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 109 毫秒
1.
The antiphospholipid syndrome (APS) is a severe autoimmune disease associated with recurrent thrombosis and fetal loss and characterized by the presence of circulating autoantibodies (aAbs) mainly recognizing the N‐terminal domain (DmI) of β2‐glycoprotein I (β2GpI). To possibly block anti‐β2GpI Abs activity, we synthesized the entire DmI comprising residues 1–64 of β2GpI by chemical methods. Oxidative disulfide renaturation of DmI was achieved in the presence of reduced and oxidized glutathione. The folded DmI (N‐DmI) was purified by RP‐HPLC, and its chemical identity and correct disulfide pairing (Cys4‐Cys47 and Cys32‐Cys60) were established by enzymatic peptide mass fingerprint analysis. The results of the conformational characterization, conducted by far‐ and near‐UV CD and fluorescence spectroscopy, provided strong evidence for the native‐like structure of DmI, which is also quite resistant to both Gdn‐HCl and thermal denaturation. However, the thermodynamic stability of N‐DmI at 37°C was remarkably low, in agreement with the unfolding energetics of small proteins. Of note, aAbs failed to bind to plates coated with N‐DmI in direct binding experiments. From ELISA competition experiments with plate‐immobilized β2GpI, a mean IC50 value of 8.8 μM could be estimated for N‐DmI, similar to that of the full‐length protein, IC50(β2GpI) = 6.4 μM, whereas the cysteine‐reduced and carboxamidomethylated DmI, RC‐DmI, failed to bind to anti‐β2GpI Abs. The versatility of chemical synthesis was also exploited to produce an N‐terminally biotin‐(PEG)2‐derivative of N‐DmI (Biotin‐N‐DmI) to be possibly used as a new tool in APS diagnosis. Strikingly, Biotin‐N‐DmI loaded onto a streptavidin‐coated plate selectively recognized aAbs from APS patients.  相似文献   

2.
S K Podder 《Biopolymers》1972,11(7):1395-1410
The interaction between poly C and (Gp)nG(n = 1,2) in dilute solution was investigated spectrophotometrically in 0.1M phosphate buffer pH 7.2 under conditions unfavorable for the formation of self-associated complexes of oligoguanylic acids. Two isosbestic points were observed when poly C was titrated gradually with GpGpG, one at 232–233 mμ(in the range of 0–33% poly C) and one around 238 mμ (in the range of 50–100% poly C). The melting temperature (Tm) of the 1:1 poly C: (Gp)nG complexes (n = 1,2) of varying concentration were determined. The equilibrium properties of the 1:1 complexes can be described by two interaction parameters, namely, (i) cooperative stacking interaction between the first nearest neighbor of the adsorbed oligomer, and (ii) intrinsic association constant of the adsorbed oligomer with its polymeric site, since the cooperative helix–coil transition particularly in the smaller oligonucleotide can be described by an “all or none” model. Based on such a model the enthalpy of stacking inteaction-dependent Tm values yielded directly the sum of the enthalpy of stacking interaction and of basepairing (which is dependent on the chain length of the oligomer) and the value of S, the stability constant of a G–C pair within a helix. The enthalpy of formation of G–C pair is then calculated as ?6.3 kcal/base pair either from the chain length dependent enthalpy term or from the temperature coefficient of S values. From the S value and the association constant of 1:1 GpGpGpC:GpCpCpC complex, other thermodynamic parameters such as nucleation parameter (β) and free energy of stacking interaction can be obtained.  相似文献   

3.
Sodium ion activity was measured using a Na-glass electrode in a solution of poly(S-carboxymethyl-L -cysteine) with no added salt at various degrees of neutralization and various concentrations for samples of different molecular weights. The conformational change from random coil to the β-structure was detected from the activity coefficient of counterions, as well as from CD. At a constant degree of neutralization, the activity coefficient is insensitive to a concentration change not only in the random-coil state, but also in the range of conformational change if the concentration is below about 3 × 10?2 monomolal. At high concentrations of about 5 × 10?2 monomolal, however, the activity coefficient becomes low, probably due to the occurrence of the stacking of the pleated sheets.  相似文献   

