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.
Thermal denaturation of Na- and Li-DNA from chicken erythrocytes was studied by means of scanning microcalorimetry in salt-free solutions at DNA concentrations (Cp) from 4.5 · 10?2 to 1 · 10?3 moles of nucleotides/liter (M). Linear dependencies of DNA melting temperature (Tm) vs lgCp were obtained: ((1)) ((2)) for Na- and Li-DNA, respectively. Microcalorimetry data were compared with the results of spectrophotometric studies at 260 nm of DNA thermal denaturation in Me-DNA + MeCl solutions at Cp ? (6–8) · 10?5 M and Cs = 0–40 mM (Me is Na or Li, Cs is salt concentration). It was found that Eqs. (1) and (2) are valid in DNA salt-free solutions over the Cp range 6 · 10?5?4.5 · 10?2M. Protonation of DNA bases due to the absorption of CO2 from air in Na-DNA + NaCl solutions affects DNA melting parameters at Cs < 4 mM. Linear dependence of Tm on lga+ is found in Na-DNA + NaCl at Cs > 0.4 mMin the absence of contact of solutions with CO2 from air (a+ is cation activity). A dependence of [dTm/dlga+] on Li+ activity was observed in Li-DNA + LiCl solutions at Cs < 10 mM: [dTm/dlga+] increases from 17°–18° at Cs > 10 mM to 28°–30° at Cs ? 0.2–0.4 mM. Spectrophotometric measurements at 282 nm show that this effect was caused by protonation of bases in fragments of denatured DNA in neutral solutions. The Poisson–Boltzmann (PB) equation was solved for salt-free DNA at the melting point. The linear dependence of Tm vs lgCp was interpreted in terms of Manning's condensation theory. PB and Manning's theories fit the experimental data if charge density parameter (ξ) of denatured DNA is in the range 1.8–2.1 (assuming for native DNA ξ = 4.2). Specificity of Li ions in interactions with DNA is discussed. © 1994 John Wiley & Sons, Inc.  相似文献   

2.
R L Ornstein  J R Fresco 《Biopolymers》1983,22(8):1979-2000
Tm values for 20 DNA duplexes with different repeating base sequences provided the data base for developing a rational and relatively simple methodology for computing apparent enthalpies for the helix → coil transitions of DNA helices, ΔH calc. The computational variables and their range of acceptable values were selected on the basis of physically plausible arguments. Over 350,000 different combinations of the variables were tested for degree of fit. It was therby possible to find a combination giving a high degree of linear fit between Tm and ΔH calc (correlation coefficient, 0.99), with Tm values deviating (on average) from the regression line by only ±2.17°C. Most of this uncertainty is attributed to experimental limitations, although computational approximations also contribute. With ΔH calc for the melting of each of the unique complementary dinucleotide fragments computed by the method developed, it is possible to estimate Tm and (relative) ΔH calc reliable for the melting of any particular DNA [with base pairs G(I)·C and A·T] given only its base sequence. The ΔHcalc values for the complementary dinucleotide fragments, together with statistical considerations, make it apparent that Tm of DNAs with repeating base sequence show only an approximate linear dependence on G·C content because A·T and G·G pairs do not contribute to helix stability independently of the base-pair sequence in which they occur. In fact, the nearestneighbor stacking interactions are so significant that certain complementary dinucleotide fragment sequences with 0,50, and 100% G·C content have the same stability.  相似文献   

