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 31 毫秒
1.
Escherichia coli ribosomal protein S1 plays a central role in initiation of protein synthesis, perhaps via participation in the binding of messenger RNA to the ribosome. S1 protein has two nucleic acid binding sites with very different properties: site I binds either single-stranded DNA or RNA, while site II binds single-stranded RNA only (Draper et al., 1977). The nucleic acid binding properties of these sites have been explored using the quenching of intrinsic protein fluorescence which results from binding of oligo- and polynucleotides, and are reported in this and the accompanying paper (Draper &; von Hippel, 1978).Site I has been studied primarily using DNA oligomers and polymers, and has been found to have the following properties. (1) The intrinsic binding constant (K) of site I for poly(dA) and poly(dC) is ~3 × 106m?1 at 0.12 m-Na+, and the site size (n, the number of nucleotide residues covered per S1 bound) is 5.1 ± 1.0 residues. (2) Binding of site I to polynucleotides is non-co-operative. (3) The K value for binding of S1 to single-stranded polynucleotides is ~103 larger than K for binding to double-stranded polynucleotides, meaning that S1 (via site I) is a potential “melting” or “double-helix destabilizing” protein. (4) The dependence of log K on log [Na+] is linear, and analysis of the data according to Record et al. (1976) shows that two basic residues in site I form charge-charge interactions with two DNA phosphates. In addition, a major part of the binding free energy of site I with the nucleic acid chain appears to involve non-electrostatic interactions. (5) Oligonucleotides bound in site II somewhat weaken the binding affinity of site I. (6) Binding affin is virtually independent of base and sugar composition of the nucleic acid ligand; in fact, the total absence of the base appears to have little effect on the binding, since the association constant for 2′-deoxyribose 5′-phosphate is approximately the same as that for dAMP or dCMP. (7) Two molecules of d(ApA) can bind to site I, suggesting the presence of two “subsites” within site I. (8) Iodide quenching experiments with S1-oligonucleotide complexes show differential exposure of tryptophans in and near the subsites of site I, depending upon whether neither, one, or both subsites are complexed with an oligonucleotide.  相似文献   

2.
In this paper we examine the specificity of the co-operative binding (in the polynucleotide mode) of bacteriophage T4-coded gene 32 protein to synthetic and natural single-stranded nucleic acids differing in base composition and sugar type. It is shown by competition experiments in a tight-binding (low salt) environment that there is a high degree of binding specificity under these (protein-limiting) conditions, with one type of nucleic acid lattice binding gene 32 protein to saturation before any binding to the competing lattice takes place; it is also shown that the same differential specificities apply at high salt concentrations. Procedures developed in the preceding paper (Kowalczykowski et al., 1980) are used to measure the net binding affinities () of gene 32 protein to a variety of polynucleotides, as well as to determine individual values of K and ω for some systems. For all polynucleotides, virtually the entire specificity and salt dependence of binding of appears to be in K. In ~0.2 m-NaCl, the net binding affinities () range from ~106 to ~1011m?1; in order of increasing affinities we find: poly(rC) < poly(rU) < poly(rA) < poly(dA) < poly(dC) < poly(dU) < poly(rI) < poly(dI) < poly-(dT). In general, for a particular homopolyribonucleotide at constant salt concentration is 101 to 104smaller than for the corresponding homopoly-deoxyribopolynucleotide. Values of for randomly copolymerized polynucleotides and for natural DNA fall at the compositionally weighted average of the values for the individual homopolynucleotides (except for poly(dT), which appears to bind somewhat tighter), indicating that the net affinity represents the sum of the binding free energy contributions of the individual nucleotides. It is shown that these results, on a competition basis under physiological salt conditions, can account quantitatively for the autogenous regulation of the synthesis of gene 32 protein at the translational level (Russel et al., 1976; Lemaire et al., 1978). In addition, these results suggest possible mechanisms by which gene 32 messenger RNA might be specifically recognized (by gene 32 protein) and functionally discriminated from the other mRNAs of phage T4.  相似文献   