4.
J Greve  M F Maestre  A Levin 《Biopolymers》1977,16(7):1489-1504
Circular dichroism (CD) spectra of poly(dA), poly(dT), poly(dA)·poly(dT), and poly[d(A-T)]·poly[d(T-A)] have been measured as a function of temperature. From these data difference spectra have been calculated by subtracting the spectrum measured at low temperature from the spectra measured at higher temperatures. The CD difference spectra obtained upon melting of the two double-stranded polymers are very similar. From a comparison of these difference spectra with calculated ones it is shown that optical transitions near 272 nm (on A) and 288 nm (most probably on T) are present. The premelting changes of the CD spectrum of poly[d(A-t)]·poly[d(T-A)] are due to a change in conformation in which the secondary structure goes from a C- to B-type spectrum by increasing the A-type nature of the polymer. Such a change is not observed for poly(dA)·poly(dT). Instead, a transition between two different B-type geometries occurs.  相似文献   

5.
The salt-induced B- to Z-DNA conformational transition is a cooperative- and time-dependent process. From a modified form of the logistic equation which describes an equilibrium between two states we have deduced a kinetic function to quantify the degree of the B to Z transition of a synthetic (dG-dC) ⋅ (dG-dC) polynucleotide. This function was obtained by introduction of time as a variable in the logistic function so that the equilibrium constant, K, is replaced by a new constant K s , characteristic of the type of salt used. This constant is defined as the salt concentration needed to reach the B-Z transition-midpoint in the time unit. The equation fits the data obtained by circular dichroism (CD) for changes in molecular ellipticity of poly(dG-m5dC) ⋅ poly(dG-m5dC) and poly(dG-dC) ⋅ poly(dG-dC) incubated with various concentrations of mono-, di-, and trivalent salts at a constant temperature. The derived expression may be a very useful tool for studying the kinetics of the B- to Z-DNA transition. Received: 1 December 1997 / Revised version: 16 March 1998 / Accepted: 27 March 1998  相似文献   

6.
Hiroshi Maeda  Kenta Ooi 《Biopolymers》1981,20(7):1549-1563
The β-coil transition of poly(S-carboxymethyl-L -cysteine) (poly[Cys(CH2CO2H)]) and poly(S-carboxyethyl-L -cysteine) (poly[Cys((CH2)2CO2H)]) was followed by CD, potentiometric titration, and viscosity in the absence of added salt. These different properties give consistent results for poly[Cys((CH2)2CO2H)]. The CD spectra of poly[Cys(CH2CO2H)] change considerably with the degree of neutralization α even for a low-molecular-weight sample incapable of forming the β-structure. Because of the superposition of this additional effect, the dependence of CD on α is inconsistent with titration data for the case of poly[Cys(CH2CO2H)], particularly when the nπ transition is used to follow the β-coil transition. The change of CD inherent to the β-coil transition is characterized by an isodichroic point: 215 nm for poly[Cys((CH2)2CO2H)] and 218 nm for poly[Cys(CH2CO2H)]. A criterion supporting the stacking of the pleated sheet is suggested based on the isodichroic point.  相似文献   

7.
The heat capacities of the single-stranded and double-stranded forms of polyadenylic acid, polyuridylic acid, and poly(uridylic and adenylic acid) were determined with a drop heat capacity calorimeter. In addition, the temperature dependence of the apparent partial heat capacity (?Cp) was measured with a newly developed differential scanning calorimeter. The calculated ΔCp at 28°C for the transition poly(A)·poly(A) ? 2 poly(A) was found to be 165 ± 24 cal/Kmol-base pair, compared with a value of 140 ± 28 for the transition poly(A)·poly(U) ? poly(A) + poly(U). The temperature dependence of ?Cp of single-stranded poly(U) was consistent with the conclusion that it is totally unstacked at temperatures above 15°C. The temperature dependence of ?Cp of single-stranded poly(A) was used to determine the base-stacking parameters for poly(A). The experimental results are consistent with a stacking enthalpy change of ?8.5 ± 0.1 kcal/mol bases and a cooperativity parameter σ of 0.57 ± 0.03 for the stacking of adenine bases. These results demonstrate that the heat capacity of single-stranded polynucleotides is greater than that of the double-stranded forms. This increased heat capacity is mainly the result of the temperature dependence of the base-stacking interactions in the single-stranded form.  相似文献   

