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 262 毫秒
1.
The hysteresis observed in cyclic acid-base titrations of the three-standed polyribonucleotide helix poly (A)-2 POLY (U) strongly depends on ionic strength. For NaCl and at 25 degrees C, hysteresis occurs in the limited concentration range between 0.03 M and 1.0 M(NaCl). The transition points associated with the cyclic conversions between the triple helix and the poly (A)-poly (A) double helix and (free) poly (U) constitute a (pH ionic strength) phase diagram covering the ranges of stability and metastability of the hysteresis system. Variations with NaCl concentration of some hysteresis parameters can be quantitatively described in terms of polyelectrolyte theories based on the cylinder-cell model for rodlike polyions. The results of this analysis suggest that the metastability is predominantly due to dlectrostatic energy barriers preventing the equilibrium transition of the partially protonated triple helix above a critical pH value. Ultraviolet absorbance and potentiometric titration data of poly (A)in the acidic pH range can be analyzed in terms of two types of double-helical structures. Spectrophotometric titrations reveal isosbestic wavelengths for structural transitions of poly (A). "Time effects" commonly observed in poly (A) titrations are suggested to reflect helix transitions between the two acidic structures.  相似文献   

2.
L C Rainen  B D Stollar 《Biochemistry》1977,16(9):2003-2007
Rabbit antibodies to the triple-helical polynucleotide poly(A)-poly(U)-poly(I) were fractionated into three major antibody populations, each recognizing a different conformational feature of the triple-helical immunogen. Two distinct populations were purified from precipitates made with poly(A)-poly(U)-poly(U) and poly(A)-poly(I)-poly(I). The former reacted with double-stranded poly(A)-poly(U) or poly(I)-poly(C), and similar populations could be purified with either double-stranded form. The second population recognized the poly(A)-poly(I) region of the triple helix, and the third required all three strands for reactivity. These immunochemical studies suggest that the poly(A) and poly(U) have the same orientation in the triple-helicical poly(A)-poly(U)-poly(I) as in the double-helical poly(A)-poly(U), in which they have Watson-Crick base pairing.  相似文献   

3.
Some properties of unprimed poly(A)-poly(U) synthesis by DNA-dependent RNA polymerase from Caulobacter crescentus were examined. The reaction required ATP and UTP as substrates and manganese as a divalent cation. Rifampicin completely inhibited the reaction at a concentration of 1 micron/ml, and the enzyme catalyzed the polymer synthesis well regardless of the presence of GTP, CTP or both. The chain length of the poly(A)-poly(U) synthesized was about one hundred base pairs, as estimated from a sedimentation velocity and the molar ratio of [3H]AMP to [gamma-32P]ATP incorporated into the poly(A)-poly(U). The reaction was dependent on the square of the enzyme concentration and the enzyme dimers formed complexes with poly(A)-poly(U) during the reaction.  相似文献   

4.
The rate of double helix formation by single-stranded poly A plus poly dT, poly dA plus poly U, poly dA plus dT, poly G plus poly dC, poly dG plus poly C, and poly dG plus poly dC have been investigated and compared to rates of ribohomopolymer helix formation rates. After correction for molecular weight, comparisons of rate data at 30°C below the melting temperature of the double helix show that:
  • 1 Rates of helix formation by all combinations of guanine plus cytosine homopolymers are the same.
  • 1 The rate of helix formation for poly dA plus poly dT is three times faster than the rate for poly A plus poly U. Rates of formation of DNA-RNA hybrid molecules are intermediate between these two rates, but closer to the poly dA plus poly dT rate.
The effect of temperature on the rate of helix formation is interpreted in terms of a steady-state model for helix propagation. The results are consistent with a mechanism in which the formation of the second base pair is the rate-determining step.  相似文献   

5.
The Zimm-Bragg theory is extended to treat the melting of the triple helix poly (A + 2U) for a solution with a 1 : 2 mole ratio of poly A to poly U. Only the case for long chains is considered. For a given set of parameters the theory predicts the fraction of segments in the triple helix, double helix, and random coil states as a function of temperature. Four nucleation parameters are introduced to describe the two order–disorder transitions (poly (A + 2U) ? poly A + 2 poly U and poly (A + U) ? poly A + poly U) and the single order–order transition (poly (A + 2U) ? poly (A + U) + poly U). A relation between the nucleation parameters is obtained which reduces the number of independent parameters to three. A method for determining these parameters from experiment is presented. From the previously published data of Blake, Massoulié and Fresco8 for [Na+] = 0.04, we find σT = 6.0 × 10?4, σD = 1.0 × 10?3, and σσ* = 1.5 × 10?3. σT and σD are the nucleation parameters for nucleating a triple helix and double helix, respectively, from a random coil region. σσ* is the nucleation parameter for nucleating a triple helix from a double helix and a single strand. Melting curves are generated from the theory and compared with the experimental melting curves.  相似文献   

