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.
Adenylyl (5′,2′)-adenosine 5′-phosphate ((2′-5′)pA-A) was detected in crude crystals of 5′-AMP prepared from Penicillium nuclease (nuclease P1) digest of a technical grade yeast RNA. While (3′–5′)A-A was split by nuclease P1, spleen phosphodiesterase, snake venom phosphodiesterase or alkali, (2′–5′)A-A was not split by a usual level of nuclease P1 or spleen phosphodiesterase. Nuclease P1 digests of 12 preparations of technical grade yeast RNA tested were confirmed to contain (2′–5′)pA-A. Its content was about 1 to 2% of the AMP component of each RNA preparation. As poly(A) was degraded completely by the Penicillium enzyme into 5′-AMP without formation of any appreciable amount of (2′–5′)pA-A, the technical grade RNA is supposed to contain 2–5′ phosphodiester linkages in addition to 3′–5′ major linkages.  相似文献   

2.
Crystals of 5‐fluorouridine (5FUrd) have unit cell dimensions a = 7.716(1), b = 5.861(2), c = 13.041(1)Å, α = γ = 90°, β = 96.70° (1), space group P21, Z = 2, ρobs = 1.56 gm/c.c and ρcalc = 1574 gm/c.c The crystal structure was determined with diffractometric data and refined to a final reliability index of 0.042 for the observed 2205 reflections (I ≥ 3σ). The nucleoside has the anti conformation [χ = 53.1(4)°] with the furanose ring in the favorite C2′–endo conformation. The conformation across the sugar exocyclic bond is g+, with values of 49.1(4) and ? 69.3(4)° for Φθc and Φ respectively. The pseudorotational amplitude τm is 34.5 (2) with a phase angle of 171.6(4)°. The crystal structure is stabilized by a network of N–H…O and O–H…O involving the N3 of the uracil base and the sugar O3′ and O2′ as donors and the O2 and O4 of the uracil base and O3′ oxygen as acceptors respectively. Fluorine is neither involved in the hydrogen bonding nor in the stacking interactions. Our studies of several 5‐fluorinated nucleosides show the following preferred conformational features: 1) the most favored anti conformation for the nucleoside [χ varies from ? 20 to + 60°] 2) an inverse correlation between the glycosyl bond distance and the χ angle 3) a wide variation of conformations of the sugar ranging froni C2′–endo through C3′–endo to C4′–exo 4) the preferred g+ across the exocyclic C4′–C5′ bond and 5) no role for the fluorine atom in the hydrogen bonding or base stacking interactions.  相似文献   

3.
W Saenger  D Suck  M Knappenberg  J Dirkx 《Biopolymers》1979,18(8):2015-2036
The cytostatic drug 6-azauridine is converted in vivo to 6-azauridine-5′-phosphate (z6Urd-5′-P), which blocks the enzyme orotidine-5′-phosphate decarboxylase (Ord-5′-Pdecase) and therefore inhibits the de novo production of uridine-5′-phosphate (Urd-5′-P). In order to relate the structure and function of z6Urd-5′-P, it was crystallized as trihydrate, space group P212121 with a = 20.615 Å, b = 6.265 Å, c = 11.881 Å, and the structure established by Patterson methods. Atomic parameters were refined by full-matrix least-squares methods to R = 0.066 using 1638 counter measured x-ray data. The ribose of z6Urd-5′-P is in a twisted C(2′)-exo, C(3′)endo conformation, the heterocycle is in extreme anti position with angle N(6)-N(1)-C(1′)-O(4′) at 86.3°, and the orientation about the C(4′)-C(5′) bond is gauche, trans in contrast to gauche, gauche found for all the other 5′-ribonucleotides. Conformational energy calculations show that z6Urd-5′-P may adopt an extreme anti conformation not allowed to Urd-5′-P, and they also predict the same unusual trans, gauche conformation about the C(4′)-C(5′) bond in orotidine-5′-phosphate (Ord-5′-P) and in z6Urd-5′-P, which renders the distances O(2)…O(5′) in z6Urd-5′-P and O(7)…O(5′) in Ord-5′-P comparable. On this basis the function of z6Urd-5′-P as an Ord-5′-Pdecase inhibitor can be explained as being due to its structural similarity with the substrate Ord-5′-P and further clarifies the inhibitory action of 5′-nucleotides bearing the heterocycles oxipurinol, xanthine, or allopurinol [J. A. Fyfe, R. L. Miller, and T. A. Krenitsky, J. Biol. Chem. 248 , 3801 (1973)]. With this in mind, new inhibitors for Ord-5′-Pdecase may be designed.  相似文献   