3.
H J Li  B Brand  A Rotter  C Chang  M Weiskopf 《Biopolymers》1974,13(8):1681-1697
Thermal denaturation of direct-mixed and reconstituted polylysine–DNA complexes in 2.5 × 10?4 M EDTA, pH 8.0 and various concentrations of NaCl has been studied. For both complexes, increasing ionic strength of the solution raises Tm, the melting temperature of free base pairs. The linear dependence of Tm on log Na+ indicates that the concept of electrostatic shielding on phosphate lattice of an infinitely long pure DNA by Na+ can be applied to short free DNA segments in a nucleoprotein. For a direct-mixed polylysine–DNA complex, the melting temperature of bound base pairs Tm′ remains constant at various ionic strengths. On the other hand, the Tm′ in a reconstituted polylysine–DNA complex is shifted to lower temperature at higher ionic strength. This phenomenon occurs for reconstituted complex with long polylysine of one thousand residues or short polylysine of one hundred residues. It is shown that such a decrease of Tm′ is not due to a reduction of coupling melting between free and bound regions in a complex when the ionic strength is raised. It is also not due to intermolecular or intramolecular change from a reconstituted to a direct-mixed complex. It is suggested that this phenomenon is due to structural change on polylysine-bound regions by ionic strength. It is suggested further that Na+ may replace water molecules and bind polylysine-bound regions in a reconstituted complex. Such a dehydration effect destabilizes these regions and lowers Tm′. This explanation is supported by circular dichroism (CD) results.  相似文献   

4.
Interaction between polylysine and DNA's of varied G + C contents was studied using thermal denaturation and circular dichroism (CD). For each complex there is one melting band at a lower temperature tm, corresponding to the helix–coil transition of free base pairs, and another band at a higher temperature tm, corresponding to the transition of polylysine-bound base pairs. For free base pairs, with natural DNA's and poly(dA-dT) a linear relation is observed between the tm and the G + C content of the particular DNA used. This is not true with poly(dG)·poly(dC), which has a tm about 20°C lower than the extrapolated value for DNA of 100% G + C. For polylysine-bound base pairs, a linear relation is also observed between the tm and the G + C content of natural DNA's but neither poly(dA-dT) nor poly(dG)·poly(dC) complexes follow this relationship. The dependence of melting temperature on composition, expressed as dtm/dXG·C, where XG·C is the fraction of G·C pairs, is 60°C for free base pairs and only 21°C for polylysine-bound base pairs. This reduction in compositional dependence of Tm is similar to that observed for pure DNA in high ionic strength. Although the tm of polylysine-poly(dA-dT) is 9°C lower than the extrapolated value for 0% G + C in EDTA buffer, it is independent of ionic strength in the medium and is equal to the tm0 extrapolated from the linear plot of tm against log Na+. There is also a noticeable similarity in the CD spectra of polylysine· and polyarginine·DNA complexes, except for complexes with poly(dA-dT). The calculated CD spectrum of polylysine-bound poly(dA-dT) is substantially different from that of polyarginine-bound poly(dA-dT).  相似文献   

5.
R D Blake  P V Haydock 《Biopolymers》1979,18(12):3089-3109
A series of high-resolution melting curves were obtained by the continuous direct-derivative method [Blake, R. D. & Lefoley, S. G. (1978) Biochim. Biophys. Acta 518 , 233–246] on lambda DNA (cI857S7 strain) under varying conditions of [Na+]. Examination of the denaturation patterns at close intervals of [Na+] indicates that frequent changes in mechanism occur below 0.04M Na+, while almost none occurs above 0.1M Na+. Changes at low [Na+] generally occur in an abrupt fashion, in most cases within a 3 mM change in [Na+], and in at least one case within 0.6 mM, indicating the balance between alternative mechanisms is frequently quite delicate. These changes involve segments of between 900 and 1500 or more base pairs in length and are therefore not insignificant. Changes at low [Na+] reflect a perturbation of the energetic balance between competing mechanisms by weakly screened long-range electrostatic forces. Some perturbation probably also arises from variations in the linear charge density of the double helix induced by the proximity of premelted loop segments; however, this contribution cannot be evaluated without a detailed denaturation map. At high [Na+] the mechanism of melting is more conserved, permitting the dependence of subtrasitional melting temperature tm(i) on [Na+] to be examined for almost all 34 ± 2 subtransitions. The G + C composition of segments responsible for each subtransition was determined by a quantitative spectral method. Analysis according to the Manning-Record expression [Manning, G. (1972) Biopolymers 11 , 937–949; Record, M. T., Jr., Anderson, C. F. & Lohman, T. M. (1978) Q. Rev. Biophysics 11 , 103–178] relating ΔHm and dtm(i)/d log[Na+] to the fraction of Na+ released during melting, appears to indicate almost 40% more Na+ is bound to the single-stranded G and/or C residues than to A and T residues. This is consistent with a much shorter mean axial spacing and higher charge density in the former, particularly single-stranded G residues, which have an extraordinary tendency to stack.  相似文献   