6.
Hydrogen-exchange studies of I · C and G · C double helices were carried out to test the generality of conclusions reached previously in studies of adenine-containing polymers (preceding paper). The cytosine amino group shows hydrogen-exchange behavior similar to the analogous group in adenine; a pH-independent pathway and a parallel general catalysis pathway require prior separation of the base-pair and pre-equilibrium protonation at the ring N. The cytosine amino group does, however, display greater sensitivity to specific and to general catalysis than found for adenine. In the G · C helix, the ring NH proton of guanine exchanges at the opening-limited rate, as does the analogous proton in A · U and A · T pairs, while the guanine amino protons exchange without a prior opening of structure. From the observed exchange rates and the known chemistry for the pH-independent reaction, one can calculate equilibrium opening constants of 4 × 10−3 for poly(rI) · poly(rC) and perhaps one tenth of that for poly(rG) · poly(rC). Also the opening rate constant for the G · C helix is 0.01 s−1.These results, when applied to published exchange curves for DNA, indicate an equilibrium opening constant of 0.005, an opening rate constant of 0.04 s−1, and a closing rate constant of 10 s−1. (All values refer to studies at 0 °C.) These values point to the same kind of traveling-loop model for base-pair opening discussed previously for the opening reactions in adenine-containing double helices.  相似文献   

7.
Using Raman spectroscopy, we examined the ribose-phosphate backbone conformation, the hydrogen bonding interactions, and the stacking of the bases of the poly(U).poly(A).poly(U) triple helix. We compared the Raman spectra of poly(U).poly(A).poly(U) in H2O and D2O with those obtained for single-stranded poly(A) and poly(U) and for double-stranded poly(A).poly(U). The presence of a Raman band at 863 cm-1 indicated that the backbone conformations of the two poly(U) chains are different in the triple helix. The sugar conformation of the poly(U) chain held to the poly(A) by Watson-Crick base pairing is C3' endo; that of the second poly(U) chain may be C2' endo. Raman hypochromism of the bands associated with base vibrations demonstrated that uracil residues stack to the same extent in double helical poly(A).poly(U) and in the triple-stranded structure. An increase in the Raman hypochromism of the bands associated with adenine bases indicated that the stacking of adenine residues is greater in the triple helix than in the double helical form. Our data further suggest that the environment of the carbonyls of the uracil residues is different for the different strands.  相似文献   

8.
When guanine and uracil form hydrogen bonds in the pairing scheme first proposed by Crick one would expect that poly(A,G) will form an unperturbed double helix with poly U at room temperature in a dilute electrolyte solution (0.1 M NaCl). We have demonstrated by Raman- and IR-spectroscopy that the secondary structure of poly(A.G) · poly U is very similar to the structure of poly A · poly U; only the thermal stability of the double helix seems slightly lower than the stability of poly A · poly U, whereas the average helix length is unaffected by the dispersed G · U base pairs. From our input ratio of guanine and adenine we estimate that about every fourth base pair is a wobble pair.  相似文献   

9.
Detailed models are presented for the triple-stranded polynucleotide helices of poly (U)-poly (A)-poly (U) (two forms), poly (U)-poly d (A) -poly (U), poly d(C)-poly d(I)-poly d(C), poly d(T)-polyd(A)-poly d(T) and poly (I)-poly (A)-poly (I). The models were genrated using a computerized, linked-atom procedure which preserves standard bond lengths, bond anglesand sugar ring conformations, constrains the helices to have the pitches and symmetries observed in X-ray diffraction experiments, and optimises the non-bonded interatomic contacts including hydrogen bonds. The possible biological sigificance of such complexes is discussed.  相似文献   