4.
Abstract

An efficient method for the synthesis of 5′-O-monomethoxytrityl-2′,3′-dideoxy-2′-fluoro-3′-thioarabinothymidine [5′-MMTaraF-T3′SH, (5)] and its 3′-phosphoramidite derivative (6) suitable for automated incorporation into oligonucleotides, is demonstrated. A key step in the synthesis involves reaction of 5′-O-MMT-2,3′-O-anhydrothymidine (4) (Eleuteri, A.; Reese, C.B.; Song, Q., J. Chem. Soc. Perkin Trans. 1 1996, 2237 pp.) with sodium thioacetate to give 5′-MMTaraF-T3′SAc (5) (Elzagheid, M.I.; Mattila, K.; Oivanen, M.; Jones, B.C.N.M.; Cosstick, Lönnberg, H. Eur. J. Org. Chem. 2000, 1987–1991). This nucleoside was then converted to its corresponding phosphoramidite derivative, 6, as described previously ((a) Sun, S.; Yoshida, A.; Piccirilli, J.A. RNA, 1997, 3, 1352–1363; (b) Matulic-Adamic, J.; Beigelman, L. Helvetica Chemica Acta 1999, 82, 2141–2150; (c) Fettes, K.J.; O’Neil, I.; Roberts, S.M.; Cosstick, R. Nucleosides, Nucleotides and Nucl. Acids 2001, 20, 1351–1354).  相似文献   

5.
The three-dimensional structure was determined by x-ray crystallography for d(T[p](CE)T), a uv photoproduct of the cyanoethyl (CE) derivative of d(TpT), having the cis-syn cyclobutane (CB) geometry and the S-configuration at the chiral phosphorus atom. The crystals of C23H30N5O12P · 2H2O belong to the orthorhombic space group P212121 (Z = 4), with cell dimensions a = 11.596 Å, b = 14.834 Å, and c = 15.946 Å, containing two water molecules per asymmetric unit. The CB ring is puckered with a dihedral angle of 151°. The two pyrimidine bases are rotated by –29° from the position of direct overlap of their corresponding atoms. This represents a major distortion of DNA, since in DNA adjacent thymines are rotated by +36°. The pyrimidine rings are puckered with Cremer–Pople parameters for T[p] and in parentheses [p]T: Q: 0.24 Å (0.31 Å); θ: 123° (120°); ?: 141° (86°). These represent half-chairs designated as 6H1 (T[p]) and 6H5 ([p]T). The CB and pyrimidine ring conformations are interrelated, and we postulate that they execute a coupled interconversion in solution. The T[p] segment has the syn glycosyl conformation, a 2T3 sugar pucker, and gauche? conformation at C4′-C5′; the [p]T segment is anti, 3T4, trans. The C5′-O5′ torsion of the [p]T unit is –124.5°, and the C3′-O3′ torsion of the T[p] unit is –152.9°. Bond angles and bond lengths involving the phosphorus atom are similar to those of other phosphotriesters. The P-O3′ and P-05′ torsion angles are –138.1° and 58.6°, respectively. Several intermolecular (but no intramolecular) hydrogen bonds are found in the crystal.  相似文献   

6.
Abstract

5′-Chloro-5′-deoxy-N,3′-O-dibenzoylthymidine (3a), 5′-chloro-5′-deoxy-N4, 3′-O-dibenzoyldeoxycytidine(3b), 5′-chloro-5′-deoxy-N6,3′-O-dibenzoyldeoxyadenosine(3c), N-benzoyl-1-(3-chloro-2,3-dideoxy-5-O-trityl-ß-D-xylofuranosyl)thymine (5a) and N6-benzoyl-9-(3-chloro-2,3-dideoxy-5-O-trityl-ß-D-xylofuranosyl)adenine (5b) have been synthesized in very high yields using a new efficient reagent, tris(2,4,6-tribrom-ophenoxy)dichlorophosphorane (BDCP). The reaction time was greatly reduced to 5–8 min. NOE data suggested an inversion of configuration at C3-position and thus an SN2 mechanism has been proposed for the chlorination reaction.

