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 93 毫秒
1.
H Singh  B Hillier 《Biopolymers》1971,10(12):2445-2457
The optical rotatory dispersion (ORD) of several diribonucleoside monophosphates (NpN) and the corresponding 2′-O-methyl substituted dinucleoside monophosphates containing 2′-O-methyl ribosyl 3′-nucleotide and a 5′-nucleoside (NmpN) were measured at pH 1, 7, and 11.2, at 0.1 ionic strength in order to examine the role of the 2′-hydroxyl group of the ribose in the conformation of the oligoribonucleotides. The optical measurements are reported from 210 to 340 mμ. The pH effect on the ORD spectra of NpN as well as NmpN are large. No dramatic changes are seen in the shapes of the ORD spectra of the NmpN to the corresponding NpN at pH 7. However, a decrease in the amplitude is seen in most of the NmpN over that of the corresponding NpN ranging from 7 percent in the case of UmpG to 46 percent in AmpA. The differences seen in the NpN and the corresponding NmpN ORD results are best explained as a consequence of a change in the ribosyl conformation on 2′-O-methylation, rather than the involvement of the 2′-hydroxyl group in intramolecular hydrogen bonding in the ribo dimmer. The NmpN behave like NpN and not dNpdN, suggesting that the geometry of the stack in NpmN and NpN depends on the oxygen at the 2′-carbon and not on what is attached to it.  相似文献   

2.
Abstract

The first-order rate constants for hydrolysis of 3′-C-methyluridylyl(2′,5′)- and -(3′,5′)adenosine and the corresponding native dinucleoside monophosphates (2′,5′- and 3′,5′-UpA) have been determined as a function of hydroxide-ion concentration (0.025 - 7 M) at 25°C. In addition to the effects on the hydrolytic stability of the compounds, the effects of the 3′-C-methyl substitution on the kinetically determined pK a values for the sugar hydroxyls of the undine moiety are discussed.  相似文献   

3.
The conformational properties of four 2′ – 5′ dinucleoside monophosphates modified with N-2-acetylaminofluorene have been studied by circular dichroism spectroscopy. Covalent binding of this chemical carcinogen at the C8 position of guanosine in the 2′ – 5′ dinucleoside monophosphates induces striking changes in their circular dichroic spectra depending on their base sequence and composition. The changes in CD spectra, redshift of the extrema and change of their polarity, not observed in the spectra of corresponding 3′ – 5′ derivatives modified with N-2-acetylaminofluorene are correlated with the difference in the configuration of 2′ – 5′ and 3′ – 5′ dinucleoside monophosphates and discussed in respect to the intramolecular stacking interactions.  相似文献   

4.
M D Topal  M M Warshaw 《Biopolymers》1976,15(9):1755-1773
A least squares analysis of the titration properties of several dinucleoside monophosphates enables calculation of the pK's for protonation. These pK's are used to resolve the spectral properties of dinucleoside monophosphates with one base charged from the apparent spectral properties of a dinucleoside monophosphate in aqueous solution. This method is applied to dinucleoside monophosphates containing adenosine and/or cytidine. Results of CD, nmr, and CD-temperature dependence measurements are presented. The results indicate that singly protonated dimers of these nucleosides stack as do their unprotonated analogs. It is suggested that this is true for all dimers with one base charged.  相似文献   

5.
The turnover rates for hydrolysis by nuclease P1 of the 16 unmodified dideoxynucleoside monophosphates were measured. In addition, the turnover rates were measured in a variety of dideoxynucleoside monophosphates containing free radical-induced base modifications. The modified bases included cis-5,6-dihydroxy-5,6-dihydrothymine (thymine glycol), 5,6-dihydrothymine, 5-hydroxymethyuracil, 8-hydroxyguanine, 5-hydroxy-5-methylhydantoin and the formamido remnant which can be derived from either a thymine or a cytosine base. The turnover rate for dinucleoside monophosphates containing 4,8-dihydro-4-hydroxy-8-oxo-guanine modifications, which are induced by singlet oxygen, were also measured. A model was devised for the hydrolysis of DNA by nuclease P1 which uses the observed turnover rates as parameters. The model predicts the abundance of monomers and dimers as hydrolysis proceeds. Whereas the level of monomers increases monotonically, the level of each dimer first increases and then falls off. There are advantages to phosphorylating dimers, as compared with monomers, using polynucleotide kinase. Consequently this model may be of interest in connection with 32P-postlabeling applied to the measurement of DNA damage in nuclease P1 partial hydrolysates of DNA.  相似文献   