6.
The binding of tiamulin with calf thymus DNA was systematically investigated using multispectroscopy and molecular modelling techniques. For DNA, once tiamulin was added, viscosity (η) and melting temperature (Tm) both exhibited an uptrend. The fluorescence performance of the tiamulin–DNA complex did not change with the ionic strength changes. The binding constant (Ka) of tiamulin for single-stranded DNA (ssDNA, 1.48 × 104 M−1) was obviously higher than that for double-stranded DNA (dsDNA, 9.51 × 103 M−1) at 291 K. The helix structure became looser and the base stack force became stronger for DNA due to the presence of tiamulin as seen from circular dichroic (CD) spectra. The intercalation binding mode of tiamulin with DNA was disclosed. Molecular modelling also revealed tiamulin inserting into the base pairs with the lowest binding free energy of −18.73 kJ mol−1 using van der Waals forces as well as hydrogen bonds.  相似文献   

7.
M T Record 《Biopolymers》1967,5(10):993-1008
The theory developed in the previous paper to discuss changes in electrostatic free energies in polynucleotide order–disorder transitions is extended to cases where one or more of the participating species is titrated to some degree α. It is shown that, for any class of transition, the melting temperature Tm at constant pH is a linear function of the logarithm of the monovalent counterion concentration M, that at high salt the logarithm of the depression of the melting temperature by pH titration is proportional to the pH change, and that the stability of the ordered form as measured by its melting temperature at neutral pH, is a monotonic function of the quantity pHm – pK, where pHm and pK are the pH of melting and the monomer base pK, both measured under similar conditions of temperature and ionic strength. For the transition from double helix to coil, the dependences of Tm and dTm/d log M on pH are determined experimentally and compared with the qualitative predictions of the theory. It is found that dTm/dlog M, a measure of – ΔF?el (the negative of the electrostatic free energy change in the transition), decreases with increasing pH. In acid solution, where the coil is more extensively prolonated than the helix, the change in electrostatic free energy in the transition is larger than at neutral pH. Conversely, in alkali the electrostatic five energy change is smaller than at neutral pH. Hence (dTm/d log M)acid > (dTm/d log M neutral) > (dTm/d log M)alkali. At Suffeciently high pH, dTm/d log M is observed to become negative, indicating that the electrostatic free energy change is positive in the transition of this region. Date from the literature on the ionic strength dependence of the melting temperature for the acid helices of poly rA, poly rC, and poly dC are also considered from the standpoint of the theory.  相似文献   

8.
Sharon S. Yu  Hsueh Jei Li 《Biopolymers》1973,12(12):2777-2788
Protamine–DNA complexes prepared by the method of direct and slow mixing in 2.5 × 10?4M EDTA, pH 8.0, have been studied by thermal denaturation and circular dichroism. The complexes show biphasic melting with Tm at about 50 °C corresponding to the melting of free DNA regions and Tm′ at about 92 °C corresponding to the melting of protamine-bound regions. In protamine-bound regions there are 1.38 amino acid residues per nucleotide, indicating a nearly completely charge neutralization. Tm is increased but Tm′ is not when the ionic strength of the buffer is raised. This also supports a full charge neutralization in protamine-bound regions. The circular dichroism of the complexes can be decomposed into two components, Δε0 of free DNA regions in B-form conformation and Δεb of protamine-bound regions in a characteristic conformation neither that of B- nor C-form but somewhere between them.  相似文献   