10.
The temperature dependence to the 31P NMR spectra of poly[d(GC)] . poly [d(GC)],d(GC)4, phenylalanine tRNA (yeast) and mixtures of poly(A) + oligo(U) is presented. The 31P NMR spectra of mixtures of complementary RNA and of the poly d(GC) self-complementary DNA provide torsional information on the phosphate ester conformation in the double, triple, and "Z" helix. The increasing downfield shift with temperature of the single-strand nucleic acids provides a measure of the change in the phosphate ester conformation in the single helix to coil conversion. A separate upfield peak (20-60% of the total phosphates) is observed at lower temperatures in the oligo(U) . poly(A) mixtures which is assigned to the double helix/triple helix. Proton NMR and UV spectra confirm the presence of the multistrand forms. The 31P chemical shift for the double helix/triple helix is 0.2-0.5 ppm upfield from the chemical shift for the single helix which in turn is 1.0 ppm upfield from the chemical shift for the random coil conformation.  相似文献   

11.
It has been shown earlier that the DNA double helix is opened due to a prolonged contact of the DNA molecule with the surface of the mercury electrode. At neutral pH, the opening process is relatively slow (around 100 s), and it is limited to potentials close to -1.2 V (against SCE). The opening of the double helix has been explained by strains in the DNA molecule due to strong repulsion of the negatively charged phosphate residues from the electrode surface where the polynucleotide chain is anchored via hydrophobic bases. Interaction of the synthetic ds polynucleotides with alternating nucleotide sequences/poly(dA-dT).poly (dA-dT), poly (dA-dU).poly (dA-dU), poly (dG-dC).poly (dG-dC)/ and homopolymer pairs/poly (dA).poly (dT), poly (rA).poly (rU) and poly (dG).poly (dC)/ with the hanging mercury drop electrode has been studied. Changes in reducibility of the polynucleotides were exploited to indicate opening of the double helix. A marked difference in the behaviour was observed between polynucleotides with alternating nucleotide sequence and homopolymer pairs: opening of the double-helical structures of the former polynucleotides occurs at a very narrow potential range (less than 100 mV) (region U), while with the homopolymer pairs containing A X T or A X U pairs, the width of this region is comparable to that of natural DNA (greater than 200 mV). In contrast to natural DNA, the region U of homopolymer pairs is composed of two distinct phases. No region U was observed with poly (dG).poly (dC). In polynucleotides with alternating nucleotide sequence, the rate of opening of the double helix is strongly dependent on the electrode potential in region U, while in homopolymer pairs, this rate is less potential-dependent. It has been assumed that the difference in the behaviour between homopolymer pairs and polynucleotides with alternating nucleotide sequence is due to differences in absorbability of the two polynucleotide chains in the molecule of a homopolymer pair (resulting from different absorbability of purine and pyrimidine bases) in contrast to equal adsorbability of both chains in a polynucleotide molecule with alternating nucleotide sequence. It has been shown that the mercury electrode is a good model of biological surfaces (e.g. membranes), and that the nucleotide sequence-dependent opening (unwinding) of the DNA double helix at electrically charged surfaces may play an important role in many biological processes.  相似文献   

12.
Abstract

Using Raman spectroscopy, we examined the ribose-phosphate backbone conformation, the hydrogen bonding interactions, and the stacking of the bases of the poly(U)·poly(A) ·poly(U) triple helix. We compared the Raman spectra of poly(U)·poly(A)·poly(U) in H2O and D2O with those obtained for single-stranded poly(A) and poly(U) and for double-stranded poly(A)·poly(U). The presence of a Raman band at 863 cm?1 indicated that the backbone conformations of the two poly(U) chains are different in the triple helix. The sugar conformation of the poly(U) chain held to the poly(A) by Watson-Crick base pairing is C3′ endo; that of the second poly(U) chain may be C2′ endo. Raman hypochromism of the bands associated with base vibrations demonstrated that uracil residues stack to the same extent in double helical poly(A)·poly(U) and in the triple-stranded structure. An increase in the Raman hypochromism of the bands associated with adenine bases indicated that the stacking of adenine residues is greater in the triple helix than in the double helical form. Our data further suggest that the environment of the carbonyls of the uracil residues is different for the different strands.  相似文献   