8.
Abstract

CD spectra of poly(dA-dT)· poly(dA-dT) in low salt (10–100 mM NaCl) and high salt (4–6 M CsF) are different i.e. 275 nm band gets inverted in going from low to high salt (Vorhickova et. al.MarJ. Mol. Biol. 166, 85, 1983). However, from CD spectra alone it is not possible to decipher any structural differences that might exist between the low and high salt forms of poly(dA-dT)? poly(dA-dT). Hence, we took recourse to high resolution NMR spectroscopy to understand the structural properties of poly(dA-dT)? poly(dA-dT) in low and high salt. A detailed analysis of shielding constants and extensive use of NOE studies under minimum spin diffusion conditions using C(8)-deuterated poly(dA-dT)? poly(dA-dT) enabled us to come up with the following conclusions (i) base-pairing is Watson-Crick under low and high salt conditions, (ii) under both the conditions of salt the experimental data can be explained in terms of an equilibrium blend of right and left-handed B-DNA duplexes with the left-handed form 70% and the right-handed 30%. In a 400 base pairs long poly(dA-dT)? polyidA-dT) (as used in this study), equilibrium between right and left-handed helices can also mean the existence of both helical domains in the same molecule with fast interchange between these domains or/and unhindered motion/propagation of these domains along the helix axis, (iii) However, there are other structural differences between the low and high salt forms of poly(dA-dT) ? poly(dA-dT); under the low salt condition, right-and left-handed B-DNA duplexes have mononucleotide as a structural repeat while under the high salt conditions, right-and left-handed B-DNA duplexes have dinucleotide as a structural repeat. In the text we provide the listing of torsion angles for the low and high salt structural forms, (iv) Salt (CsF) induced structural transition in poly(dA-dT)? poly(dA-dT) occurs without any breakage of Watson- Crick pairing, (v) The high salt form of poly(dA-dT)? poly(dA-dT) is not the left-handed Z-helix.

Although the results above from NMR data are quite unambiguous, a question still remains i.e. what does the salt (CsF) induced change in the CD spectra of poly(dA-dT)? poly(dA-dT) really indicate? Interestingly, we could show that the salt (CsF) induced change in poly(dA-dT)? poly(dA-dT) is quite similar to that caused by a basic polypeptide viz. poly-L(Lys2-Ala)n i.e. both the agents induced a ψ-structure in DNA. And it was also demonstrated that the changes in poly(dA-dT)? poly(dA-dT) as caused by CsF and poly-L-(Lys2-Ala)n could be reverted back by ethidium bromide-a relaxing agent.

To minimize complications from spin diffusion in this study we have used very small presaturation pulse lengths and C(8)-deuterated poly(dA-dT)? poly(dA-dT) of 400 ± 150 bp long. Even though deuteration of a primary site of diffusion such as C(8)H substantially decreases diffusion, in order to make sure that our conclusions are not compromised by possible diffusion in such a long fragment under small presaturation times, we have repeated our experiments using the six base pair long duplex of d(A-T-A-T-A-T) and found the results to be strikingly similar to that from the polymer.  相似文献   

9.
The binding of initiation Factor 3 (IF3) to poly (1,N6-ethenoadenylic acid) [poly(epsilon A)] was investigated by fluorescence spectroscopy. At low salt concentrations, IF3 evokes an increase in the fluorescence intensity of poly(epsilon A) due to the unstacking of the nucleotide bases. The poly(epsilon A) fluorescence enhancement titrates to an endpoint of 13 +/- 2 nucleotide residues per IF3. The maximum poly(epsilon A) fluorescence enhancement, at lattice saturation, decreases with increasing salt concentration. Even though IF3 does not produce a large fluorescence increase between 75 and 200 mM NaCl concentration, the protein still binds to poly(epsilon A) at these salt concentrations as measured by sedimentation partition chromatography; the value of Kobs for the IF3-poly(epsilon A) interaction is comparable to that of other synthetic polynucleotides. The binding of IF3 to poly(A) at 150 and 200 mM NaCl induces an increase in nucleotide base-base separation as determined by CD, yet IF3-induced disruption of base stacking of poly(epsilon A) at these same salt concentrations is not detected by fluorescence. It is likely that IF3 binds primarily to the phosphate backbone of poly(epsilon A) at low salt concentrations, producing an increase in the fluorescence intensity. But, at higher salt concentrations, the aromatic amino acids intercalate between the nucleotide bases quenching the poly(epsilon A) fluorescence.  相似文献   

10.

Background

RNA has attracted recent attention for its key role in gene expression and hence targeting by small molecules for therapeutic intervention. This study is aimed to elucidate the specificity of the alkaloid coralyne to poly(G), poly(C), poly(I) and poly(U) in the light of its ability in inducing self-structure in poly(A).

