data were acquired in the same buffer with the addition of 2% glycerol to inhibit dimerization, and the resultant monomeric aIF5B in solution did indeed adopt a structure very similar to the one reported earlier for the protein in crystalline form. The p(r) function indicated an elongated conformation supported by a radius of gyration of 37.5±0.2 ? and a maximum dimension of ~130 ?. The effects of glycerol on the formation of dimers are discussed. This new model of aIF5B in solution shows that there are universal structural differences between aIF5B and the homologous protein IF2 from Escherichia coli.  相似文献   

7.
8.
Sedimentation equilibrium (SE) analytical ultracentrifugation (AUC) is a gold standard for the rigorous determination of macromolecular buoyant molar masses and the thermodynamic study of reversible interactions in solution. A significant experimental drawback is the long time required to attain SE, which is usually on the order of days. We have developed a method for time-optimized SE (toSE) with defined time-varying centrifugal fields that allow SE to be attained in a significantly (up to 10-fold) shorter time than is usually required. To achieve this, numerical Lamm equation solutions for sedimentation in time-varying fields are computed based on initial estimates of macromolecular transport properties. A parameterized rotor-speed schedule is optimized with the goal of achieving a minimal time to equilibrium while limiting transient sample preconcentration at the base of the solution column. The resulting rotor-speed schedule may include multiple over- and underspeeding phases, balancing the formation of gradients from strong sedimentation fluxes with periods of high diffusional transport. The computation is carried out in a new software program called TOSE, which also facilitates convenient experimental implementation. Further, we extend AUC data analysis to sedimentation processes in such time-varying centrifugal fields. Due to the initially high centrifugal fields in toSE and the resulting strong migration, it is possible to extract sedimentation coefficient distributions from the early data. This can provide better estimates of the size of macromolecular complexes and report on sample homogeneity early on, which may be used to further refine the prediction of the rotor-speed schedule. In this manner, the toSE experiment can be adapted in real time to the system under study, maximizing both the information content and the time efficiency of SE experiments.  相似文献   

9.
The time-derivative method for deriving the sedimentation coefficient distribution, g(s*), from sedimentation velocity data that was developed by Walter Stafford has many advantages and is now widely used. By fitting Gaussian functions to the g(s*) distribution both sedimentation and diffusion coefficients (and therefore molecular masses) for individual species can be obtained. However, some of the approximations used in these procedures limit the accuracy of the results. An alternative approach is proposed in which the dc/dt data are fitted rather than g(s*). This new approach gives improved accuracy, extends the range to sedimentation coefficients below 1 S, and enhances resolution of multiple species. For both approaches the peaks from individual species are broadened when the data cover too wide a time span, and this effect is explored and quantified. An alternative algorithm for calculating ?(s*) from the dc/dt curves is presented and discussed. Rather than first averaging the dc/dt data for individual scan pairs and then calculating ?(s*) from that average, the ?(s*) distributions are calculated for every scan pair and then subsequently averaged. This alternative procedure yields smaller error bars for g(s*) and somewhat greater accuracy for fitted hydrodynamic properties when the time span becomes large.  相似文献   

