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.
We have investigated the effect of reduced water activity on the pressure-stability of double-stranded DNA polymers, poly[d(A-T)] and poly[d(I-C)]. Water activity was modulated by the addition of ethylene glycol and glycerol. The ionic strength of the medium was such that pressure had a destabilising effect on the polymers in the absence of cosolvents. The molar volume change of the heat-induced helix to coil transition (ΔVT) becomes more positive as the activity of water was reduced, suggesting that the pressure-induced denaturation of DNA polymers would not occur at very low water activity. This would imply that water plays a crucial role in the pressure denaturation of DNA, much like that in pressure denaturation of proteins where the driving force of the process is the penetration of water molecules into the protein core [Hummer et al., Proc Natl Acad Sci USA 1998, 95, 1552–1555].  相似文献   

2.
The kinetics of renaturation of heat- or formamide-denatured DNA have been studied by following the change of optical density at a constant temperature. Solvents of different ionic strength and various DNA samples have been used. At the lower ionic strengths studied, the reaction follows second-order kinetics, substantiating the hypothesis that strands of native DNA separate upon denaturation and recombine during renaturation. As the ionic strength is increased at a constant temperature, the reaction deviates from simple second-order behavior. This appears to be the result of the inhibition to rewinding caused by short helical segments in the denatured DNA which are more stable at the higher ionic strenth.  相似文献   

3.
We studied the ability of frozen-thawed mouse morulae to develop in vitro when the cryoprotectant proteins were substituted with one of the following nonorganic macromolecules: polyvinyl alcohol (PVA), polyvinyl pyrrolidone (PVP), and ficoll. We also determined how these agents interacted with 3 different cryoprotectants: glycerol (GLY), propylene glycol (PG), and ethylene glycol (EG). The influence of both of the above factors was measured on the basis of post-thaw morphological appearance, the percentage of development to the expanded blastocyst stage and the total cell count. Morulae (n=950) were collected from superovulated mice. Those classified as good or excellent were distributed among the 12 different freezing solutions, obtained by combining the 3 cryoprotectants with the 4 macromolecules (the 3 mentioned above, plus a control of 5% fetal calf serum) in phosphate buffered saline (PBS). Embryos frozen in PVA, PVP and ficoll tended to be a little difficult to recover from the straws. Development to the expanded blastocyst stage was significantly lower (P<0.05) in propylene glycol (43.6%) than in ethylene glycol (79.5%) or in glycerol (76.1%). Polyvinyl alcohol provided a higher survival rate when combined with glycerol (90.3) or ethylene glycol (95.0), but when it was combined with propylene glycol, only 56.5% of embryos survived after thawing. A positive interaction was observed between glycerol and PVA and between ethylene glycol and PVA or ficoll. The results indicate that fetal serum could be successfully substituted for any of the 3 chemically defined macromolecules. However, our findings also suggest that the use of PG as a cryoprotectant should be avoided when mouse morulae are frozen using the quick freezing method.  相似文献   

4.
Two different methods were used to isolate high-molecular-weight heavy satellite DNA from the spermatozoa of the sea-urchin Lytechinus variegratus. Purification of satellite rDNA (ribosomal DNA) by selective denaturation of the spermatozoal DNA followed by a separation of the native and denatured DNA in an aqueous mixture of poly(ethylene glycol)/dextran is affected by the molecular weight of the spermatozoal DNA. Enrichment of satellite rDNA by selective precipitation of main-band DNA by poly-L-lysine is cumbersome but more suitable.  相似文献   

5.
A M Schwartz  G D Fasman 《Biopolymers》1979,18(5):1045-1063
Chromatin was thermally denatured in the presence and absence of 1M ethylene glycol using a technique whereby both the hyperchromism and ellipticity are monitored simultaneously. Model complexes containing poly(L -Lys) or poly(L -Lys, L -Ala, Gly) and DNA were similarly melted in order to assist in interpreting the chromatin results. In both cases a general pattern emerged whereby ethylene glycol perturbed the resultant melting profile, showing increased hyperchromicity, ellipticity, and premelt slope. In addition, ethylene glycol destabilizes and reduces the high melting region of polypeptide-bound DNA and the extent of higher ordered structure in model complexes and chromatin. These results emphasize the importance of hydrophobic forces in polypeptide–polypeptide and polypeptide–DNA interactions.  相似文献   