6.
In order to obtain information about the conformational features of a 2′-O-methylated polyribonucleotide at the nearest neighbor level, a detailed nuclear magnetic resonance study of AmpA was undertaken. AmpA was isolated from alkali hydrolysates of yeast RNA, and proton spectra were recorded at 100 MHz in the Fourier transform mode in D2O solutions, 0.01 M, pH 5.4 and 1.5 at 25°C. 31P spectra were recorded at 40.48 MHz. Complete, accurate sets of nmr parameters derived for each nucleotidyl unit by simulation iteration methods. The nmr data were translated into conformational parameters for all the bonds using procedures developed in earlier studies from these laboratories. It is shown that AmpA exists in aqueous solution with a flexible molecular framework, which shows preferences for certain orientations. The ribose rings exist as a 2E ? 3E equilibrium with the —pA ribose showing a bias for the 3E pucker. The C(4′)—C(5′) bonds of both nucleotidyl units show significant preference (75–80%) to exist in gg conformation. The dominant conformer (80%) about C(5′)—O(5′) of the 5′-nucleotidyl unit is gg′. Even though an unambiguous determination of the orientation of the 3′-phosphate group cannot be made, tentative evidence shows that it preferentially occupies g+ domains [O(3′)—P trans to C(3′)—C(2′)] in which the H(3′) —C(3′)—O(3′)—P(3′) dihedral angle is about 31°. There is reasonable evidence that the 2′-O-methyl preferentially occupies the domain in which the O(2′)—CH3 bond is trans to C(2′)—C(1′). Lowering of pH to 1.5, which results in protonation of both the adenine moieties, causes destacking of AmpA. Such destacking is accompanied by small, but real, perturbations in the conformations about most of the bonds in the backbone. A detailed comparison of the solution conformations of ApA and AmpA clearly shows that 2′-O-methylation strongly influences the conformational preference about the C(3′)—O(3′) bond of the 3′-nucleotidyl unit, in addition to inducing small changes in the overall ribophosphate backbone conformational equilibria. The effect of 2′-O-methylation is such that the C(3′)—O(3′) is forced to occupy preferentially the g+ domain rather than the normally preferred g? domain [O(3′)—P trans to C(3′)—C(4′)] in ApA. The data on ApA and AmpA further reveal that the extent of stacking interaction is less in AmpA compared to ApA. It is suggested that stacked species of AmpA exist as right-handed stacks where the magnitude of ω and ω′ about O(5′)—P and P—O(3′) is about 290°. The reason for the lesser degree of stacking in AmpA compared to ApA is intramolecular interaction between 2′-O-methyl and the flexible O(3′)—P—O(5′) bridge, the interaction causing some perturbation in the magnitudes of ω/ω′, causing destacking. The destacking will lead to an increase in χCN by a few degrees, causing an increase in 2E populations; the latter in turn will shift the 3′ phosphate group from g? to g+ domains. In short, a coupled series of conformational events is envisioned at the onset of destacking, made feasible by the interaction between the 2′-O-methyl group and the swivel O(3′)—P—O(5′) bridge.  相似文献   

7.
The content of 2′-O-methylated dinucleotides from 17 S rRNA of yeast and 18 S rRNAs of chicken cells and Novikoff rat cells was determined by a new procedure using homochromatography. The procedure is simple and can be used to compare the extent of 2′-O-methylation simultaneously in various RNAs. It was of interest that 18 S rRNAs of chicken and rat have two times more 2′-O-methylated dinucleotides as compared to yeast 17 S rRNA.  相似文献   

8.
A dinucleoside bearing an amide internucleotide C3′-CH2-C(O)-NH-C5′ bond was synthesized by the interaction of 3′-deoxy-3′-carboxylmethylribothymidine-2′,3′-lactone obtained by hydrolysis of 2′-O-acetyl-5′-O-benzoyl-3′-deoxy-3′-ethoxycarboxylmethylribothymidine with 5′-deoxy-5′-amino-3′-O-(tert-butyldimethylsilyl)thymidine. After standard manipulations with protective groups, the dinucleoside was converted into 3′-O-(2-cyanoethyl-N,N′-diisopropylphosphoroamidite), which was used for the synthesis of modified oligonucleotides on an automatic synthesizer. Duplex melting curves formed by modified and complementary natural oligonucleotides were measured and the melting temperatures and thermodynamic parameters of duplex formation were calculated. The introduction of one modified bond into oligonucleotides caused only an insignificant decrease in the duplex melting temperatures compared with the nonmodified ones.  相似文献   