3.
The DeVoe polarizability theory is used to calculate vibrational circular dichroism (VCD) and infrared (IR) absorption spectra of four polyribonucleotides: poly(rA) x poly(rU), poly(rU) x poly(rA) x poly(rU), poly(rG) x poly(rC), and poly(rC+) x poly(rI) x poly(rC). This is the first report on the use of the DeVoe theory to calculate VCD, oriented VCD, IR absorption, and IR linear dichroism (LD) spectra of double- and triple-stranded polyribonucleotides. Results are reported for DeVoe theory calculations--within the base-stretching 1750-1550 cm(-1) spectral region--on several proposed multistranded polyribonucleotide geometries. The calculated spectra obtained from these proposed geometries are compared with previously reported measured and calculated VCD and IR spectral results. Base-base hydrogen-bonding effects on the frequencies and magnitudes of the base carbonyl stretching modes are explicitly considered. The good agreements found between calculated and measured spectra are proposed to be further evidence of the usefulness of the DeVoe theory in drawing three-dimensional structural conclusions from measured polyribonucleotide VCD and IR spectra.  相似文献   

4.
In this paper we examine molecular details of the interaction of bacteriophage T4-coded gene 32 protein with oligo- and polynucleotides. It is shown that the binding affinity (Koligo) of oligonucleotides of length (l) from two to eight nucleotide residues for gene 32 protein is essentially independent of base composition or sugar type. This binding also shows little dependence on salt concentration and on oligonucleotide length; even the expected statistical length factor in Koligo is not observed, suggesting that binding occurs at the end of the oligonucleotide lattice and that the oligonucleotide is not free to move across the binding site. Co-operative (contiguous) or isolated binding of gene 32 protein to polynucleotides is very different; here binding is highly salt dependent (? log Kω? log [NaCl] ~- ?7) and essentially stoichiometric at salt concentrations less than ~0.2 m (for poly(rA)). Binding becomes much weaker and the binding isotherms appear typically co-operative (sigmoid) in protein concentration at higher salt concentrations. We demonstrate, by fitting the co-operative binding isotherms to theoretical plots at various salt concentrations and also by measuring binding at very low protein binding density (ν), that the entire salt dependence of is in the intrinsic binding constant (K); the co-operativity parameter (ω) is essentially independent of salt concentration. Furthermore, by determining titration curves in the presence of salts containing a series of different anions and cations, it is shown that the major part of the salt dependence of the gene 32 protein-polynucleotide interaction is due to anion (rather than to cation) displacement effects. Binding parameters of oligonucleotides of length sufficient to bind two or more gene 32 protein monomers show behavior intermediate between the oligonucleotide and the polynucleotide binding modes. These different binding modes probably reflect different conformations of the protein; the results are analyzed to produce a preliminary molecular model of the interactions of gene 32 protein with nucleic acids in its different binding modes.  相似文献   

5.
S1 is an acidic protein associated with the 3′ end of 16S RNA; it is indispensable for ribosomal binding of natural mRNA. We find that S1 unfolds single stranded stacked or helical polynucleotides (poly rA, poly rC, poly rU). It prevents the formation of poly (rA + rU) and poly (rI + rC) duplexes at 10–25 mM NaCl but not at 50–100 mM NaCl. Partial, salt reversible denaturation is also seen with coliphage MS2 RNA, E. coli rRNA and tRNA. Generally, only duplex structures with a Tm greater than about 55° are formed in the presence of S1. The protein unfolds single stranded DNA but not poly d(A·T).  相似文献   

6.
We demonstrate that the isotropic absorption and linear dichroism in an unknown flow field can be used to determine base tilt in polynucleotides if three transitions are measured and the directions of the corresponding dipoles are known. The method is applied here to reach conclusions about the base tilt in poly(rA), poly(rA)+·poly(rA), and poly(rC). The respective values are: 28° tilt about the axis + 50° toward C8 from the C1′ → N9, and 25° tilt about the axis + 118° toward C5 from C1′ → N1. The results for poly(rA)+·poly(rA) are consistent with the accepted model. Spectra were measured for poly(rC)+·poly(rC), but definite conclusions must await reliable directions for transition dipoles. The dipole direction for the 218-nm transition in rC is found to be +13° or +43° toward C5 from C1′ → N1. The CD spectra to about 168 nm are presented and discussed.  相似文献   