6.
The accumulation of the cosolutes ethylene glycol, urea, glycine, sarcosine, and glycine betaine at the single-stranded DNA surface exposed upon melting the double helix has been quantified for DNA samples of different guanine-cytosine (GC) content using the local-bulk partitioning model [Record, M. T., Jr., Zhang, W., and Anderson, C. F. (1998) Adv. Protein Chem. 51, 281-353]. Urea and ethylene glycol are both locally accumulated at single-stranded DNA relative to bulk solution. Urea exhibits a stronger affinity for adenine (A) and thymine (T) bases, leading to a greater net dehydration of these bases upon DNA melting; ethylene glycol local accumulation is practically independent of base composition. However, glycine, sarcosine, and glycine betaine are not necessarily locally accumulated at single strands after melting relative to bulk solution, although they are locally accumulated relative to double-stranded DNA. The local accumulation of glycine, sarcosine, and glycine betaine at single strands relative to double-stranded DNA decreases with bulk cosolute molality and increases with GC content for all N-methylated glycines, demonstrating a stronger affinity for G and C bases. Glycine also shows a minimum in melting temperature T(m) at 1-2 m for DNA samples of 50% GC content or less. Increasing ionic strength attenuates the local accumulation of urea, glycine, sarcosine, and glycine betaine and removes the minimum in T(m) with glycine. This attenuation in local accumulation results in counterion release during the melting transition that is dependent on water activity and, hence, cosolute molality.  相似文献   

7.
Native calf thymus DNA was gamma-irradiated at 500 mug/ml in 0.01 M NaCl in the presence or absence of oxygen. By irradiation, an increasing amount of DNA becomes reactive with a water-soluble carbodiimide-derivative (CMEC). In the DNA sections reactive with CMEC the nucleotide strands are separated, a phenomenon previously described as radiation-induced denaturation. The dose-effect curve for the formation of denatured DNA shows an upward-bent form; a distinct oxygen effect of about 2 is observed. By a comparative study with DNA samples, degraded partially with DNAse I, it was shown that a minor part of the radiation-induced denaturation results from the formation of the radiation-induced single strand breaks, whereas the major part is a local denaturation independent of the strand breaks. In these locally denatured regions 20 to 50 nucleotide pairs are separated.  相似文献   

8.
We have studied the linear dichroism (LD) of rat liver chromatin oriented by flow. Soluble chromatin, prepared by brief nuclease digestion, is found to exhibit a positive LD at low ionic strength (1 mM NaCl), with a constant LD/A over the absorption band centered at 260 nm (A, isotropic absorbance). Several previous dichroism studies on soluble chromatin have been performed on sonicated materials and have given negative LD, probably due to the presence of uncoiled DNA. The positive dichroism can be interpreted in terms of a supercoil of DNA in chromatin with a pitch angle larger than 55°, and is, for example, consistent with a model where the cylindrical nucleosome core particles are stacked face to face in the chromatin filament. In contrast to the nuclease-digested chromatin, sonicated chromatin was confirmed to exhibit negative LD. This difference can be attributed to a partial uncoiling of the linker regions between the nucleosomes due to the shearing. The structural transition of chromatin to a compact form can be observed as a reduction of the positive LD of the nuclease-digested chromatin to almost zero in 0.1 M NaCl or in 0.1 mM MgCl2. This transition is due to a decreased electrostatic repulsion between negative phosphate groups on the DNA chain. In the case of Na+, this can be explained as a screening effect due to the bulk concentration of Na+. With Mg2+ a considerably stronger effect may indicate a more localized binding to the phosphates. At ionic strengths higher than 0.5M NaCl, the dissociation of the histones from DNA leads to uncoiling of chromatin. The change in LD during this process shows that histone H1 contributes only to a small degree to the coiling of the DNA chain, whereas histones H3 and H4 play the major role in the coiling.  相似文献   