Methods

Multifaceted experimental techniques like competition dialysis, absorption, fluorescence, circular dichroism and calorimetry were employed. Salt dependence and temperature dependence of the binding was also elucidated.

Results

Results of competition dialysis, absorption and fluorescence studies revealed that coralyne binds strongly to the polypurines, poly(G) and poly(I) compared to the polypyrimidines, poly(U) and poly(C). Partial intercalative binding due to the stacking of the molecules between the bases was envisaged. The binding was predominantly enthalpy driven with favourable entropy term with a large favourable non-electrostatic contribution revealed from salt dependent data and the dissection of the free energy. The heat capacity change of − 125 and − 119 cal/mol K− 1 respectively for poly(G) and poly(I) and the partial enthalpy–entropy compensation phenomenon observed confirmed the involvement of multiple weak noncovalent interactions. Circular dichroism studies provided evidence for significant perturbation of the conformation of the RNAs, but no self-structure induction was evident in any of the polymers under the condition of the study.

Conclusions

This study presents a complete structural and thermodynamic profile of coralyne interaction to four single stranded RNA polymers.

General significance

The study for the first time elucidates the base specificity of coralyne–RNA complexation at the single stranded level.  相似文献   

11.
The conformation of the valinomycin–lithium complex has been studied using CD and nmr techniques. The lithium ion induced significant changes in the chemical shifts of the NH and CαH protons, as well as in the CD spectra of valinomycin. From the analysis of the lithium ion titration data, it is concluded that valinomycin forms a 1:1 type weak complex with lithium, having a stability constant of 48 L mol?1 at 25°C. This conformation is different from the familiar valinomycin–potassium complex. The nature of the interaction at low and high concentrations of lithium ions with valinomycin (ionophore) and gramicidin-S (nonionophore) has been compared. At high salt concentrations, there was a further change in the CD and nmr spectra of valinomycin, giving a second plateau region at > 3M of the salt. In the case of gramicidin-S, no significant changes in the nmr or CD spectra were observed in the lower concentration range corresponding to where changes were observed in the case of valinomycin. However, the addition of lithium salt at concentrations greater than 3M induced changes in both the CD and nmr spectra of gramicidin-S, and the titration graph of molar ellipticity versus concentration of lithium perchlorate gave a plateau region at concentrations greater than this. These results indicate that the effects of lithium at low and high concentrations are independent of each other. The conformational transitions at very high salt concentrations (denaturation) are more likely due to solvent structural perturbations rather than to the consequences of ion binding.  相似文献   

12.
Circular dichroism (CD) curves are reported for poly dA, (pdA)6, (pdA)2, poly A, ApAp, ApA, AMP, dApA, pdApA, A-2′-O-methyl pA, and A-2′-O-methyl pAp. Analysis of these curves indicated the presence of single CD bands at 228–230 mμ and at 278–280 mμ in oligomers longer than dinucleotides. In the case of dinucleotides and mononucleotides (from the literature, in addition to those studied here), the 230 mμ CD of band appears but the 280 mμ CD band does not. We assign the 230 mμ band to a very weak π–π* transition at this wavelength. From theoretical considerations, we show that the 280 mμ band is not an exciton component of the strong π–π* transition at 260 mμ in adenine. We conclude that the 280 mμ CD band must be assigned to a distinct absorption, not previously reported, which we suggest arises from an n–π* transition. The fact that the n–π* CD band at 280 mμ is not seen in mononucleotides or dinucleotides is ascribed to solvation of the adenine ring by water, which shifts the band to shorter wavelengths. Therefore, only interior residues of oligomers have the 280 mμ band, and the optical activity of a polymer cannot be computed from that of a dinucleotide, by using a nearest-neighbor approximation. The existence of this end effect hag been tested, by taking it into account in computing the rotational strengths of the 278 mμ n–π* transition for several oligomers; it is pointed out that a more sensitive test of this end effect would require CD data for the oligo dA series of 3 to 5 residues. We speculate about the structural and optical differences between poly dA and poly A, and point out the need for a theoretical treatment of n–π* Cotton effects in polynucleotides.  相似文献   

