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 187 毫秒
1.
Double-stranded DNA fragments terminated at their 5′-ends by the singlestranded sequence pA-A-T-T-, generated by digestion of DNA with EcoRI restriction endonuclease, were ligated with Escherichia coli polynucleotide ligase under various conditions of temperature, concentration and time. The linear and circular products of ligation were separated by electrophoresis in agarose gel and quantitated by densitometry. The rate of ligation of (EcoRI-cleaved) simian virus (SV40) DNA at a concentration of 100 μg/ml increased from 0 °C to 5 °C to 10 °C (6-fold increase overall); raising the temperature to 15 °C did not further increase the rate of ligation. At the appropriate DNA concentrations, the predominant products of ligation are either linear concatemers that are integral multimers of the starting DNA fragment, or covalently closed circular structures of the monomeric DNA fragment. Ligating a mixture of two different length DNA fragments gives rise to all of the possible expected recombinant molecules.Linear or circular products of ligation were predicted by consideration of the total concentration of DNA termini, i, and the local concentration of one terminus in the neighborhood of the other on the same DNA molecule, j. The parameter j is a function of the length of a DNA molecule, providing this length is greater than the random coil segment of DNA. Experimentally it was found that circular structures are formed in significant amounts only under conditions when the value of j is several times greater than that of i. When j = i, equal amounts of linear and circular products would be expected, but most of the molecules were ligated into linear concatemers. No circular structure of a DNA fragment whose contour length l (6 × 10−2 μm) is smaller than the random coil segment value b (7·17 × 10−2 μm) was observed, while circular structures of the dimer of the same molecule (12 × 10−2 μm) were detected.  相似文献   

2.
We have investigated the binding of the biologically important flavonoid fisetin with the carrier protein bovine serum albumin using multi-spectroscopic and molecular docking methods. The binding constants were found to be in the order of 104 M?1 and the number of binding sites was determined as one. MALDI-TOF analyses showed that one fisetin molecule binds to a single bovine serum albumin (BSA) molecule which is also supported by fluorescence quenching studies. The negative Gibbs free energy change (?G°) values point to a spontaneous binding process which occurs through the presence of electrostatic forces with hydrophobic association that results in a positive entropy change (+51.69 ± 1.18 J mol?1 K?1). The unfolding and refolding of BSA in urea have been studied in absence and presence of fisetin using steady-state fluorescence and lifetime measurements. Urea denaturation studies indicate that fisetin is gradually released from its binding site on the protein. In the absence of urea, an increase in temperature that causes denaturation of the protein results in the release of fisetin from its bound state indicating that fisetin binds only to the native state of the protein. The circular dichroism (CD) and Fourier transform infrared (FTIR) spectroscopic studies showed an increase in % α-helix content of BSA after binding with fisetin. Site marker displacement studies in accordance with the molecular docking results suggested that fisetin binds in close proximity of the hydrophobic cavity in site 1 (subdomain IIA) of the protein. The PEARLS (Program of Energetic Analysis of Receptor Ligand System) has been used to estimate the interaction energy of fisetin with BSA and the results are in good correlation with the experimental findings.  相似文献   

3.
The interaction between aniline and ferriprotoporphyrin IX in alkaline solution has been investigated using pyridine and the N-methyl pyridinium ion as site-specific inhibitors of oxygen activation. Pyridine inhibits oxygen activation in a noncompetitive manner with respect to aniline (K1 = 1.24 mol −1 dm3 at 30°C) while the N-methyl pyridinium ion inhibited in a manner consistent with two sites for aniline binding, only one of which was competitively inhibited (K1 = 317mol-l dm3 at 30°C). A comprehensive reinvestigation of the interaction of pyridine and N-methyl pyridinium ions with alkaline ferriprotoporphyrin IX has shown that two molecules of each ligand bind per hemin dimer in a strongly cooperative manner. The association constant for the first pyridinium ion bound is K1a = 176 mol−1 dm3 at 30°C, while that for the first pyridine molecule bound is K1a = 0.580 mol−1 dm3 at 30°C; these are both close to the observed inhibition association constants (K1). Thermodynamic parameters for the interactions have been evaluated and compared to previous literature values. On the basis of these results a model is proposed for aniline interaction with the ferriprotoporphyrin dimer IX which involves the binding of two molecules of aniline to the ferriprotoporphyrin IX tetrapyrrole ring system by planar π bonding interactions with the rings having the propionate groups attached.  相似文献   