9.
Steven Spiker  Roger Chalkley 《Planta》1972,102(4):362-367
Summary Pea (Pisum sativum L.) nucleohistone and chromatin which had been treated both in vivo and in vitro with indoleacetic acid and gibberellic acid were thermally denatured in low and high ionic strength media. Contrary to previous reports, the hormones had no effect on thermal denaturation profiles. The previously observed biphasic nature of the profiles obtained at high ionic strength is explained.  相似文献   

10.
Summary DNA molecules of B. subtilis phage SPP1 exhibit terminal redundancy and are partially circularly permuted. This was established by the hybridization of selected EcoRI restriction fragments to single strands of SPP1 DNA and by an analysis of the distribution of denaturation loops in partially denatured SPP1 DNA molecules. Deletions in SPP1 DNA are not compensated by an increase in terminally repetitious DNA. This finding, which is unique to SPP1, is discussed in terms of a modification of the Streisinger/Botstein model of phage maturation.  相似文献   

11.
SYNOPSIS. Cryoprotectants were tested in both complex and semidefined media for the trypanosomatid Crithidia fasciculata. Near log-phase or end-of-log-phase cultures were frozen for 24–48 hr at ∼ -20 C, then warmed in air to room temperature. Immediate motility was correlated with viability. The best protectant of the 83 tested was glycerol at ∼ 10% (w/v). Survival without cryoprotectant was rare. Outstanding cryoprotectants (perhaps also useful solvents for drugs poorly soluble in water) were: ethylene glycol; 2,2'-dioxyethanol (diethylene glycol); 1,2,4-butanetriol; 1,4-cyclohexanediol; dimethylsulfoxide; propylene glycol; and N -acetylethanolamine. Several sugars were active, e.g., D-arabinose, sucrose, and sorbitol. Trypanosomes tolerated cryoprotectants much less; tolerance was better in growth media than in suspension media. Trypanosoma gambiense was grown in blood-enriched media + 2-2.5% glycerol, suspended in 20% (w/v) glycerol. then frozen; this permitted 3-week survival. T. conorhini survived 4 weeks after growth in media containing glycerol 2.5%+ ethylene glycol 4%+ rutin 1.0 mg per 100 ml.  相似文献   

12.
The intrinsic viscosity and sedimentation coefficient, of native and single-stranded T7 DNA have been determined at 25°C as a function of ionic strength in neutral and alkaline NaCl. The relationship between [η] and S,w is well represented by the Mandelkern-Flory equation over the entire range of conditions between 0.0013 and 1M Na+. An apparent discrepancy between the two methods at moderate to high ionic strengths is probably due to a change in V with ionic strength. It appears that [η] is a more sensitive and reliable measure of molecular expansion for native DNA, S,w but is a better index of conformational change in single strands, since [η] becomes too small to measure conveniently at high ionic strengths. At moderate to high ionic strengths, denaturation leads to a decrease in [η], although unfolded single strands retain considerable viscosity. At sufficiently low ionic strength, the intrinsic viscosity of the single strands becomes higher than that of native DNA, and the effective volume of a single strand approaches that of the native molecule.  相似文献   

13.
Vitrification enhancement by synthetic ice blocking agents   总被引:7,自引:0,他引:7  
Small concentrations of the synthetic polymer polyvinyl alcohol (PVA) were found to inhibit formation of ice in water/cryoprotectant solutions. Ice inhibition improved with decreasing molecular weight. A PVA copolymer of molecular weight 2 kDa consisting of 20% vinyl acetate was found to be particularly effective. PVA copolymer concentrations of 0.001, 0.01, 0.1, and 1% w/w decreased the concentration of glycerol required to vitrify in a 10-ml volume by 1, 3, 4, and 5% w/w, respectively. Dimethyl sulfoxide concentrations required for vitrification were also reduced by 1, 2, 2, and 3% w/w, respectively. Crystallization of ice on borosilicate glass in contact with cryoprotectant solutions was inhibited by only 1 ppm of PVA copolymer. Devitrification of ethylene glycol solutions was also strongly inhibited by PVA copolymer. Visual observation and differential scanning calorimeter data suggest that PVA blocks ice primarily by inhibition of heterogeneous nucleation. PVA thus appears to preferentially bind and inactivate heterogeneous nucleators and/or nascent ice crystals in a manner similar to that of natural antifreeze proteins found in cold-hardy fish and insects. Synthetic PVA-derived ice blocking agents can be produced much less expensively than antifreeze proteins, offering new opportunities for improving cryopreservation by vitrification.  相似文献   