10.
The glucocorticoid receptor from mouse AtT-20 pituitary tumor cells exists in three forms. The largest form is an untransformed (non-DNA-binding), oligomeric species (9.1 S, 8.3 nm, Mr 319 000). Two transformed (DNA-binding) forms can be generated. One is an oligomeric protein (5.2 S, 6-8.3 nm, Mr 132 000-182 000), while the other is the monomeric, hormone-binding subunit (3.8 S, 6 nm, Mr 96 000). The composition of the oligomeric, transformed receptor and its relationship to the monomeric protein were examined. The 3.8S monomer can be isolated from DEAE-cellulose (0.12 M step elution) in a form that continues to sediment at about 3.8 S on molybdate-containing sucrose gradients and at about 4.2 S on molybdate-free gradients. Addition of a non-hormone-binding component isolated from the same DEAE-cellulose column (0.5 M KCl step) can apparently interact with the 3.8-4.2 S monomer, increasing its sedimentation coefficient to 5.2 S (on molybdate-containing gradients) or 6.6 S (on low-salt, molybdate-free gradients). This factor is a macromolecule (nondialyzable) and is heat-stable (100 degrees C, 20 min). A dose-dependent shift to the higher sedimentation coefficient is observed when increasing quantities of the 0.5 M step material are added to the receptor monomer. This activity is abolished when the 0.5 M step material is treated with ribonuclease A. Further, when RNA is purified from the 0.5 M step by phenol/chloroform extraction, its ability to increase the S value of the monomer is retained. Ribonuclease treatment of the untransformed, 9.1S, oligomeric complex does not cause a significant decrease in sedimentation rate, while the same treatment of the 5.2S, oligomeric, transformed receptor (obtained after Sephadex G-25 transformation) causes a decrease in sedimentation rate to about 3.8 S. The addition of bovine liver mRNA and rRNA does not cause a shift in sedimentation rate of the receptor monomer to a discrete, higher sedimenting receptor form. However, the addition of total rabbit liver tRNA or three distinct tRNA species causes a shift in sedimentation to a similar, but not identical, form as that with the 0.5 M step material. We propose that the 5.2S, oligomeric transformed glucocorticoid receptor is composed of one monomeric hormone-binding, protein subunit (Mr 96 000) and a low molecular weight RNA (Mr 36 000). This interaction may be important for the role of the receptor in regulating gene expression.  相似文献   

11.
Direct fitting of sedimentation velocity data with numerical solutions of the Lamm equations has been exploited to obtain sedimentation coefficients for single solutes under conditions where solvent and solution plateaus are either not available or are transient. The calculated evolution was initialized with the first experimental scan and nonlinear regression was employed to obtain best-fit values for the sedimentation and diffusion coefficients. General properties of the Lamm equations as data analysis tools were examined. This method was applied to study a set of small peptides containing amphipathic heptad repeats with the general structure Ac-YS-(AKEAAKE)nGAR-NH2, n = 2, 3, or 4. Sedimentation velocity analysis indicated single sedimenting species with sedimentation coefficients (s(20,w) values) of 0.37, 0.45, and 0.52 S, respectively, in good agreement with sedimentation coefficients predicted by hydrodynamic theory. The described approach can be applied to synthetic boundary and conventional loading experiments, and can be extended to analyze sedimentation data for both large and small macromolecules in order to define shape, heterogeneity, and state of association.  相似文献   

12.
Visual acuity and contrast sensitivity progressively diminish with increasing viewing eccentricity. Here we evaluated how visual enumeration is affected by visual eccentricity, and whether subitizing capacity, the accurate enumeration of a small number (~3) of items, decreases with more eccentric viewing. Participants enumerated gratings whose (1) stimulus size was constant across eccentricity, and (2) whose stimulus size scaled by a cortical magnification factor across eccentricity. While we found that enumeration accuracy and precision decreased with increasing eccentricity, cortical magnification scaling of size neutralized the deleterious effects of increasing eccentricity. We found that size scaling did not affect subitizing capacities, which were nearly constant across all eccentricities. We also found that size scaling modulated the variation coefficients, a normalized metric of enumeration precision, defined as the standard deviation divided by the mean response. Our results show that the inaccuracy and imprecision associated with increasing viewing eccentricity is due to limitations in spatial resolution. Moreover, our results also support the notion that the precise number system is restricted to small numerosities (represented by the subitizing limit), while the approximate number system extends across both small and large numerosities (indexed by variation coefficients) at large eccentricities.  相似文献   