9.
Ultrathin (black) lipid membranes were made from sheep red cell lipids dissolved in n-decane. The presence of aliphatic alcohols in the aqueous solutions bathing these membranes produced reversible changes in the ionic permeability, but not the osomotic permeability. Heptanol (8 mM), for example, caused the membrane resistance (Rm) to decrease from >108 to about 105 ohm-cm2 and caused a marked increase in the permeability to cations, especially potassium. In terms of ionic transference numbers, deduced from measurements of the membrane potential at zero current, T cat/T Cl increased from about 6 to 21 and T K/T Na increased from about 3 to 21. The addition of long-chain (C8ndash;C10) alcohols to the lipid solutions from which membranes were made produced similar effects on the ionic permeability. A plot of log Rm vs. log alcohol concentration was linear over the range of maximum change in Rm, and the slope was -3 to -5 for C2 through C7 alcohols, suggesting that a complex of several alcohol molecules is responsible for the increase in ionic permeability. Membrane permselectivity changed from cationic to anionic when thorium or ferric iron (10-4 M) was present in the aqueous phase or when a secondary amine (Amberlite LA-2) was added to the lipid solutions from which membranes were made. When membranes containing the secondary amine were exposed to heptanol, Rm became very low (103–104 ohm-cm2) and the membranes became perfectly anion-selective, developing chloride diffusion potentials up to 150 mv.  相似文献   

10.
Two newly isolated obligate methanol-utilizing bacteria (strains IvaT and LapT) with the ribulose monophosphate pathway of C1 assimilation are described. The isolates are strictly aerobic, Gram negative, asporogenous, motile rods multiplying by binary fission, mesophilic and neutrophilic, synthesize indole-3-acetate. The prevailing cellular fatty acids are straight-chain saturated C16:0 and unsaturated C16:1 acids. The major ubiquinone is Q-8. The predominant phospholipids are phosphatidylethanolamine, phosphatidylglycerol and cardiolipin. Ammonia is assimilated by glutamate dehydrogenase. The DNA G+C contents of strains IvaT and LapT are 54.0 and 50.5 mol% (Tm), respectively.Based on 16S rRNA gene sequence analysis and DNA–DNA relatedness (38–45%) with type strains of the genus Methylobacillus, the novel isolates are classified as the new species of this genus and named Methylobacillus arboreus IvaT (VKM B-2590T, CCUG 59684T, DSM 23628T) and Methylobacillus gramineus LapT (VKM B-2591T, CCUG 59687T, DSM 23629T).The GenBank accession numbers for the 16S rRNA gene and mxaF gene sequences of the strains IvaT and LapT are GU937479, GU937478 and HM030736, HM030735, respectively.  相似文献   

11.
Abstract

The protease from Aspergillus tamarii Kita UCP1279 extraction by aqueous two-phase PEG-Citrate (ATPS) systems, using a factorial design 24, was investigated. Then, the variables studied were polyethylene glycol (PEG) molar mass (MPEG), concentrations of PEG (CPEG) and citrate (CCIT), and pH. The responses analyzed were the partition coefficient (K), activity yield (Y) and purification factor (PF). The thermodynamic parameters of the ATPS partition were estimated as a function of temperature. ATPS was able to pre-purify the protease (PF = 1.6) and obtained 84% activity yield. The thermodynamic parameters ΔG°m (?10.89?kJ mol?1), ΔHm (?5.0?kJ?mol?1) and partition ΔSm (19.74?J mol?1 K?1) showed that the preferential migration of almost all protein contaminants of the crude extract to the salt-rich phase, while the preferred protease was the PEG rich phase. The extracted enzyme presents optimum temperature and pH at range of 40–50?°C and 9.0–11.0, respectively. Moreover, the enzyme was identified as serine protease based on inhibition profile. ATPS showed the satisfactory performance as the first step for Aspergillus tamarii Kita UCP1279 protease pre-purification.  相似文献   