4.
P E Grebow  T M Hooker 《Biopolymers》1974,13(11):2349-2366
Semiempirical conformational energy calculations were carried out for the cyclic dipeptide L -alanyl-L -histidine diketopiperazine. The results indicate that electrostatic effects are probably significant in determining the conformation assumed by this molecule. When the imidazole group is in its uncharged state the most stable conformations of the molecule are those with the imidazole ring folded over the diketopiperazine ring (χ1 = 60°). Upon protonation of the imidazole group the folded conformation may be destabilized relative to conformations characterized by χ1 positions near 180°.  相似文献   

5.
A number of copper(II) complexes of tridentate ligands with various donor atoms have been studied in an attempt to duplicate the unusual reactivity patterns and accompanying spectral changes of the copper(II) center in galactose oxidase. Results indicate that in order to match the optical and electron spin resonance spectral change observed upon CN? binding by the enzyme, an equatorial, negative ligand must be displaced in a small molecule model. The crystal and molecular structure of the best model complex was solved by a single crystal X-ray diffraction study. The compound, monoacetato-1,3-bis(2-(4-methyl-pyridyl)imino)isoindolatocopper(II), crystallizes in the centro-symmetric triclinic space group Pī with a = 7.392(3) Å, b = 13.782(5) Å, c = 23.422(12) Å, α = 92.08(3)°, β = 104.11(5)°, γ = 109.98(4)°, V = 2156(1) Å3, d(obsd.)(calc.)=(1.43)(1.44) g/cm?3 for mol wt of 466.7 and Z = 4. Diffraction data were collected with a Syntex Pl diffractometer using graphite-monochromatized Cu radiation (λ = 1.5418 Å). The copper atoms were located from a Patterson synthesis; all other nonhydrogen atoms were located via difference. Fourier techniques, and hydrogen atoms were placed in calculated positions. Final refinement resulted in discrepancy indices of R = 0.089 and “Goodness to Fit” = 3.68 for all 3608 reflections having (I) ? 3σ(I) (5°<2θ<100°). There are two unique molecules in the asymmetric unit that are monomeric and well separated. The geometry around the copper atom is approximately square pyramidal, with the coordination sphere derived from three nitrogens of the tridentate ligand, one oxygen from the acetate unit, and an oxygen atom of a water molecule occupying an axial position. The structure is surprising both in that an axial water molecule is present and that the remaining four ligand atoms to the copper atom are rather distorted from a planar configuration. The plane defined by the copper, N5, and N3 atoms intersects the plane defined by the copper, Nl, and Ol, atoms forming a “twist angle” of 25.0° (0.0° would be ideal for a planar inner coordination sphere). The stereoelectronics of the inner coordination spheres of the type II Cu(II) enzymes galactose oxidase and superoxide dismutase are discussed and appropriate comparisons are made with emphasis on the origin of spectral changes observed upon anion binding.  相似文献   

6.
The molecular dynamics of the rhodopsin chromophore (11-cis-retinal) has been followed over a 3-ns path, whereby 3 × 106 discrete conformational states of the molecule were recorded. It is shown that within a short time, 0.3–0.4 ns from the start of simulation, the retinal β-ionone ring rotates about the C6–C7 bond through ~60° relative to the initial configuration, and the whole chromophore becomes twisted. The results of ab initio quantum chemical calculations indicate that for the final conformation of the chromophore center (t = 3 ns) the rhodopsin absorption maximum is shifted by 10 nm toward longer wavelengths as compared with the initial state (t = 0). In other words, the energy of transition of such a system into the excited singlet state S1 upon photon capture will be lower than that for the molecule where the β-ionone ring of the chromophore is coplanar to its polyene chain.  相似文献   

7.
Electron transfer between horse heart and Candida krusei cytochromes c in the free and phosvitin-bound states was examined by difference spectrum and stopped-flow methods. The difference spectra in the wavelength range of 540–560 nm demonstrated that electrons are exchangeable between the cytochromes c of the two species. The equilibrium constants of the electron transfer reaction for the free and phosvitin-bound forms, estimated from these difference spectra, were close to unity at 20°C in 20 mM Tris-HCl buffer (pH 7.4). The electron transfer rate for free cytochrome c was (2–3) · 104 M?1 · s?1 under the same conditions. The transfer rate for the bound form increased with increase in the binding ratio at ratios below half the maximum, and was almost constant at higher ratios up to the maximum. The maximum electron exchange rate was about 2 · 106 M?1 · s?1, which is 60–70 times that for the free form at a given concentration of cytochrome c. The activation energy of the reaction for the bound cytochrome c was equal to that for the free form, being about 10 kcal/mol. The dependence of the exchange rate on temperature, cytochrome c concentration and solvent viscosity suggests that enhancement of the electron transfer rate between cytochromes c on binding to phosvitin is due to increase in the collision frequency between cytochromes c concentrated on the phosvitin molecule.  相似文献   