7.
8.
Herein, we evaluate the binding of Pb(II) and Bi(III) to cysteine-substituted versions of the TRI peptides [AcG-(LKALEEK)4G-NH2] which have previously been shown to bind Hg(II) and Cd(II) in unusual geometries as compared with small-molecule thiol ligands in aqueous solutions. Studies of Pb(II) and Bi(III) with the peptides give rise to complexes consistent with the metal ions bound to three sulfur atoms with M–S distances of 2.63 and 2.54 Å, respectively. Competition experiments between the metal ions Pb(II), Cd(II), Hg(II) and Bi(III) for the peptides show that Hg(II) has the highest affinity, owing to the initial formation of the extremely strong HgS2 bond. Cd(II) and Pb(II) have comparable binding affinities at pH > 8, while Bi(III) displays the weakest affinity, following the model, M(II) + (TRI LXC)3 3? → M(II)(TRI LXC)3 ?. While the relevant equilibria for Hg(II) binding to the TRI peptides corresponds to a strong first step forming Hg(TRI LXC)2(HTRI LXC), followed by a single deprotonation to give Hg(TRI LXC)3 ?, the binding of Cd(II) and Pb(II) is consistent with initial formation of M(II)(TRI LXC)(HTRI LXC)2 + at pH < 5 followed by a two-proton dissociation step (pK a2) yielding M(II)(TRI LXC)3 ?. Pb(II)(TRI LXC)(HTRI LXC)2 + converts to Pb(II)(TRI LXC)3 ? at slightly lower pH values than the corresponding Cd(II)–peptide complexes. In addition, Pb(II) displays a lower pK a of binding to the “d”-substituted peptide, (TRI L12C, pK a2 = 12.0) compared with the “a”-substituted peptide, (TRI L16C, pK a2 = 12.6), the reverse of the order seen for Hg(II) and Cd(II). Pb(II) also showed a stronger binding affinity for TRI L12C (K bind = 3.2 × 107 M?1) compared with that with TRI L16C (K bind = 1.2 × 107 M?1) at pH > 8.  相似文献   

9.
In a search for eucaryotic enzymes which might process the heterogenous nuclear RNA (HnRNA) from animal cells into messenger RNA, a ribonuclease called RNAse D analogous to E. coli RNAse III in its ability to cleave specifically synthetic or viral double-stranded polyribonucleotides has been detected and extensively purified from the cytosol of Krebs II mouse ascites cells. The purification procedure involved cellular fractionation followed by DEAE-and CM-cellulose chromatography and resulted in an RNAas D preparation contaminated with trace amounts of single-strand specific RNAse (equivalent to less than 0.3 ng per ml) as assayed against poly (rC). Significant levels of RNAse H activity against poly (rA)-poly (dT) were still present in these preparations.  相似文献   

10.
Acetohexamide is a drug used to treat type II diabetes and is tightly bound to the protein human serum albumin (HSA) in the circulation. It has been proposed that the binding of some drugs with HSA can be affected by the non-enzymatic glycation of this protein. This study used high-performance affinity chromatography to examine the changes in acetohexamide–HSA binding that take place as the glycation of HSA is increased. It was found in frontal analysis experiments that the binding of acetohexamide to glycated HSA could be described by a two-site model involving both strong and weak affinity interactions. The average association equilibrium constant (Ka) for the high affinity interactions was in the range of 1.2–2.0 × 105 M−1 and increased in moving from normal HSA to HSA with glycation levels that might be found in advanced diabetes. It was found through competition studies that acetohexamide was binding at both Sudlow sites I and II on the glycated HSA. The Ka for acetohexamide at Sudlow site I increased by 40% in going from normal HSA to minimally glycated HSA but then decreased back to near-normal values in going to more highly glycated HSA. At Sudlow site II, the Ka for acetohexamide first decreased by about 40% and then increased in going from normal HSA to minimally glycated HSA and more highly glycated HSA. This information demonstrates the importance of conducting both frontal analysis and site-specific binding studies in examining the effects of glycation on the interactions of a drug with HSA.  相似文献   