13.
Epitopes recognized by five mAb which block the binding of human IgE antibodies to grass group I (GpI) Ag were characterized and partially mapped. Site specificity studies defined four apparently non-overlapping blocking antibody binding sites on the meadow fescue GpI molecule, Fes e I. One of these sites (site A) was localized to a 14,000 m.w. fragment designated P3 generated by CNBr cleavage of purified Fes e I. The P3 peptide possessed human IgE binding sites as well as other epitopes (non-site A) defined by 19 other anti-GpI mAb. All of the P3 reactive antibodies recognized cross-reactive determinants found on GpI Ag isolated from five different grasses suggesting that P3 is a conserved portion of grass GpI molecules. The P3 fragment from Fes e I was used to immunize mice and induced antibodies which reacted with intact GpI Ag from all 5 different grasses currently being studied in this laboratory.  相似文献   

14.
The CD spectra and melting profiles have been measured for nine synthetic double-stranded RNAs containing I · C instead of G · C base pairs: poly[r(I) · r(C)], poly[r(I-C) · r(I-C)], poly[r(A-I-C) · r(I-C-U)], poly[r(A-C) · r(I-U)], poly[r(A-I) · r(C-U)], poly[r(A-C-C) · r(I-I-U)], poly[r(A-A-C) · r(I-U-U)], poly[r(A-C-U) · r(A-I-U)], and poly[r(A-U-C) · r(I-A-U)]. CD spectra have not previously been reported for the latter six of these polymers. The substitution of inosinate for guanylate led to recognizable CD differences, with all but two of the polymers having two resolved positive bands above 230 nm. Also, the I-containing RNAs differed from their G-containing counterparts in the almost complete absence of negative CD bands at long wavelengths and in the reduction of negative CD bands near 210 nm. First-neighbor comparisons showed that the CD spectra of the I-containing RNAs were consistent with the nearest-neighbor sequences of the polymers, as previously shown for G-containing RNAs (D. M. Gray, J.-J. Liu, R. L. Ratliff, and F. S. Allen, Biopolymers (1981) 20 , 1337–1382). Moreover, two of the first-neighbor comparisons involved spectra of poly[r(A) · r(U)] and poly[r(I) · r(C)], polymers known to be in the A family of conformations in fibers (S. Arnott, D. W. L. Hukins, S. D. Dover, W. Fuller, and A. Hodgson, (1973) J. Mol. Biol. 81 , 107–122). Thus, differences in the CD spectra of I- and G-containing RNAs could be simply explained as resulting from differences in the hypoxanthine and guanine chromophores, without invoking differences in conformation. Finally, melting temperatures of the I-containing RNAs were found to vary much less with base composition than do the melting temperatures of G-containing RNAs, since A · U base pairs are closer to I · C than to G · C base pairs in stability.  相似文献   

15.
16.
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   

17.
H Y Wu  M J Behe 《Nucleic acids research》1985,13(11):3931-3940
Salt induced transitions between four conformations of the methylated ribo-deoxyribo co-polymer poly (rG-m5dC).poly (rG-m5dC) have been studied using phosphorous-NMR, Raman spectroscopy, and circular dichroism. A high salt A-Z transition is observed for the polymer. However, the methylated polymer does not enter the high salt Z form more readily than the analogous unmethylated polymer, unlike the effect of methylation on the fully deoxy polymer poly (dG-dC).poly (dG-dC). The methylated polymer fails to undergo a low salt A-Z transition in 5 mM Tris buffer, unlike the unmethylated poly (rG-dC).poly (rG-dC). However, if the counterion is changed to triethanolamine buffer, an A-Z transition does take place. In 5 mM Tris buffer the phosphorous-NMR spectrum of poly (rG-m5dC).poly (rG-m5dC) shows one resonance in the absence of NaCl that splits into two closely spaced resonances as the NaCl level is increased to 30 mM. The Raman spectrum of poly (rG-m5dC).poly (rG-m5dC) shows that it is in the A conformation at intermediate salt concentrations. From this we conclude that poly (rG-m5dC).poly (rG-m5dC) is in a regular A conformation in Tris buffer at low Na+ levels, shifting to an alternating A conformation with a dinucleotide repeat at intermediate salt concentrations.  相似文献   

18.
Abstract

Using CD measurements we show that the interaction of netropsin to poly(dA-dT)·poly(dA-dT) involves two binding modes at low ionic strength. The first and second binding modes are distinguished by a defined shift of the CD maximum and the presence of characteristic isodichroic points in the long wavelength range from 313 nm to 325 nm. The first binding mode is independent of ionic strength and is primarily determined by specific interaction to dA·dT base pairs. Employing a netropsin derivative and different salt conditions it is demonstrated that ionic contacts are essential for the second binding mode. Other alternating duplexes and natural DNA also exhibit more or less a second step in the interaction with netropsin observable at high ratio of ligand per nucleotide. The second binding mode is absent for poly(dA)·poly(dT). The presence of a two-step binding mechanism is also demonstrated in the complex formation of poly(dA-dT)·poly(dA-dT) with the distamycin analog consisting of pentamethylpyrrolecarboxamide. While the binding mode I of netropsin is identical with its localization in the minor groove, for binding mode II we consider two alternative interpretations.  相似文献   