8.
《BBA》2023,1864(4):149002
In cyanobacteria that undergo far red light photoacclimation (FaRLiP), chlorophyll (Chl) f is produced by the ChlF synthase enzyme, probably by photo-oxidation of Chl a. The enzyme forms homodimeric complexes and the primary amino acid sequence of ChlF shows a high degree of homology with the D1 subunit of photosystem II (PSII). However, few details of the photochemistry of ChlF are known. The results of a mutational analysis and optically detected magnetic resonance (ODMR) data from ChlF are presented. Both sets of data show that there are significant differences in the photochemistry of ChlF and PSII. Mutation of residues that would disrupt the donor side primary electron transfer pathway in PSII do not inhibit the production of Chl f, while alteration of the putative ChlZ, P680 and QA binding sites rendered ChlF non-functional. Together with previously published transient EPR and flash photolysis data, the ODMR data show that in untreated ChlF samples, the triplet state of P680 formed by intersystem crossing is the primary species generated by light excitation. This is in contrast to PSII, in which 3P680 is only formed by charge recombination when the quinone acceptors are removed or chemically reduced. The triplet states of a carotenoid (3Car) and a small amount of 3Chl f are also observed by ODMR. The polarization pattern of 3Car is consistent with its formation by triplet energy transfer from ChlZ if the carotenoid molecule is rotated by 15° about its long axis compared to the orientation in PSII. It is proposed that the singlet oxygen formed by the interaction between molecular oxygen and 3P680 might be involved in the oxidation of Chl a to Chl f.  相似文献   

9.
The proteins S6 and S18 from the 30 S ribosomal subunit of Escherichia coli were isolated to a purity of greater than 95%, characterized in solution, and investigated by sedimentation equilibrium for possible intermolecular interactions in a dilute salt reconstitution buffer. It was observed that neither protein S6 nor S18 has a tendency to self-associate in the concentration range studied. An analysis of solution mixtures containing proteins S6 and S18 revealed a species of molecular weight greater than either of the proteins. Proteins S6 and S18 were found to interact with an equilibrium constant of association of 6.6 ± 4.2 × 104m?1 at 3 °C with a Gibbs free energy of interaction, ΔG ° = ?6.1 kcal/mol. These data are part of those collected to help in building a map of the energetics in the 30 S ribosomal subunit, which provides for the stabilization of the structure.  相似文献   

10.
Two esteroproteolytic enzymes (A and D) have been isolated from the mouse submaxillary gland and shown to be pure by ultracentrifugation, immunoelectrophoresis, acrylamide-gel electrophoresis, and amino acid analyses. The enzymes have molecular weights of approximately 30,000 and are structurally and antigenically related. Narrow pH optima between 7.5 and 8.0 are exhibited by both enzymes. The “pK1's” are between 6.0 and 6.5 and the “pK2's” are near 9.0. A marked preference for arginine-containing esters is shown by both enzymes. The maximum specific activity of enzyme A on p-tosylarginine methyl ester (TAME) at pH 8 was 2500–3000 μm min?1 mg?1 and for enzyme D, 400–600 μm min?1 mg?1. With TAME as substrate, the Km for enzyme A was 8 × 10?4m at 25 °C and 6 × 10?4m at 37 °C. For D, Km was 3 × 10?4 at 25 °C and 2 × 10?4m at 37 °C.An apparent activation of enzyme D by tosylarginine (TA), a product of TAME hydrolysis, and all α-amino acids examined was due to removal of an inhibitor by chelation. This effect could be duplicated by 8-hydroxyquinoline and diethyldithiocarbamate but not by EDTA. Enzyme A was not affected by these substances to any remarkable extent. Several divalent ions proved to be potent inhibitors of enzyme D. Both enzymes are inactivated by the active site reagents diisopropyl phosphofluoridate and tosyllysine chloromethylketone but much less rapidly than is trypsin. Nitrophenyl-4-guanidionobenzoate reacts with a burst of nitrophenol liberation but with a rapid continuing hydrolysis. One active site per molecule is indicated. Enzyme D is inactivated by urea, reversibly at 10 m and with maximal permanent losses at 6 m. Autolysis of the unfolded form by the native enzyme when they coexist at intermediate urea concentrations appears to occur.Identity of enzyme D and the epithelial growth factor binding protein is demonstrated.  相似文献   