9.
Synthesis of a novel ribo-MMI dimer with 2′-OH and 2′-OMe in 5′- and 3′-nucleosides, respectively is presented. The synthesis was accomplished by reductive coupling of 3′-deoxy-3′-C-formyluridine and 2′-O-methyl-5′-O-methylaminouridine via a thioacetal as the key intermediate for the top part of the dimer. Incorporation of ribo- MMI dimers into oligonucleotides increased binding affinity for target RNA.  相似文献   

10.
Seven dinucleoside monophosphates containing 2′-halogeno-2′-deoxypurine nucleoside residue, dAfl-U, dAcl-U, dAbr-U, dAio-U, dGfl-U, and dIfl-C, were chemically synthesized and investigated by 1H-nmr spectroscopy at 300 MHz. The sugar and backbone conformations of these compounds were analyzed by the spectral pattern of furanose proton resonances; and the extents of base-base interaction were estimated from chemical shifts and their temperature-dependent changes of base-proton resonances. It is found that the population of C3′-endo conformer and the extent of base-base interaction decrease as the electronegativity of 2′-substituent decreases in dAx-U (x = fl, cl, br, and io) series. The C3′-endo (3E) population and the base-base interaction in Nfl-U (N = A,G)-type dimers as well as dIfl-C are relatively higher than the corresponding natural ribo-dimers but can be recognized as grossly similar to the conformation of regular RNA dimers.  相似文献   

11.
Five dinucleoside monophosphates containing I-N6-ethenoadenosine (?A) have been studied using fluorescence measurements. The fluorescence spectra of these dinucleoside monophosphates are almost the same as the fluorescence spectrum of ?AMP. Fluorescence quantum yields of these dimers are greatly reduced compared to that of ?AMP. Intramolecular base-base interactions may be responsible for fluorescence quenching. It is found that the fluorescence decay kinetics does not obey a simple decay law but that the decay data can be well described as a sum of three exponentials. This implies that these dimers cannot be characterized as a two-state system, but can be described as systems consisting of three or more conformational states. Sequence effects upon the fluorescence behavior are observed. The fluorescence quenching and decay parameters of Gp?A and Up?A indicate a higher degree of base-base interaction than in their ?ApG and ?ApU counterparts.  相似文献   

12.
RNA 2′-O-methylation is widely distributed and plays important roles in various cellular processes. Mycoplasma genitalium RNase R (MgR), a prokaryotic member of the RNase II/RNB family, is a 3′-5′ exoribonuclease and is particularly sensitive to RNA 2′-O-methylation. However, how RNase R interacts with various RNA species and exhibits remarkable sensitivity to substrate 2′-O-methyl modifications remains elusive. Here we report high-resolution crystal structures of MgR in apo form and in complex with various RNA substrates. The structural data together with extensive biochemical analysis quantitively illustrate MgR’s ribonuclease activity and significant sensitivity to RNA 2′-O-methylation. Comparison to its related homologs reveals an exquisite mechanism for the recognition and degradation of RNA substrates. Through structural and mutagenesis studies, we identified proline 277 to be responsible for the significant sensitivity of MgR to RNA 2′-O-methylation within the RNase II/RNB family. We also generated several MgR variants with modulated activities. Our work provides a mechanistic understanding of MgR activity that can be harnessed as a powerful RNA analytical tool that will open up a new venue for RNA 2′-O-methylations research in biological and clinical samples.  相似文献   

13.
Potential energy calculations were employed to examine the effect of ribose 2′-O-methylation on the conformation of GpC. Minimum energy conformations and allowed conformational regions were calculated for 2′MeGpC and Gp2′MeC. The two lowest energy conformations of 2′MeGpC and Gp2′MeC are similar to those of GpC itself. The helical RNA conformation (sugar pucker-C(3′)-endo, ω′ and ω,g?g?, bases-anti) is the global minimum, and a helix-reversing conformation with ω′, ω in the vicinity of 20°, 80° is next in energy. However, subtle differences between the three molecules are noted. When the substitution is on the 5′ ribose (Gp2′MeC), the energy of the helical conformation is less than that of GpC, due to favorable interactions of the added methyl group. When the substitution is at the 3′ ribose (2′MeGpC) these stabilizing interactions are outweighed by steric restrictions, and the helical conformation is of higher energy than for GpC. Furthermore, the statistical weight of the 2′MeGpC g? g? helical region is substantially less than the corresponding weight for Gp2′MeC. In addition, 2′MeGpC′s methoxy group is conformationally restricted to a narrow range centered at 76°. This group has a broadly allowed region between 50 and 175° in Gp2′MeC. These differences occur because the appended methyl group in 2′MeGpC is located in the interior of the helix cylinder, as it would be in polynucleotide, while it hangs unimpeded in Gp2′MeC. These findings suggest that 2′-O-methylation has both stabilizing and destabilizing influences on the helical conformation of RNA. For 2′MeGpC the destabilizing steric hindrance imposed by the nature of the guanine base dominates.  相似文献   