13.
[3H]Palmitate, PA, exchange efflux kinetics is recorded from human erythrocyte ghosts to buffer with bovine serum albumin, BSA, at 0 degrees C. The effects have been investigated of three medium/ghost volume ratios: 36, 80 and 500, of six BSA concentrations, [BSA]: 0.01, 0.02, 0.05, 0.2, 1 and 2% (1.5, 3.0, 7.5, 30, 150 and 300 microM) and of various v, molar ratios of palmitate to BSA, between 0.15 and 0.94. Data are analyzed in terms of a virtually closed three-compartment model. In theory, the tracer efflux is biexponential and the rate coefficients differ at least 20 fold [1]. The efflux rate at 2% BSA is monoexponential beyond our resolution time of about 1 s, but nearly biexponential at or below 0.2% BSA with a well-defined smallest-rate coefficient beta. beta depends strongly on [BSA] but is remarkably v independent. The medium/ghost volume ratio has no effect on beta when [BSA] > or = 0.2%, although beta measured at 2% BSA is almost 2-fold higher than at 0.2%. This suggests the presence of an unstirred layer, USL. According to our model, the observations are understood quantitatively on basis of our previously published dissociation rate constants of the PA-BSA complex, as well as PA equilibrium bindings to ghost membranes (Bojesen, I.N. and Bojesen, E. (1991) Biochim. Biophys. Acta 1069, 297-307). Essentially, beta is theoretically a function of two terms, one comprising the membrane transport parameters and the other the medium-dependent variables. Most important is the clearance with respect to monomer concentration adjacent to the membrane. The clearance is calculated on basis of quasi-stationary diffusion in USL. The data are compatible with a planar USL of 6 microns depth and with the same area as a ghost but not with a spherical USL.  相似文献   

14.
Sedimentation analysis and light-scattering measurements were made with the two forms of pig pancreas pro-(carboxypeptidase A), in order to determine some of their physical properties. The following values were found (the first value applies to the binary complex and the second one to the monomer). The A 1%/280.1 cm values were 19.9 +/- 0.3 and 16.3 +/- 0.3. The partial specific volumes v -0 were 0.707 +/- 0.016 cm3/g and 0.714 +/- 0.015 cm3/g. The sedimentation coefficients S 0/20,w were 4.90 +/- 0.15S and 3.75 +/- 0.15 S. The diffusion coefficients D 0/20,w were (5.8 +/- 0.1) X 10(-7) cm2/s and (6.95 +/- 0.15) X 10(-7) cm2/s. From these data the following values were calculated. Relative molecular masses Mr were 71 000 +/- 4000 and 46 000 +/- 3000. The frictional ratios f/fmin. were 1.37 +/- 0.06 and 1.31 +/- 0.07; assuming a value for the solvation of the molecules (delta = 0.5 g/g) the asymmetry values range from 3 to 5 for the binary complex and from 2 to 4 for the monomer. The Mr values found in the present work coincide with those found by means of polyacrylamide-gel electrophoresis in the presence of sodium dodecyl sulphate [Martínez, Avilés, SanSegundo & Cuchillo (1981) Biochem. J. 197, 141-147]. Therefore the low values obtained by those authors when using gel-filtration chromatography must be the result of the interaction of the zymogens with the gel matrix, as the asymmetry is too small to justify the large discrepancies found.  相似文献   