11.
Binding of [3H]aflatoxin B1 to rat plasma was investigated in vivo and vn vitro. Column chromatographic and polyacrylamide gel electrophoretic analysis clearly demonstrated that aflatoxin B1 bound primarily plasma albumin. Very little binding activity was shown by other plasma proteins. Spectrofluorimetric studies were undertaken to gain some insight into the nature of the aflatoxin-albumin interaction. Quenching of the lone tryptophan fluorescence intensity upon aflatoxin binding was due, at least in part, to a ligand-induced conformational change in the albumin molecule. Aflatoxin B1 binds an apolar site with an association constant of 30 mM−1 at pH 7.4 and 20°C. Neither charcoal treatment of rat albumin nor the presence of 0.15 M NaCl had many significant effect on the interaction. The association constant was pH-dependent, increasing about 1.7-fold as the pH increased from 6.1 to 8.4. This pH dependence is ascribed to a pH-induced conformational change in the albumin molecule. Thermodynamic studies indicated that the aflatoxin-albumin interaction was exothermic (ΔH = −29.3 kJ·mol), with a ΔS value of −13.8 J·mol−1·K−1.  相似文献   

12.
Unspecific binding of a protamine, namely fluorescein-labelled clupeine Z, to double-stranded calf thymus DNA was studied using fluorescence titration methods and chemical relaxation techniques. Both equilibrium and kinetic data have been analysed using general theoretical approaches discussed in the accompanying paper. The results agree well with the predictions made on the basis of a standard co-operative binding model.Basic parameters evaluated are the co-operative binding constant (K), the coefficient measuring co-operative interaction between nearest neighbours (q), the number of nucleotides occupied by one protamine molecule (n) and the rate constant of dissociation at the ends of bound ligand sequences (KD). Values obtained at 20 °C, pH 7.5 and 0.4 m-NaCl were K = 5.8 × 107m?1, q = 1700, n = 20 and KD = 0.29 s?1. They have been found to be sensitive to the concentration of added salt (NaCl). This effect apparently reflects the essentially electrostatic nature of the binding process. The results can be satisfactorily described in terms of competitive binding of sodium ions.  相似文献   

13.
The interaction of ochratoxin A, a mycotoxin produced by Aspergillus ochraceus, with isolated rat liver mitochondria and plasma membranes has been studied. Cell membranes bind [14C]ochratoxin A poorly and do not show saturation in the concentration range examined. The uptake of the toxin by mitochondria is saturable, with an apparent Km at 0 °C of 30 nmol/mg of protein. Sonication or freeze-thawing reduces the extent of incorporation by 88%. Ochratoxin A uptake is energy dependent, resulting in a depletion of intramitochondrial ATP. Uncouplers such as m-chlorocarbonylcyanide phenylhydrazone or the respiratory inhibitors rotenone and antimycin A inhibit uptake 60–85%, while ATP reverses the antimycin and rotenone inhibition. Phosphate transport is sensitive to inhibition by the toxin, as measured by Ca2+ plus Pistimulated respiration and [32P]Pi incorporation. In turn, phosphate inhibits nearly completely [14C]ochratoxin A uptake at 22 °C and causes a concomitant mitochondrial swelling yet is not incorporated into the matrix space. Thus, the saturable uptake of ochratoxin A is accompanied by a decrease in the energy state and inhibition of Pi transport, which results in deteriorative changes of the mitochondria, as evidenced by large-amplitude swelling.  相似文献   