14.
To study the alkali denaturation of supercoiled DNA, plasmid pBR322 was treated with gradient concentrations of NaOH solution. The results of gel electrophoresis showed that the alkali denaturation of the supercoiled DNA occurred in a narrow range of pH value (12.88-12.90). The alkali-denatured supercoiled DNA ran, as a sharp band, faster than the supercoiled DNA. The supercoiled plasmid DNA of pBR322, pACYC184 and pJGX15A were denatured by NaOH, and then visualized by atomic force microscopy. Compared with the supercoiled DNA, the atomic force microscopy images of the alkali-denatured supercoiled DNA showed rough surface with many kinks, bulges on double strands with inhomogeneous diameters. The apparent contour lengths of the denatured DNA were shortened by 16%, 16% and 50% for pBR322, pACYC184 and pJGX15A, respectively. All evidence suggested that the alkali-denatured supercoiled DNA had a stable conformation with unregistered, topologically constrained double strands and intrastrand secondary structure.  相似文献   

15.
Dissociation of tetrameric l-asparaginase from Escherichia coli B was examined in the presence of urea containing one of the following polyhydric alcohols: ethylene glycol, 1,2-propanediol, 1,3-propanediol, glycerol, erythritol, arabitol, adonitol, mannitol, sorbitol, inositol, glucose, sucrose, and polyethylene glycol. Low concentrations of these compounds accelerate the rate of subunit dissociation, and, with the exception of the propanediols and polyethylene glycol, higher concentrations decrease the rate at which the oligomeric enzyme dissociates. The specific concentration at which this transition occurs is related to the length of the carbon chain of the polyhydric alcohols and to the steric configuration of the hydroxyl groups about the asymmetric carbon atoms. In addition, the rate at which the oligomeric enzyme dissociates decreases as the number of hydroxymethyl groups per molecule polyol increases and reaches a maximum with the six carbon members.Low concentrations (1% by volume) of methanol, ethanol, ethylene glycol, propylene glycol, or glycerol contained in the renaturation buffer slightly accelerate the rate of reassembly of denatured subunits. The rate at which reassociation to the tetramer occurs also increases as the number of hydroxymethyl groups per molecule of polyhydric alcohol increases.  相似文献   

16.
Using immunochemical technique thermal denaturation of soybean 11S globulin, dissolved in different ionic strength solutions (µ=0~4.0) and heated at 100°C for 5 min, has been quantitatively studied. The curves of the percentage of antigenicity remaining were obtained as a function of salt concentration. The 11S globulin became strongly resistant to thermal denaturation with increasing both KCl and potassium phosphate. The stabilizing effect (in terms of percent antigenicity) was separated into three regions. At ionic strength below 0.7, potassium phosphate had no stabilizing effect while KCl had aslightly effect. The rise in stabilizing effect up to about 50%, near 1.0~1.5 µ, represented a second transition to a different denatured state which retains undissociated molecule. At rises up to 75~95%, near 2.5~3.5µ, a different conformational state resulted in which thermally denatured 11S globulin maintained almost intact native conformation after heating. The selection of an adequate ionic strength of protein solution has enabled preparation of thermally denatured 11S globulins which have desired-residual amounts of structured regions.  相似文献   

17.
Dissociation and reconstitution of a DNA polymerase alpha-primase complex   总被引:3,自引:0,他引:3  
The conditions for dissociation of the DNA polymerase alpha-primase complex (DNA polymerase alpha 1) have been examined. It was revealed that 50% ethylene glycol effectively dissociated the complex. The dissociated DNA polymerase and primase were purified to eliminate cross-contaminating activities by column chromatography using buffers containing 50% ethylene glycol. The sedimentation coefficients of the purified DNA polymerase and primase were 7.1S and 5.7S, respectively. These two enzymes were mixed in the presence of 20% ethylene glycol and the mixture was sedimented through a glycerol gradient containing no ethylene glycol. The DNA polymerase and primase activities co-sedimented at 9.1S which corresponds to the S value of intact alpha 1, indicating the reconstitution of the DNA polymerase alpha-primase complex.  相似文献   