12.
Spectroscopic and calorimetric melting studies of 28 DNA hairpins were performed. These hairpins form by intramolecular folding of 16 base self‐complementary DNA oligomer sequences. Sequence design dictated that the hairpin structures have a six base pair duplex linked by a four base loop and that the first five base pairs in the stem are the same in every molecule. Only loop sequence and identity of the duplex base pair closing the loop vary for the set of hairpins. For these DNA samples, melting studies were carried out to investigate effects of the variables on hairpin stability. Stability of the 28 oligomers was ascertained from their temperature‐induced melting transitions in buffered 115 mM Na+ solvent, monitored by ultraviolet absorbance and differential scanning calorimetry (DSC). Experiments revealed the melting temperatures of these molecules range from 32.4 to 60.5°C and are concentration independent over strand concentrations of 0.5 to 260 μM; thus, as expected for hairpins, the melting transitions are apparently unimolecular. Model independent thermodynamic transition parameters, ΔHcal, ΔScal, and ΔGcal, were determined from DSC measurements. Model dependent transition parameters, ΔHvH, ΔSvH, and ΔGvH were estimated from a van't Hoff (two‐state) analysis of optical melting transitions. Results of these studies reveal a significant sequence dependence to DNA hairpin stability. Thermodynamic parameters evaluated by either procedure reveal the transition enthalpy, ΔHcalHvH) can differ by as much as 20 kcal/mol depending on sequence. Similarly, values of the transition entropy ΔScalSvH) can differ by as much as 60 cal/Kmol (eu) for different molecules. Differences in free energies ΔGcalGvH) are as large as 4 kcal/mol for hairpins with different sequences. Comparisons between the model independent calorimetric values and the thermodynamic parameters evaluated assuming a two‐state model reveal that 10 of the 28 hairpins display non‐two‐state melting behavior. The database of sequence‐dependent melting free energies obtained for the hairpins was employed to extract a set of n‐n (nearest‐neighbor) sequence dependent loop parameters that were able to reproduce the input data within error (with only two exceptions). Surprisingly, this suggests that the thermodynamic stability of the DNA hairpins can in large part be reasonably represented in terms of sums of appropriate nearest‐neighbor loop sequence parameters. © 1999 John Wiley & Sons, Inc. Biopoly 50: 425–442, 1999  相似文献   

13.
Thermal transitions in mixtures of polydeoxyribodinucleotides   总被引:8,自引:0,他引:8  
Oligo d(C-A) and oligo d(T-G) of known average lengths, prepared by a combination of chemical and enzymatic procedures, have been mixed in 0.02 M and 0.07 M Na+, and absorbance has been studied as a function of increasing temperature. The transitions have been analyzed for the temperature of maximum slope Tm, the breadth of the transition, the value of the slope at Tm, and the maximum hyperchromicity. Linear expressions have been found relating the inverse of the length in nucleotide units (n?1) of the shorter oligomer, irrespective of its identity, to Tm and also to the transition breadth. From a difference in slope between the Tm versus n?1 expressions for the two molarities, the entropy and enthalpy of melting have been calculated as a function of n?1.  相似文献   

14.
Studies on poly(L-lysine50, L-tyrosine50)-DNA interaction   总被引:3,自引:0,他引:3  
R M Santella  H J Li 《Biopolymers》1974,13(9):1909-1926
Interaction between poly(Lys50, Tyr50) and DNA has been studied by absorption, circular dichroism (CD), and fluorescence spectroscopy and thermal denaturation in 0.001M Tris, pH 6.8. The binding of this copolypeptide to DNA results in an absorbance enhancement and fluorescence quenching on tyrosine. There is also an increase in the tyrosine CD at 230 nm. The CD of DNA above 250 nm is slightly shifted to the longer wavelength which is qualitatively similar to, but quantitatively much smaller than, that induced by polylysine binding. At physiological pH the poly(Lys50, Tyr50)–DNA complex is soluble until there is one lysine and one tyrosine per nucleotide in the complex. The same ratio of amino acid residues to nucleotide has also been observed in copolypeptide-bound regions of the complex. The addition of more poly(Lys50, Tyr50) to DNA yields a constant melting temperature, Tm′, for bound base pairs at 90°C which is close to that of polylysine-bound DNA under the same condition. The melting temperature, Tm, of free base pairs at about 60°C on the other hand, is increased by 10°C as more copolypeptide is bound to DNA. As the temperature is raised, both absorption and CD spectra of the complexes with high coverage are changed, suggesting structural alteration, perhaps deprotonation, on bound tyrosine. The results in this report also suggest that intercalation of tyrosine in DNA is unlikely to be the mode of binding.  相似文献   