15.
A rapid, sensitive and specific liquid chromatography-tandem mass spectrometry (LC-MS-MS) method has been developed and validated for the simultaneous analysis of hydrocodone (HYC) and its metabolite hydromorphone (HYM) in human plasma. A robotic liquid handler and a 96-channel liquid handling workstation were used to aliquot samples, to add internal standard (I.S.), and to extract analytes of interest. A 96-well mixed-mode solid-phase cartridge plate was used to extract the analytes and I.S. The chromatographic separation was on a silica column (50 x 3 mm, 5-microm) with a mobile phase consisting of acetonitrile, water and trifluoroacetic acid (TFA) (92:8:0.01, v/v). The run time for each injection was 2.5 min with the retention times of approximately 2.1 and 2.2 min for HYC and HYM, respectively. The tandem mass spectrometric detection was by monitoring singly charged precursor-->product ion transition 300-->199 (m/z) for HYC, and 28-->185 (m/z) for HYM. The validated calibration curve range was 0.100-100 ng/ml, based on a plasma volume of 0.3 ml. The correlation coefficients were greater than or equal to 0.9996 for both HYC and HYM. The low limit of quantitation (LLOQ) was 0.100 ng/ml for both HYC and HYM with signal-to-noise ratio (S/N) of 50 and 10. respectively. The deuterated analytes, used as internal standards, were monitored at mass transitions 303-->199 (m/z) for HYC-d3 and 289-->185 (m/z) for HYM-d3. The inter-day (n= 17) precision of the quality control (QC) samples were < or = 3.5% RSD (relative standard deviation) for HYC and < or = 4.7% RSD for HYM, respectively. The inter-day accuracy of the QC samples were < or = 2.1% RE (relative error) for HYC and < or = 1.8% RE for HYM. The intra-day (n=6) precision and accuracy of the QC samples were < or = 2.6% RSD and < or = 3.0% RE for HYC, and < or = 4.7% RSD and < or = 2.4% RE for HYM. There was no significant deviation from the nominal values after a 5-fold dilution of high concentration QC samples by blank matrix. The QC samples were stable when kept at room temperature for 24-h or experienced three freeze-thaw cycles. The extraction recoveries were 86% for HYC and 78% for HYM. No detectable carryover was observed when a blank sample was injected immediately after a 2500 ng/ml sample that was 25-fold more concentrated than the upper limit of quantitation (ULOQ).  相似文献   

16.
J de Gunzburg  M Veron 《Biochemistry》1981,20(16):4547-4554
Three adenosine 3',5'-phosphate (cAMP) binding proteins were separated and partially purified from cytoplasmic extracts of Dictyostelium discoideum cells developed to aggregation competence. Two species, A and B, representing respectively 50% and 20% of the total activity, bind cAMP with very rapid kinetics and high specificity. Species A (Kd = 7.5 nM) is a monomeric protein of 36 000 daltons with a sedimentation coefficient of 2.3 S. Species B, which binds cAMP with positive cooperativity, also displays a high affinity for the ligand (Kd = 3.2 nM). This protein is present in the extracts as an equilibrium between monomeric, dimeric, and tetrameric forms with respective sedimentation coefficients of 2.4, 4.5, and 6.5 S; binding of cAMP to the monomer induces the appearance of the multimeric forms. A third cAMP binding protein (species C, Kd - 9.5 nM) was characterized as a larger protein (Mr 190 000, sedimentation coefficient of 9.2 S) which also binds adenosine and adenosyl derivatives. Species C represents 30% of the activity in the extracts and resemble the "adenosine analogue binding proteins" described in mammalian cells. The relevance of the properties of these proteins to the developmental process of D. discoideum amoebas is discussed.  相似文献   

17.
The physical properties of the cardiac muscarinic acetylcholine receptor (mAcChR) purified from porcine atria as recently described [Peterson, G.L., Herron, G.S., Yamaki, M., Fullerton, D.S., & Schimerlik, M.I. (1984) Proc. Natl. Acad. Sci. U.S.A. 81, 4993-4997] have been examined by D2O/H2O sucrose gradient sedimentation and Sephacryl S-300 gel filtration in Triton X-405 and by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). From the sedimentation experiments the partial specific volume and sedimentation constant for the mAcChR-Triton X-405 complex were determined to be 0.813 cm3/g and 5.30 S, respectively, which lead to an estimate of the molecular weight of the complex of 143 000. Gel filtration in Triton X-405 gave an estimate of the Stokes radius (4.29 nm) and an apparent molecular weight of 116 000. Combination of sedimentation and gel filtration gave an apparent molecular weight of 137 000 and a frictional ratio (f/f0) of 1.21 for the complex. The partial specific volume of the receptor calculated from composition was 0.717 cm3/g assuming 26.5% by weight carbohydrate. The amount of bound Triton X-405 was estimated at 1.011 g/g of mAcChR, which gave an apparent molecular weight of 70 900 (sedimentation) or 68 200 (sedimentation plus gel filtration) for the uncomplexed receptor. SDS-PAGE experiments at acrylamide concentrations ranging from 6% T [monomer plus bis(acrylamide)] to 17% T gave a linear range of apparent molecular weight from 67 600 (6% T) to 98 600 (17% T), and calibration against the retardation coefficient, Kr, determined from Ferguson plots gave an apparent molecular weight of 89 100 +/- 6700. From a newly developed, novel evaluation scheme the anomalous migration of the mAcChR in SDS-PAGE was found to be due to both an excess charge density and an abnormally large shape parameter (Kr), and the true molecular weight of the protein portion of the mAcChR ligand binding polypeptide was estimated to be between 50 000 and 60 000.  相似文献   