13.
T R Fink  D M Crothers 《Biopolymers》1972,11(1):127-136
We prepared helices with noncomplementary bases by N1-oxidation of poly A, followed by reaction with poly U. Mixing curves indicate that doubly and triply helical structures form, with only the unmodified adenines involved in base pair formation. Circular dichroism spectra were examined particularly at the absorbance maximum of the adenine N1-oxide (A*). In the single strand poly (A,A*), there is a relatively strong pair of positive and negative CD bands from the A*. These are greatly reduced in the double helix, and abolished in the triple helix. We conclude that A* stacks in a conventional manner with A in the single strand, but is rotated out of the double and triple helix. In the double helix the A* probably maintains a preferred orientation with respect to the helix, but rotates randomly in the triple helix.  相似文献   

14.
The Green's function technique is applied to a study of breathing modes in a DNA double helix which contains a region of different base pairs from the rest of the double helix. The calculation is performed on a G-C helix in the B conformation with four consecutive base pairs replaced by A-T. The average stretch in hydrogen bonds is found amplified around the A-T base pair region compared with that of poly(dG)-poly(dC). This is likely related to the A-T regions lower stability against hydrogen bond melting. The A-T region may be considered to be the initiation site for melting in such a helix.  相似文献   

15.
An ethidium-induced double helix of poly(dA)-poly(rU).   总被引:4,自引:2,他引:2       下载免费PDF全文
Equilibrium dialysis, relaxation kinetic, melting, and continuous variation mixing experiments on complexes of poly(dA) and poly(rU) demonstrate that ethidium induces conversion of a 1:1 mixture of these homopolymers (at one molar salt and 19 degrees C) from a three stranded to a two stranded helix. This is the first demonstration of a double helix of poly(dA)-poly(rU) in solution.  相似文献   

16.
The binding of echinomycin to deoxyribonucleic acid.   总被引:20,自引:4,他引:16       下载免费PDF全文
Echinomycin is a peptide antibiotic which binds strongly to double-helical DNA up to a limit of approximately one molecule per five base-pairs. There is no detectable interaction with rRNA and only extremely feeble non-specific interaction with poly(rA)-poly(rU). Heat denaturation of DNA greatly decreases the binding, and similarly limited interaction is observed with naturally occurring single-stranded DNA. Association constants for binding to nine double-helical DNA species from different sources are presented; they vary by a factor of approximately 10, but are not simply related to the gross base composition. The interaction with DNA is ionic-strength-dependent, the binding constant falling by a factor of 4 when the ionic strength is raised from 0.01 to 0.10mol/litre. From the effect of temperature on the association constant for calf thymus DNA, the enthalpy of interaction is calculated to be about -13kJ/mol (-3kcal/mol). Binding of echinomycin persists in CsCl gradients and the buoyant density of nicked bacteriophage PM2 DNA is decreased by 25 mg/ml. Echinomycin interacts strongly with certain synthetic poly-deoxynucleotides, the binding constant decreasing in the order poly(dG)-poly(dC) greater than poly(dG-dC) greater than poly(dA-dT). For the latter two polymers the number of base-pairs occluded per bound antibiotic molecule is calculated to be three, whereas for poly(dG)-poly(dC) it is estimated to be four to five. Poly(dA)-poly(dT) and poly(dI)-poly(dC) interact only very weakly with the antibiotic. Poly(dI-dC) interacts to a slightly greater extent, but the binding curve is quite unlike that seen with the three strongly binding synthetic polynucleotides. Echinomycin affects the supercoiling of closed circular duplex bacteriophage PM2 DNA in the characteristic fashion of intercalating drugs. At low ionic strength the unwinding angle is almost twice that of ethidium. Likewise the extension of the helix, determined from changes in the viscosity of rod-like sonicated DNA fragments, is nearly double that expected for a simple (monofunctional) intercalation process. On this basis the interaction process is characterized as bifunctional intercalation. At higher ionic strength the unwinding angle relative to that of ethidium and the helix extension per bound echinomycin molecule fall, indicating a smooth progression towards more nearly monofunctional intercalation. Two simpler compounds which act as analogues of the quinoxaline chromophores of echinomycin, quinoxaline-2-carboxamide and the trypanocidal drug Bayer 7602, interact with DNA very much more weakly than does echinomycin, showing that the peptide portion of the antibiotic plays an essential role in determining the strength and specificity of the interaction.  相似文献   