14.
The interaction of a water-soluble dinuclear nickel(II) complex, [Ni2(EGTB)(CH3CN)4](ClO4)4·4H2O (EGTB = ethylene glycol-bis(β-aminoethyl ether) N,N,N′,N′-tetrakis(2-benzimidazoyl)) (1), and bovine serum albumin (BSA) was investigated under physiological conditions using fluorescence, synchronous fluorescence, UV–vis absorption and circular dichroism (CD). The experimental results suggested that the nickel(II) complex could bind to BSA with binding constant (K) ~ 104 M?1 and quench the intrinsic fluorescence of BSA through a static quenching mechanism. The thermodynamic parameters, ΔG°, ΔH°, and ΔS°, calculated at different temperatures, indicated that the binding reaction was spontaneous and electrostatic interactions played a major role in this association. Based on the number of binding sites, it was considered that one molecule of complex 1 could bind to a single site or two sites of the BSA molecule or the two binding modes coexisted. In view of the results of site marker competition experiments, the reactive sites of BSA to complex 1 mainly located in subdomain IIA (site I) and subdomain IIIA (site II) of BSA. Moreover, the binding distance, r, between donor (BSA) and acceptor (complex 1) was 5.13 nm according to Förster nonradiation energy transfer theory. Finally, as shown by the UV–vis absorption, synchronous fluorescence and CD, complex 1 could induce conformation and microenvironmental changes of BSA. The results obtained herein will be of biological significance in toxicology investigation and anticancer metallodrug design.  相似文献   

15.
RNA extracted by cold phenol from the large cytoplasmic ribosomal subunit of the trypanosomatid flagellate Crithidia fasciculata and analyzed by polyacrylamide gel electrophoresis at 4 °C consisted of one species with a molecular weight of 1.3 × 106 (relative to ribosomal RNA from E. coli MRE 600). When extracted with hot phenol (65 °C), the large ribosomal subunit gave rise to two components with molecular weights of 0.72 and 0.56 × 106. On heating for 60 s, followed by rapid cooling, the single cold-phenol-extracted 1.30 × 106-dalton species completely dissociated into two components of molecular weights 0.72 and 0.56 × 106, present in equimolar amounts. When analyzed by polyacrylamide-agarose gel electrophoresis in the presence of SDS, RNA extracted by cold phenol from the large cytoplasmic ribosomal subunit consisted of three components of molecular weights 1.3, 0.72, and 0.56 × 106, present in apparently equimolar amounts. RNA from the small cytoplasmic ribosomal subunit consisted of one species with a molecular weight of 0.84 × 106, independent of extraction or analytical conditions. It is proposed that under high salt and low temperature conditions, the large ribosomal RNA molecule is held together by its secondary structure, and that denaturing extraction or analytical conditions reveal an otherwise “hidden” lesion present in the molecule in vivo.  相似文献   

16.
The kinetics of the hydrogen-deuterium exchange reactions of double-helical poly (rI) · poly (rC), single-stranded poly(rC) and poly(rI), inosine, and cytosine- 5′-phosphoric acid have been examined, at various temperatures in the range 20 °C to 52 °C, by stopped-flow ultraviolet spectrophotometry, in the region 270 to 300 nm. For the solution of double-helical poly(rI) · poly(rC), two first-order deuteration reactions were found: a fast one and a slow one. At 25 °C and at pH 7.0, the rate constant was 12.3 s?1 for the fast reaction, and 0.13 s?1 for the slow reaction. The rate constant of the fast reaction is nearly equal to that of the single-stranded poly(rC) (12.6 s?1), and is assigned to the deuteration at the amino hydrogen (that is, free from the C · I hydrogen bond) of the cytosine residue. The slow reaction is attributable to the deuteration of the two hydrogens: the amino hydrogen of rC and imide hydrogen of rI, which are rapidly exchanging with each other within every rC · rI base-pair. From the observed temperature effect on this slow reaction rate, it has been concluded that there are two types of “opening process” that are relevant to the hydrogen exchange reaction; one of them is predominent in the range 47 °C to 52 °C and the other in the temperature region lower than 47 °C. The enthalpy (H) and entropy (S) differences of the “open” and “closed” forms in the former type process are ΔH = 167 kcal per mole and ΔS = 507 e.u., while in the latter ΔH = 8.1 kcal per mole and ΔS = 10 e.u..  相似文献   

17.
The kinetics of binding of 1-naphthylacetic acid to particulate fractions from tobacco-pith callus were studied. This binding site does not bind auxin at 0° C. Binding experiments performed at 25° C demonstrated an apparent K a of approx. 6.5·106 M-1. A filtration method was developed in order to study non-equilibrium kinetics of this binding. Dissociation of the complex of auxin and binding site indicates the presence of at least two binding components with dissociation rate constants (k off) of 6.1·10-3 min-1 and 6.0·10-2 min-1. This binding behaviour was not independent, indicating that the binding of auxin to the particulate fractions was more complex than binding of one hormone molecule to one binding site. This complexity was further confirmed by experiments in which the initial velocity of complex formation was measured. A model was worked out into which our data fit without contradictions. It involves the binding of four hormone molecules to one receptor molecule.  相似文献   