18.
The four molecular forms of chick embryo leg muscle acetylcholinesterase have been isolated by velocity sedimentation; their apparent sedimentation coefficients are 19.5 S, 11.5 S, 7.1 S, and 5.4 S. All four forms are glycoproteins, exhibit the same Km for acetylcholine, and are inhibited to the same extent by specific inhibitors of acetyl- and buryrylcholinesterase. Treatment of the 19.5 S form of acetylcholinesterase with trypsin generates an array of molecular forms, several of which have sedimentation coefficients identical with the naturally occurring forms. Collagenase treatment of the 19.5 S acetylcholinesterase results in a somewhat different pattern of acetylcholinesterase forms including a novel 20.6 S form. Only the 19.5 S acetylcholinesterase is sensitive to collagenase treatment. Our results indicate that the several acetylcholinesterase forms share a common catalytic subunit, and suggest that the molecular forms of acetylcholinesterase in the chick represent different ensembles of a common monomer. In culture, the muscle cells contain only the 11.5 and 7.1 S acetylcholinesterase forms; however, they also secrete substantial amounts of enzyme into the medium. These secreted acetylcholinesterases have sedimentation coefficients of 9 S and 15 S. The relative abundance of the different acetylcholinesterase molecular forms changes during muscle development, both in vivo and in vitro, suggesting that the assembly and distribution of this family of membrane glycoproteins is developmentally regulated.  相似文献   

19.
We have examined in detail analytical solutions of expressions for sedimentation equilibrium in the analytical ultracentrifuge to describe self-association under nonideal conditions. We find that those containing the radial dependence of total solute concentration that incorporate the Adams-Fujita assumption for composition-dependence of activity coefficients reveal potential shortcomings for characterizing such systems. Similar deficiencies are shown in the use of the NONLIN software incorporating the same assumption about the interrelationship between activity coefficients for monomer and polymer species. These difficulties can be overcome by iterative analyses incorporating expressions for the composition-dependence of activity coefficients predicted by excluded volume considerations. A recommendation is therefore made for the replacement of current software packages by programs that incorporate rigorous statistical-mechanical allowance for thermodynamic nonideality in sedimentation equilibrium distributions reflecting solute self-association.  相似文献   

20.
I H Brown  J Vinograd 《Biopolymers》1971,10(10):2015-2028
Catenated molecules of closed circular DNA have been isolated from the mitochondrial DNA of HeLs cells. The sedimentation coefficients of several purified species have been investigated. The catenated dimer, made up of two interlocked duplex circles, sediments at 51 S in its superhelical (closed) form. Treatment with pancreatic DNase to relax the duplex circles converts the 51 S doubly closed dimer to a 42 S singly open species, then to a 36 S doubly open catenated dimer. The triply closed trimer sediments at 63 S and is converted to a 45 S triply open form by DNase. Electron microscopy of the DNA samples before and after DNase treatment shows that under the conditions used DNase does not change the catenated nature of the DNA. The measured sedimentation coefficients, have been compared with those estimated from previously proposed correlations of sedimentation coefficient and molecular weight, and with the sedimentation coefficients for catenated DNA presented by Wang. When all the interlocked circles in a catenane are relaxed, the DNA sediments about 5–10% faster than a relaxed multiple-length circular molecule of the same molecular weight. The sedimentation coefficient, 36 S, of the fully relaxed catenated dimer is 1.4 times that of the relaxed monomer.  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