15.
Thermoregulatory sweating [total body (m sw,b), chest (m sw,c) and thigh (m sw,t) sweating], body temperatures [oesophageal (T oes) and mean skin temperature (T sk)] and heart rate were investigated in five sleep-deprived subjects (kept awake for 27 h) while exercising on a cycle (45 min at approximately 50% maximal oxygen consumption) in moderate heat (T air andT wall at 35° C. Them sw,c andm sw,t were measured under local thermal clamp (T sk,1), set at 35.5° C. After sleep deprivation, neither the levels of body temperatures (T oes,T sk) nor the levels ofm sw, b,m sw, c orm sw, t differed from control at rest or during exercise steady state. During the transient phase of exercise (whenT sk andT sk,1 were unvarying), them sw, c andm sw, t changes were positively correlated with those ofT oes. The slopes of them sw, c versusT oes, orm sw, t versusT oes relationships remained unchanged between control and sleep-loss experiments. Thus the slopes of the local sweating versusT oes, relationships (m sw, c andm sw, t sweating data pooled which reached 1.05 (SEM 0.14) mg·cm–2·min–1°C–1 and 1.14 (SEM 0.18) mg·cm–2·min–1·°C–1 before and after sleep deprivation) respectively did not differ. However, in our experiment, sleep deprivation significantly increased theT oes threshold for the onset of bothm sw, c andm sw, t (+0.3° C,P<0.001). From our investigations it would seem that the delayed core temperature for sweating onset in sleep-deprived humans, while exercising moderately in the heat, is likely to have been due to alterations occurring at the central level.  相似文献   

16.
Thermodynamics of the B to Z transition in poly(dGdC)   总被引:1,自引:0,他引:1  
The thermodynamics of the B to Z transition in poly(dGdC) was examined by differential scanning calorimetry, temperature-dependent absorbance spectroscopy, and CD spectroscopy. In a buffer containing 1 mM Na cacodylate, 1 mM MgCl2, pH 6.3, the B to Z transition is centered at 76.4°C, and is characterized by ΔHcal = 2.02 kcal (mol base pair)?1 and a cooperative unit of 150 base pairs (bp). The tm of this transition is independent of both polynucleotide and Mg2+ concentrations. A second transition, with ΔHcal = 2.90 cal (mol bp)?1, follows the B to Z conversion, the tm of which is dependent upon both the polynucleotide and the Mg2+ concentrations. Turbidity changes are concomitant with the second transition, indicative of DNA aggregation. CD spectra recorded at a temperature above the second transition are similar to those reported for ψ(–)-DNA. Both the B to Z transition and the aggregation reaction are fully and rapidly reversible in calorimetric experiments. The helix to coil transition under these solution conditions is centered at 126°C, and is characterized by ΔHcal = 12.4 kcal (mol bp)?1 and a cooperative unit of 290 bp. In 5 mM MgCl2, a single transition is seen centered at 75.5°C, characterized by ΔHcal = 2.82 kcal (mol bp)?1 and a cooperative unit of 430 bp. This transition is not readily reversible in calorimetric experiments. Changes in turbidity are coincident with the transition, and CD spectra at a temperature just above the transition are characteristic of ψ(–)-DNA. A transition at 124.9°C is seen under these solution conditions, with ΔHcal = 10.0 kcal (mol bp)?1 and which requires a complex three-step reaction mechanism to approximate the experimental excess heat capacity curve. Our results provide a direct measure of the thermodynamics of the B to Z transition, and indicate that Z-DNA is an intermediate in the formation of the ψ-(–) aggregate under these solution conditions.  相似文献   

17.
Estimates of nuclear DNA base pair composition by determination of thermal denaturation temperatures (Tm) indicated guanine + cytosine (G + C) levels of 35–56% for 17 species of marine green algae. Tm values were found to be reproducible with coefficients of variation among samples and replicates of generally less than 1 percent. G + C % values in four species of Enteromorpha varied within a narrow range of 53–56%, whereas values for three species of Ulva showed substantially greater variation, ranging from 35–55%. Ulva fasciata collections from two geographically separate North Carolina sites had mean G + C composition of 44.8 and 35.6 respectively, suggesting that these populations may be genetically distinct. Enteromorpha linza, which has been treated as a species of Ulva, had a G + C composition of 53.2, typical of the Enteromorpha species investigated. Nuclear DNA base pair composition data for species of Cladophorales and Caulerpales are given as well.Center for Marine Science Research, UNC-W contribution No. 009.  相似文献   