14.
Cell-free extracts of calamondin orange (Citrus mitis) catalysed the O-methylation of almost all hydroxyls of a number of flavonoids, indicating the existence in citrus tissues of ortho, meta, para and 3-O-methyltransferases. The latter, hitherto unreported enzyme, catalysed the formation of 3-O-methyl ethers of galangin and quercetin. The stepwise O-methylation of a number of compounds, especially quercetin and quercetagetin, tends to suggest a coordinated sequence of O-methylations on the surface of a multienzyme complex. The methyl acceptor abilities of the flavonoid substrates used are discussed in relation to their hydroxyl substitution patterns and their negative electron density distribution.  相似文献   

15.
Abstract

5′-O-(methylphosphonyl)-N-(phenylacetyl)-2 ′-deoxycytidine, deoxyadenosine and deoxyguanosine were pyrophosphorylated and the resulting N-protected P α-methyl nucleoside triphosphates were deblocked by treatment with penicillin amidase at pH 7.8, 25°C to give P α-methyl nucleoside triphosphates.  相似文献   

16.
Abstract

The preponderance of 3′-5′ phosphodiester links in nucleic acids is well known. Albeit less prevalent, the 2′-5′ links are specifically utilised in the formation of ‘lariat’ in group II introns and in the msDNA-RNA junction in myxobacterium. As a sequel to our earlier study on cytidylyl-2′,5′-adenosine we have now obtained the crystal structure of adenylyl-2′,5′-adenosine (A2′p5′A) at atomic resolution. This dinucleoside monophosphate crystallises in the orthorhombic space group P212121 with a= 7.956(3)Å, b = 12.212(3)Å and c = 36.654 (3) Å. CuKα intensity data were collected on a diffractometer. The structure was sloved by direct methods and refined by full matrix least squares methods to R = 10.8 %. The 2′ terminal adenine is in the commonly observed anti (χ2 =?161°) conformation and the 5′ terminal base has a syn (χ1 = 55°) conformation more often seen in purine nucleotides. A noteworthy feature of A2′p5′ A is the intranucleotide hydrogen bond between N3 and 05′ atoms of the 5′ adenine base. The two furanose rings in A2′ p5′ A show different conformations-C2′ endo, C3′ endo puckering for the 5′ and 2′ ends respectively. In this structure too there is a stacking of the purine base on the ribose 04′ just as in other 2′-5′ dinucleoside structures, a feature characteristically seen in the left handed ZDNA. In having syn, anti conformation about the glycosyl bonds, C2′ endo, C3′ endo mixed sugar puckering and N3–05′ intramolecular hydrogen bond A2′p5′ A resembles its 3′-5′ analogue and several other 2′-5′ dinucleoside monophosphate structures solved so far. Striking similarities between the 2′-5′ dinucleoside monophosphate structures suggest that the conformation of the 5′-end nucleoside dictates the conformation of the 2′ end nucleoside. Also, the 2′-5′ dimers do not favour formation of miniature classical double helical structures like the 3′-5′ dimers. It is conceivable, 2–5(A) could be using the stereochemical features of A2′p5′ A which accounts for its higher activity.  相似文献   

17.
The structural properties of uracil photohydrates at the monomer and dimer level in aqueous solution have been examined in detail by nmr spectroscopy. Based on such evidence, the absolute configurations of the two possible diastereomers have been assigned, and the conformational perturbations induced by photohydration have been evaluated. In all instances, photohydration shifts the 2E ? 3E puckering equilibrium of the sugar ring of the uridylyl fragment towards 2E (from 12–18%). In addition, for both dimers examined in detail, ho6UpA and Apho6hU, the effect of dimerization on sugar pucker is such that the 3′-terminal unit shows a clear increase in the percentage of 3E (relative to the appropriate 5′-mononucleotide), whereas the percentage 3E of the 5′-terminal unit shows no change. This is contrary to the findings in the normal dinucleoside monophosphates, where an increased preference for 3E pucker occurs in both residues on dimerization and increased base stacking. Significant base–base interactions were observed in both hydrated dimers despite the loss of the planar π-system in the uracil fragment. In addition, the rate of photohydration for a particular dimer pair (e.g., ApU and UpA or GpU and UpG) is shown to be inversely dependent on the amount of base stacking in the parent dimer. This latter parameter has also been correlated with the ratio of the two possible diastereomers formed in the reaction and is associated with a preferential attack at one face of the pyrimidine base ring. The shift of the sugar puckering equilibrium towards 2E has been compared with similar shifts observed when adenosine and guanosine are methylated at N(1) and N(7), respectively. The possible biological significance of the above-mentioned conformational aspects is discussed.  相似文献   