18.
Paul S  Punam S  Chaudhuri TK 《The FEBS journal》2007,274(22):6000-6010
In vitro refolding of maltodextrin glucosidase, a 69 kDa monomeric Escherichia coli protein, was studied in the presence of glycerol, dimethylsulfoxide, trimethylamine-N-oxide, ethylene glycol, trehalose, proline and chaperonins GroEL and GroES. Different osmolytes, namely proline, glycerol, trimethylamine-N-oxide and dimethylsulfoxide, also known as chemical chaperones, assist in protein folding through effective inhibition of the aggregation process. In the present study, it was observed that a few chemical chaperones effectively reduced the aggregation process of maltodextrin glucosidase and hence the in vitro refolding was substantially enhanced, with ethylene glycol being the exception. Although, the highest recovery of active maltodextrin glucosidase was achieved through the ATP-mediated GroEL/GroES-assisted refolding of denatured protein, the yield of correctly folded protein from glycerol- or proline-assisted spontaneous refolding process was closer to the chaperonin-assisted refolding. It was also observed that the combined application of chemical chaperones and molecular chaperone was more productive than their individual contribution towards the in vitro refolding of maltodextrin glucosidase. The chemical chaperones, except ethylene glycol, were found to provide different degrees of protection to maltodextrin glucosidase from thermal denaturation, whereas proline caused the highest protection. The observations from the present studies conclusively demonstrate that chemical or molecular chaperones, or the combination of both chaperones, could be used in the efficient refolding of recombinant E. coli maltodextrin glucosidase, which enhances the possibility of identifying or designing suitable small molecules that can act as chemical chaperones in the efficient refolding of various aggregate-prone proteins of commercial and medical importance.  相似文献   

19.
The Cu(II) ion interaction with calf-thymus DNA was studied by means of differential pulse polarography and sweep voltammetry as well as chromatography and viscosimetry. Most of the complexes formed at high ionic strength (0.2 M) and lower Cu(II) concentrations are of a nondenaturing nature. Their formation has but a minor effect on unwinding process of the DNA double helix. The excess of Cu(II) (P = 5) leads, however, to distinct denaturation of the DNA structure. Metal ions have little effect on the denaturation induced by the polarographic reduction of DNA on the mercury electrode. This conclusion is consistent with the character of the polarographic process and with the fact that Cu(II) ions are not very effective in the interaction with AT pairs. Cupric ions have no renaturing ability towards thermally denatured DNA at 0.2 M ionic strength but distinct renaturation was observed at low ionic strength (0.05 M).  相似文献   

20.
M J Tunis  J E Hearst 《Biopolymers》1968,6(9):1325-1344
The hydration of DNA is an important factor in the stability of its secondary structure. Methods for measuring the hydration of DNA in solution and the results of various techniques are compared and discussed critically. The buoyant density of native and denatured T-7 bacteriophage DNA in potassium trifluoroacetate (KTFA) solution has been measured as a function of temperature between 5 and 50°C. The buoyant density of native DNA increased linearly with temperature, with a dependence of (2.3 ± 0.5) × 10?4 g/cc-°C. DNA which has been heat denatured and quenched at 0°C in the salt solution shows a similar dependence of buoyant density on temperature at temperatures far below the Tm, and above the Tm. However, there is an inflection region in the buoyant density versus T curve over a wide range of temperatures below the Tm. Optical density versus temperature studies showed that this is due to the. inhibition by KTFA of recovery of secondary structure on quenching. If the partial specific volume is assumed to be the same for native and denatured DNA, the loss of water of hydration on denaturation is calculated to be about 20% in KTFA at a water activity of 0.7 at 25°C. By treating the denaturation of DNA as a phase transition, an equation has immmi derived relating the destabilizing effect of trifluoroacetate to the loss of hydration on denaturation. The hydration of native DNA is abnormally high in the presence of this anion, and the loss of hydration on denaturation is greater than in CsCl. In addition, trifluoroacetate appears to decrease the ΔHof denaturation.  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