11.
Brief treatment of gene 32 protein with proteolytic enzymes produces two specific digestion products in good yield (Moise & Hosoda, 1976). One, representing the native protein with ~60 amino acid residues removed from the C-terminus, is G32P1I. The other, for which ~20 amino acid residues have been removed from the N-terminus in addition to the 60 residues from the C-terminus, is G32P1III. Both of these specific “core” fragments of gene 32 protein have been isolated and purified, and their binding properties to single-stranded oligo- and polynucleotides have been studied. We find that the binding properties of G32P1I are relatively little changed from those characteristic of the native gene 32 protein: (1) the apparent binding constants to short (l = 2 to 8) oligonucleotides are independent of lattice length and essentially independent of base and sugar composition, but do show an increased salt dependence of binding relative to that of the native protein; (2) the intrinsic association constants (K) for polynucleotides binding in the co-operative mode show the same binding specificities as seen with the native protein, but with absolute values increased two to fourfold; (3) the polynucleotide binding co-operativity parameter (ω?2 × 103) and the binding site size (n ~-7 nucleotide residues) are the same as for the native protein; (4) essentially the entire salt dependence of the net affinity () remains in K. However, unlike native gene 32 protein, G32P1I can melt native DNA to equilibrium (Hosoda et al., 1974; Greve et al., 1978); this suggests that the kinetic pathways for DNA melting by these two species must differ, since the changes in equilibrium binding parameters measured here are far too small to account for the differences in melting behavior. In contrast to G32P1I, for G32P1III we find that: (1) binding is non-cooperative (ω ~-1); (2) the binding site size (n) for the protein has decreased by one to two nucleotide residues relative to that characteristic of the native protein and G32P1I; (3) binding to short (l = 2 to 8) oligonucleotides is length and salt concentration dependent; (4) while binding to polynucleotides continues to show approximately the same base composition dependence as the native protein, the absolute values of K are somewhat different and the salt concentration dependencies of K are less. Polynucleotide ultraviolet light and circular dichroism spectra obtained in the presence of G32P1I and G32P1III are indistinguishable from those measured with the native protein at similar binding densities, indicating that all three protein species distort the polynucleotide lattice to comparable extents.These results are combined with the equilibrium binding data for native gene 32 protein (Kowalczykowski et al., 1980a: Newport et al., 1980) to obtain further insight into the molecular details of the interactions of this protein with its nucleic acid binding substrates.  相似文献   

12.
Lipocortin I-like protein with a molecular weight of 94,000 Da as judged by Western analysis was found to bind to ssDNA rather than to dsDNA in a Ca(2+)-dependent manner. This protein was also bound to [(32)P]poly(rA) and [(32)P]poly(rG) as measured by EMSA. Poly(rG), poly(rA), poly(dC), and poly(dT) were competitive against binding of either [(32)P]poly(rA) or [(32)P]poly(rG), while poly(rC), poly(rU), and poly(dA) were less effective binding competitors. The binding of this protein to poly(rA) or poly(rG) was inhibited by immunoprecipitable anti-lipocortin I (calpactin II) and anti-S100 protein antibodies, but not by an anti-Ig antibody. Phospholipids such as phosphatidylserine and phosphatidylinositol enhanced the binding of lipocortin I to poly(rA). Taken together, our present observations suggest that the lipocortin I-S100 protein heterotetramer binds to either purine RNAs or pyrimidine ssDNAs in a Ca(2+)- and phospholipid-dependent manner.  相似文献   