19.
Design and synthesis of peptides capable of specific binding to DNA   总被引:1,自引:0,他引:1  
In the present communication, design, synthesis and DNA binding activities of the following two peptides are reported: Dns-Gly-Ala-Gln-Lys-Leu-Ala-Cly-Lys-Val-Gly-Thr-Lys-Val-Lys-Val-Gl y-Thr-Lys-Thr - Val-OH (I) and [(H-Ala-Lys-Leu-Ala-Thr-Lys-Ala-Gly-Val-Lys-Gln-Gln-Ser-Ile-Gln-Leu-Ile- Thr- Ala-Aca-Lys-Aca)2Lys-Aca]2Lys-Val-OH (II), where Aca = NH(CH2)5CO--; Dns is a residue of 5-dimethylaminonaphtalene-1-sulfonic acid. Peptide I contains a large fraction (ca.30%) of valyl and threonyl residues, which possess a high potential for beta structure formation. Peptide II contains four repeats of the amino acid sequence present in the presumed DNA binding helix-turn-helix unit of 434 Cro repressor. These four domains are linked in such a way that two domains can interact with two halves a 14 base pair long operator site on DNA. From CD studies we have found that peptide I is in a random coil conformation in the aqueous solution in the presence of 20% trifluoroethanol. By contrast, amino acid residues of peptide II assume alpha helical, beta and random coiled conformations under the same conditions. A change in the secondary structure of the two peptides upon binding to DNA is observed. The difference CD spectra obtained by subtracting the spectra of free DNA from the spectra of peptide I--DNA complexes gives rise to a beta-like pattern. The difference CD spectra obtained for complexes of peptide II with various natural and synthetic DNAs suggest that alpha-beta-transition takes place in the presumed helix-turn-helix repeat units of peptide II upon binding to DNA. Peptide I binds more strongly to poly(dG).poly(dC) than to poly(dA).poly(dT) and poly[d(GC)].poly[d(GC)]. The binding takes place in the minor DNA groove because minor groove binding antibiotic sibiromycin can displace peptide I from a complex with poly(dG).poly(dC). Analysis of footprinting diagramms shows that peptide I specifically protects phosphodiester bonds within operator sites OR1 and OR2 of phage lambda from nuclease cleavage. By contrast, peptide II does not react specifically with operators OR1, OR2 and OR3 of phage 434 although it forms very tight complexes with DNA which are stable in the presence of 1M NH4F.  相似文献   

20.
The acid-base titration (pH 8 --> pH 2.5 --> pH 8) of eleven mixing curve samples of the poly(dG) plus poly(dC) system has been performed in 0.15 M NaCl. Upon protonation, poly(dG).poly(dC) gives rise to an acid complex, in various amounts according to the origin of the sample. We have established that the hysteresis of the acid-base titration is due to the non-reversible formation of an acid complex, and the liberation of the homopolymers at the end of the acid titration and during the base titration: the homopolymer mixtures remain stable up to pH 7. A 1G:1C stoichiometry appears to be the most probable for the acid complex, a 1G:2C stoichiometry, as found in poly(C(+)).poly(I).poly(C) or poly(C(+)).poly(G).poly(C), cannot be rejected. In the course of this study, evidence has been found that the structural consequences of protonation could be similar for both double stranded poly(dG).poly(dC) and G-C rich DNA's: 1) protonation starts near pH 6, dissociation of the acid complex of poly(dG).poly(dC) and of protonated DNA take place at pH 3; 2) the CD spectrum computed for the acid polymer complex displays a positive peak at 255 nm as found in the acid spectra of DNA's; 3) double stranded poly(dG).poly(dC) embedded in triple-stranded poly(dG).poly(dG).poly(dC) should be in the A-form and appears to be prevented from the proton induced conformational change. The neutral triple stranded poly(dG).poly(dG).poly(dC) appears therefore responsible, although indirectly, for the complexity and variability of the acid titration of poly(dG).poly(dC) samples.  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