18.
When yeast tRNAPfPhe, a derivative of tRNAPhe in which proflavine replaces the Y base, is bound simultaneously to both the peptidyl and aminoacyl sites of a 70 S Escherichia coli ribosome, there is a rapid mutual energy transfer between the two bound tRNAs. Analysis of this energy transfer yields an upper limit for the proflavine-proflavine distance of 20 Å. It also allows an unequivocal measurement of the emission spectrum of tRNAPfPhe bound at the aminoacyl site. In the presence of message this spectrum is very different from that seen in the peptidyl site, implying that in the two sites the hypermodified bases exist in significantly different environments. The rapid energy transfer leads to some loss of fluorescence anisotropy. This can be analyzed to obtain an estimate of the angle between the two proflavines: 28 ° ± 10 ° or 152 ° ± 10 °. Taken together all of these results place severe constraints on possible models of codon-anticodon complexes. The mutual energy transfer seen and analyzed on the ribosome is a convenient aspect of fluorescence spectroscopy, and it is one that should see broad application where multiple copies of a fluorescent ligand interact on a macromolecular substrate.  相似文献   

19.
Ab initio RHF/4–31G molecular-orbital calculations have been conducted on methoxymethyl formate and methoxymethyl acetate as models for examining the anomeric effect and stereochemistry of 1-O-acetylglycopyranoses. The results indicate that, as with the methyl glycopyranosides, the α-4C1(D) configurations are more stable than the β-4C1(D), except that the energy difference is more dependent on the disposition about the glycosidic bond. The lowest-energy conformations occur with glycosidic torsion-angles of ?  180°, where the anomeric energy is about 4 kcal/mol. There is a secondary energy-minimum at ?  90°, for which the anomeric energy is less, about 2 kcal/mol. This orientation corresponds to the conformation most commonly observed in the crystal structures of peracetylated glycopyranoses. Small differences in the CO single-bond lengths, which are observed experimentally in both the α and β anomers, are reproduced by the theoretical calculations.  相似文献   

20.
Following dialysis against distilled water, the 16 S ribosomal RNA of Escherichia coli is unable to interact with 30 S subunit protein S4 at 0 °C. The dialysed RNA recovered this capacity, however, when heated at 40 °C in the presence of 0.02m-MgCl2 prior to addition of the protein. Furthermore, its sensitivity to ribo-nuclease markedly declined and its sedimentation rate increased as a consequence of this treatment. Although no concomitant changes in secondary structure were detected by absorbance and fluorescence techniques, the rearrangement of a small number of base-pairs was not excluded. Kinetic measurements revealed that binding site reactivation satisfies the first-order rate law and that the process is highly temperature-dependent, exhibiting an Arrhenius activation energy of 40,800 cal/mol. Together, these data suggest that dialysed RNA undergoes a unimolecular conformational transition upon pre-incubation in Mg2+-containing buffers and that this transition leads to renaturation of the binding site for protein S4.Similar results were obtained for several other proteins of the 30 S subunit. In particular, S7, S16/S17 and S20 all failed to interact efficiently with dialysed 16 S RNA at 0 °C. These proteins bound normally to the RNA, however, after it had been incubated at 40 °C in the presence of Mg2+ ions. By contrast, prior dialysis of the 16 S RNA did not affect its ability to associate with S8 and S15 at 0 °C. These two proteins interacted equally well with dialysed and pre-incubated 16 S RNA, indicating that their binding sites are not susceptible to the reversible alterations in conformation which influence the attachment of the other RNA-binding proteins to the nucleic acid molecule. The effects of dialysis and pre-incubation on the interaction of 16 S RNA with an unfractionated mixture of 30 S subunit proteins were also investigated. The dialysed RNA bound only S6, S8, S15 and S18 at 0 °C whereas, after heating at. high Mg2+ concentrations, the RNA associated with S4, S7, S9, S13, S16/S17, S19 and S20 as well. These results leave little doubt that the protein-binding capacities of the 16 S RNA are intimately related to its three-dimensional configuration, although individual binding sites appear to differ significantly in their stability to small changes in structure.  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