13.
The synthesis and interactions of the d- and l-enantiomers of the amino acid amide derivatives [Formula: see text] (I) and lysyl dipeptides [Formula: see text] (II) with poly rI.poly rC, poly rA.poly rU and calf thymus DNA is reported. The following results were found. (1) The degree of stabilization of the helices as measured by the T(m) (;melting' temperature) of the helix-coil transition was dependent on the nature of the amino acid. (2) For the poly rI.poly rC helix, the l-enantiomers of salts (I) and (II) stabilized more than the d-enantiomers. The same was true for calf thymus DNA in the presence of salts (II) and for poly rA.poly rU in the presence of salts (II) and the proline derivatives of salts (I). (3) As R increased in size and became more apolar, the amount of stabilization of the poly rI.poly rC helix in the presence of salts (I) decreased. On the other hand, the amount of stabilization increased with more polar substituents. An attempt was then made to determine whether the difference in stabilization of the double-stranded helices at the T(m) in the presence of the l- and d-enantiomers of salts (I) is due to the interaction with the helix, the random coil or both. A new method was developed for determining the binding of small ions to polyions that involves a competition between an insoluble polystyrene ion-exchange resin and the soluble polyion for the counterion. Dissociation constants are obtained for the complexes of single- and double-stranded helices with the salts (I). The results are illuminating and indicate that with certain helices, i.e. poly rA.poly rU, the interactions of salts (I) with the single strands may not be ignored. It is concluded that the high optical specificity found in Nature, i.e. d-ribose in nucleic acids and l-amino acids in proteins, cannot be attributed solely to monomer-polymer interactions described by Gabbay (1968).  相似文献   

14.
Gel filtration studies demonstrate that retinol receptors of chick retinal and pigment epithelial cytosols are (1) of very similar nature (2) of small molecular size (about 18 000 daltons) and are different in character from serum proteins. Citral inhibits the binding of [3H] retinol to the retinal 2 S receptor. Retinol acetate competes with retinol for binding to 2 S receptor in both retina and pigment epithelium whereas retinol palmitate is an effective competitor only in the pigment epithelium. Dithiothreitol maximizes 2 S binding in retina and pigment epithelial cytosol; its absence does not lead to receptor aggregation however. A limited number of high affinity binding sites (2 S receptor) appear to be present in retina and pigment epithelium. A 5 S binding species is also present in pigment epithelium; it is similar in character to [3H] retinol binding in serum and may arise from serum contamination of the pigment epithelial preparation. Binding affinity in retina is high with possibly two classes of retinol binding sites present of KD about 1·10?9 and 4·10?8.  相似文献   

15.
16.
Uno T  Aoki K  Shikimi T  Hiranuma Y  Tomisugi Y  Ishikawa Y 《Biochemistry》2002,41(43):13059-13066
The binding of the copper(II) complex of water-soluble meso-tetrakis(N-methylpyridinium-4-yl)porphyrin (TMPyP) to double-helical polynucleotides has been studied by optical absorption, circular dichroism (CD), and resonance Raman spectroscopic methods. The target polymers were RNA and RNA.DNA hybrids consisting of rA.rU, rI.rC, rA.dT, and rI.dC base pairs. Relative to the metal-free H(2)TMPyP [Uno, T., Hamasaki, K., Tanigawa, M., and Shimabayashi, S. (1997) Inorg. Chem. 36, 1676-1683], CuTMPyP binds to poly(rA).poly(dT) and poly(rA).poly(rU) with a greatly increased binding constant. The external self-stacking of the porphyrin on the surface of the polymers was evident from the strong conservative-type induced CD signals. The signal intensity correlated almost linearly with the number of stacking sites on the polymer except for poly(rA).poly(dT), which showed extraordinarily strong CD signals. Thus, the bound porphyrin may impose an ordered architecture on the polymer surface, the stacking being facilitated by the more planar nature of the CuTMPyP than the nonmetal counterpart. Resonance Raman spectra of the stacked CuTMPyP were indistinguishable from those of the intercalated one with positive delta(Cbeta-H) and negative delta(Cm-Py) bending shifts, and hence the stacked porphyrins are suggested to adopt a similar structure to that of intercalated ones. Porphyrin flattening by copper insertion opens a new avenue for medical applications of porphyrins, blocking biological events related to RNA and hybrids in malignant cells.  相似文献   