18.
Tm values of 16 fully complementary RNA duplexes with repeating base sequence have been employed as the empirical basis for developing a reliable and practical method for computing apparent enthalpies (ΔH calc) for their helix → coil transitions. The approach taken is the same as in the accompanying investigation of DNA duplexes, although some of the computational variables of the “best-fit” function are necessarily different due to the distinguishing structural properties of the RNA-type helix. An excellent linear correlation was thus obtained between experimental Tm and ΔH calc values. An equally good fit was obtained between Tm and ΔH calc for five unrelated (to the 16 RNAs) decaribonucleotide duplexes. The differences in computational variables between the best-fit methods for RNA and DNA duplexes are shown to be a reflection of differences in cation binding and the effective local dielectric. The greater Tm dependence on G·C content of RNA helices than of DNA helices is shown to be due to a greater latitude of stacking stabilities of complementary dinucleotide fragments containing A·T than A·U base pairs.  相似文献   

19.
Detection of hydA genes of Clostridia spp. using degenerative and species specific primers for C. butyricum were optimized by the addition of bovine serum albumin (BSA) to polymerase chain reaction (PCR) and quantitative PCR (qPCR) reactions. BSA concentrations ranging from 100 to 400 ng/μl were examined using pure cultures and a variety of environmental samples as test targets. A BSA concentration of 100 ng/μl, which is lower than previously reported in the literature, was found to be most effective in improving the detection limit. The brightness of amplicons with 100 ng/μl BSA increased in ethidium bromide-treated gels, the minimum detection limit with BSA was at least one log greater, and cycle threshold (C T) values were lower than without BSA in qPCR indicating improved detection of target deoxyribonucleic acid for most samples tested. Although amplicon visualization was improved at BSA concentrations greater than or equal to 100 ng/μl, gene copy numbers detected by qPCR were less, CT values were increased, and T m values were altered. SYBR Green dissociation curves of qPCR products of DNA from pure culture or sludge samples showed that BSA at 100 ng/μl reduced the variability of peak areas and T m values.  相似文献   

20.
Plasmalemma-rich microsomal vesicles were prepared from whole leaf and acid-washed epidermal tissue of Vicia faba L. cv. Osnabrücker Markt by aqueous two-phase partitioning in dextran T-500 and polyethylenglycol 1350 aqueous phases. These vesicles were tightly sealed and predominantly right-side out, and contained a K+ -stimulated, mg2+-dependent and vanadate-sensitive ATPase. The enzyme from both tissues exhibited nearly identical properties: pH optimum 6.4, Km for ATP 0.60 mM(whole leaf) and 0.67 mM (epidermis). Vmax -480 nmol (mg protein)1 min1 (whole leaf) and 510 nmol (mg protein)1 min1 (epidermis), I50 (Na3,VO4) 7.5 μM (whole leaf) and 15 μM (epidermis). The enzyme was not inhibited by NO3(50 mM)or sodium azide (I mM). DCCD (20 μM) reduced enzyme activity to 50% (whole leaf) and 58% (epidermis), gramicidin S (20 μM) to 36% (whole leaf) and 41%(epidermis). Ca2+ inhibited the ATPase [I50, C2+: 0.5 mM(whole leaf) and 0.8 mM(epidermis)]. Ca2+ inhibited the ATPase [I50, C2+ 0.5 mM(whole leaf) und 0.8 (epidermis)]. The vanadate-sensitive ATPase from whole leaf and epidermal tissue was slightly but significantly stimulated by fusicoccin (FC) at a concentration (0.13 μM) promoting stomatal opening. The stimulation was not seen in the solubilized ATPase. Stomata of the cultivar used here were insensitive lo (±)ABA up to 2 μM level which is effective in most other cultivars and species. Likewise, at this concentration no effect of ABA on the activity of the epidermal ATPase was observed. The data are discussed with respect to the interaction of FC and ABA with the ATPase.  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