17.
The polymerization of 2-fluoroadenosine 5'-diphosphate by polynucleotide phosphorylase to give high molecular weight poly(2-fluoroadenylic acid), poly(fl2A), is described. Both the single-stranded and double-stranded (acid) forms of poly(fl2A) exhibit strikingly similar ultraviolet and circular dichroism spectra to those of poly(A), and the enzymatic polymerization rates and thermal hyperchromicities of the two polymers are also very similar. However, the pKa of poly(fl2A) for protonation at N-1 is 2.9 compared to 5.9 for poly(A) under similar conditions. Poly(fl2A) forms a triple-stranded helix with poly(U) which has ultraviolet and cd spectra very reminiscent of poly(A) . 2 poly(U), but no conditions could be found which permitted the formation of a double helix. In the Escherichia coli ribosome system poly(fl2A) codes for the synthesis of polylysine, as does poly(A), although the rate and extent of incorporation were less in the former case. The role of basicity of adenine N-1 in these interactions is discussed.  相似文献   

18.
The inclination angle between the base normal and the helix axis is measured for ribo-AU polymers by using flow linear dichroism (LD), and compared to measurements for deoxyribo-AT polymers under dehydrating conditions. The CD of the DNA polymers under the dehydrating conditions is not the same as the corresponding RNA polymers, which are presumed to be in the A form. However, the LD indicates that poly(dAdT)-poly(dAdT) can assume the A form in 80% 2,2,2-trifluoroethanol, although poly(dA)-poly(dT) retains B form structure in this dehydrating solvent. The inclination angles are similar for B form poly(dAdT)-poly(dAdT) and poly(dA)-poly(dT), and these parameters are also similar for A form poly(rArU) -poly(rArU) and poly(rA) -poly(rU). All the inclination axes are similar. © 1995 John Wiley & Sons, Inc.  相似文献   

19.
Polyglutamine (poly(Q)) expansion is associated with protein aggregation into β-sheet amyloid fibrils and neuronal cytotoxicity. In the mutant poly(Q) protein huntingtin, associated with Huntington's disease, both aggregation and cytotoxicity may be abrogated by a polyproline (poly(P)) domain flanking the C terminus of the poly(Q) region. To understand structural changes that may occur with the addition of the poly(P) sequence, we synthesized poly(Q) peptides with 3-15 glutamine residues and a corresponding set of poly(Q) peptides flanked on the C terminus by 11 proline residues (poly(Q)-poly(P)), as occurs in the huntingtin sequence. The shorter soluble poly(Q) peptides (three or six glutamine residues) showed polyproline type II-like (PPII)-like helix conformation when examined by circular dichroism spectroscopy and were monomers as judged by size-exclusion chromatography (SEC), while the longer poly(Q) peptides (nine or 15 glutamine residues) showed a β-sheet conformation by CD and defined oligomers by SEC. Soluble poly(Q)-poly(P) peptides showed PPII-like content but SEC showed poorly defined, overlapping oligomeric peaks, and as judged by CD these peptides retained significant PPII-like structure with increasing poly(Q) length. More importantly, addition of the poly(P) domain increased the threshold for fibril formation to ≈ 15 glutamine residues. X-ray diffraction, electron microscopy, and film CD showed that, while poly(Q) peptides with ≥ 6 glutamine residues formed β-sheet-rich fibrils, only the longest poly(Q)-poly(P) peptide (15 glutamine residues) did so. From these and other observations, we propose that poly(Q) domains exist in a “tug-of-war” between two conformations, a PPII-like helix and a β-sheet, while the poly(P) domain is conformationally constrained into a proline type II helix (PPII). Addition of poly(P) to the C terminus of a poly(Q) domain induces a PPII-like structure, which opposes the aggregation-prone β-sheet. These structural observations may shed light on the threshold phenomenon of poly(Q) aggregation, and support the hypothesized evolution of “protective” poly(P) tracts adjacent to poly(Q) aggregation domains.  相似文献   

20.
The rapid penetration of poly(A)-poly(U) into cell nuclei is shown by radioautography, by recovery of acid-precipitable material from isolated nuclei and by sucrose gradient centrifugation of nuclear lysates. The majority of poly(A)-poly(U) remains intact in the nuclei for at least h. This penetration is increased 20-fold by pretreatment of the cells with DEAE Dextran. In cells treated with DEAE Dextran, DNA and RNA syntheses are stimulated by poly(A)-poly(U) from the time the polymer complex is added and for at least h.  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