18.
Two new 5-methyl ether flavone glucosides (7,4′,5′-trihydroxy-5,3′-dimethoxyflavone 7-O-β-D-glucopyranoside and 7,4′-dihydroxy-5-methoxyflavone 7-O-β-D-glucopyranoside) were isolated from the leaves of Thai mangrove Bruguiera gymnorrhiza together with 7,3′,4′,5′-tetrahydroxy-5-methoxyflavone, 7,4′,5′-trihydroxy-5,3′-dimethoxyflavone, luteolin 5-methyl ether 7-O-β-D-glucopyranoside, 7,4′-dihydroxy-5,3′-dimethoxyflavone 7-O-β-D-glucopyranoside, quercetin 3-O-β-D-glucopyranoside, rutin, kaempferol 3-O-rutinoside, myricetin 3-O-rutinoside and an aryl-tetralin lignan rhamnoside. The structure of a lignan rhamnoside was found to be related to racemiside, an isolated compound from Cotoneaster racemiflora, and also discussed. Structure determinations were based on analyses of physical and spectroscopic data including 1D- and 2D-NMR.  相似文献   

19.
Abstract

The synthesis of Methylene(methylimino) or MMI linked nucleoside dimers in all sixteen possible configurations has been accomplished via a reductive coupling of a nucleosidic aldehyde with an hydroxylamine. This has allowed us to prepare all of the necessary 2′-O-methyl MMI dimer building blocks necessary for use in an antisense motif.  相似文献   

20.
Addition of 5-bromo-2′,3′-O-isopropylidene-5′-O-trityluridine (2) in pyridine to an excess of 2-lithio-1,3-dithiane (3) in oxolane at 78° gave (6R)-5,6-dihydro-(1,3-dithian-2-yl)-2′,3′-O-isopropylidene -5′-O-trityluridine (4), (5S,6S)-5-bromo-5,6-dihydro-(1,3-dithian-2-yl)-2′,3′-O-isopropylidene-5′-O-trityluridine (5), and its (5R) isomer 6 in yields of 37, 35, and 10%, respectively. The structure of 4 was proved by Raney nickel desulphurization to (6S)-5,6-dihydro-2′,3′-O-isopropylidene-6-methyl-5′-O-trityluridine (7) and by acid hydrolysis to give D-ribose and (6R)-5,6-dihydro-6-(1,3-dithian-2-yl)uracil (9). Treatment of 4 with methyl iodide in aqueous acetone gave a 30&%; yield of (R,S)-5,6-dihydro-6-formyl-2′,3′-O-isopropylidene-5′-O-trityl-uridine (10), characterized as its semicarbazone 11. Both 5 and 6 gave 4 upon brief treatment with Raney nickel. Both 5 and 6 also gave 6-formyl-2′,3′-O-isopropylidene-5′- O-trityluridine (12) in ~41%; yield when treated with methyl iodide in aqueous acetone containin- 10%; dimethyl sulfoxide. A by-product, identified as the N-methyl derivative (13) of 12 was also formed in yields which varied with the amount of dimethyl sulfoxide used. Reduction of 12 with sodium borohydride, followed by deprotection, afforded 6-(hydroxymethyl)uridine (17), characterized by hydrolysis to the known 6-(hydroxymethyl)uracil (18). Knoevenagel condensation of a mixture of the aldehydes 12 and 13 with ethyl cyanoacetate yielded 38%; of E- (or Z-)6-[(2-cyano-2-ethoxycarbonyl)ethylidene]-2′,3′-O-isopropylidene-5′-O-trityluridine (19) and 10%; of its N-methyl derivative 20. Hydrogenation of 19 over platinum oxide in acetic anhydride followed by deprotection gave R (or S)-6-(3-amino-2-carboxypropyl)uridine (23).  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