17.
E J Gabbay 《Biopolymers》1967,5(8):727-747
Information concerning the structures of rA–rU, rA–rU2 rI–rC, rA–rI2, and acid rA helices in solutions is reported. Through the use of diquaternary ammonium salts of the general structure, \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm R}_1 {\rm R}_2 {\rm R}_3 \mathop {\rm N}\limits^ + ({\rm CH}_2 )n\mathop {\rm N}\limits^ + {\rm R}_1 {\rm R}_2 {\rm R}_3 \cdot 2{\rm Br}^ - $\end{document} (I), it is shown that (1) the distances between adjacent negatively charged oxygen atoms on the helix increases in the following order rA–rI2 < rI–rC < rA–rU ? rA–rU2; (2) the density of the helices increases in the order. rA–rI2 < rA–rU < rA–rU2 < rI–rC; (3) there is a large hydrophobia site in rA–rI2 and possibly also in rA–rU, rA–rU2, and rI–rC helices; (4) the results of the interactions between the salts of type I and the helices may be formulated in semi-quantitative terms by the use of two parameters, α, and β which are shown to be related to the charge separation and the density of the helices, respectively; (5) the studies in solutions compare favorably with the x-ray studies on the fibers; and (6) the acid rA helix differs significantly from the other helices by the fact that the electrostatic interstrand interactions between the negatively charged oxygen atom of a phosphate group and the positively charged 10-amino group of adenine contribute significantly to the stabilization of the helix, and thus it is found that the presence of the salts, I, leads to a significant destabilization of the acid rA helix.  相似文献   

18.
Interaction of sanguinarine with A-form RNA structures of poly(rI)poly(rC) and poly(rA).poly(rU) has been studied by spectrophotometric, spectrofluorimetric, UV melting profiles, circular dichroism and viscometric analysis. The binding of sanguinarine to A-form duplex RNA structures is characterised by the typical bathochromic and hypochromic effects in the absorption spectrum, increasing steady state fluorescence intensity, an increase in fluorescence quantum yield of sanguinarine, an increase in fluorescence polarization anisotropy, an increase of thermal transition temperature, an increase in the contour length of sonicated rod-like RNA structure and perturbation in circular dichroic spectrum. Scatchard analysis indicates that sanguinarine binds to each polymer in a non-cooperative manner. Comparative binding parameters determined from absorbance titration by Scatchard analysis, employing the excluded site model, indicate a higher binding affinity of sanguinarine to poly(rI).poly(rC) structure than to poly(rA).poly(rU) structure. On the basis of these observations, it is concluded that the alkaloid binds to both the RNA structures by a mechanism of intercalation.  相似文献   

19.
20.
A cyclic AMP-adenosine binding protein, whose binding sites are activated by preincubation in the presence of Mg+-ATP, has been purified to apparent homogeneity from mouse liver (P.M. Ueland and S.O. Døskeland, 1977, J. Biol. Chem.,252, 677–686). The degree of activation of both the cyclic AMP binding site and a high-affinity site for adenosine depends on the concentration of ATP during the preincubation. The velocity and the degree of activation are dependent on the temperature and the presence of Mg2+ and K+. The NH4+ ion can be substituted for K+, whereas Na+ is inefficient. Low pH promotes the conversion from the inactive to the active form. The apparent affinity for adenosine to the high-affinity site for this adenine derivative and the affinity for cyclic AMP to the site specific for this nucleotide are independent of the degree of activation as judged from the slope of Scatchard plots. The activation of the cyclic AMP binding site by ATP (6 mm) was determined at pH 7 in the presence of 10 μm cyclic AMP, AMP, ADP, or adenosine. Adenosine specifically inhibits the activation and does not promote the inactivation of the binding protein. The possibility that the apparent inhibition of activation was effected by interference with cyclic AMP binding by adenosine was ruled out.  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