  相似文献   

7.
We have shown previously that the translation of Melon necrotic spot virus (MNSV, family Tombusviridae, genus Carmovirus) RNAs is controlled by a 3′‐cap‐independent translation enhancer (CITE), which is genetically and functionally dependent on the eukaryotic translation initiation factor (eIF) 4E. Here, we describe structural and functional analyses of the MNSV‐Mα5 3′‐CITE and its translation initiation factor partner. We first mapped the minimal 3′‐CITE (Ma5TE) to a 45‐nucleotide sequence, which consists of a stem‐loop structure with two internal loops, similar to other I‐shaped 3′‐CITEs. UV crosslinking, followed by gel retardation assays, indicated that Ma5TE interacts in vitro with the complex formed by eIF4E + eIF4G980–1159 (eIF4Fp20), but not with each subunit alone or with eIF4E + eIF4G1003–1092, suggesting binding either through interaction with eIF4E following a conformational change induced by its binding to eIF4G980–1159, or through a double interaction with eIF4E and eIF4G980–1159. Critical residues for this interaction reside in an internal bulge of Ma5TE, so that their mutation abolished binding to eIF4E + eIF4G1003–1092 and cap‐independent translation. We also developed an in vivo system to test the effect of mutations in eIF4E in Ma5TE‐driven cap‐independent translation, showing that conserved amino acids in a positively charged RNA‐binding motif around amino acid position 228, implicated in eIF4E–eIF4G binding or belonging to the cap‐recognition pocket, are essential for cap‐independent translation controlled by Ma5TE, and thus for the multiplication of MNSV.  相似文献   

8.
Low temperature combined with low light (LL) affects crop production, especially the yield and quality of peppers, in northwest China during the winter and spring seasons. Zeaxanthin (Z) is a known lipid protectant and active oxygen scavenger. However, whether exogenous Z can mitigate LL-induced inhibition of photosynthesis and oxidative stress in peppers remains unclear. In this study, we investigated the effects of exogenous Z on photosynthesis and the antioxidant machinery of pepper seedlings subject to LL stress. The results showed that the growth and photosynthesis of pepper seedlings were significantly inhibited by LL stress. In addition, the antioxidant machinery was disturbed by the uneven production and elimination of reactive oxygen species (ROS), which resulted in damage to the pepper. For example, membrane lipid peroxidation increased ROS content, and so on. However, exogenous application of Z before LL stress significantly increased the plant height, stem diameter, net photosynthetic rate (Pn), and stomata, which were obviously closed at LL. The activities of antioxidant enzymes superoxide dismutase (SOD), catalase (CAT), mono de-hydroascorbate reductase (MDHAR), de-hydroascorbate reductase (DHAR), ascorbate peroxidase (APX), and ascorbate oxidase (AAO) improved significantly due to the increased expression of CaSOD, CaCAT, CaAPX, CaMDHAR, and CaDHAR. The ascorbic (AsA) and glutathione (GSH) contents and ascorbic/dehydroascorbate (AsA/DHA) and glutathione/oxidized glutathione (GSH/GSSG) ratios also increased significantly, resulting in the effective removal of hydrogen peroxide (H2O2) and superoxide anions (O2•−) caused by LL stress. Thus, pre-treatment with Z significantly reduced ROS accumulation in pepper seedlings under LL stress by enhancing the activity of antioxidant enzymes and accumulation of components of the ascorbate–glutathione (AsA–GSH) cycle and upregulated key genes in the AsA–GSH cycle.  相似文献   

9.
R. Krishnan  T. P. Seshadri 《Biopolymers》1994,34(12):1637-1646
Sequence analysis of msDNA from bacterium such as Stigmatella aurantiaca, Myxococcus xanthus and Escherichia coli B revealed that the guanine residue of the single-stranded RNA is linked to the cytosine residue of the msDNA through a 2′–5′ instead of a conventional 3′–5′ phosphodiester bond. We have now obtained the crystal structure of the self-complementary dimer guanylyl-2′,5′-cytidine (G2′p5′C) that occurs at the msDNA-RNA junction. G2′p5′C crystallizes in the orthorhombic space group P212121 with a = 8.376(2), b = 16.231(5), c = 18.671(4). CuK ∝ intensity data were collected on a diffractometer in the ω ?2θ scan mode. The amount of 1699 out of 2354 reflections having I ≥ 3σ (F) were considered observed. The structure was solved by direct methods and refined by full-matrix least squares to a R factor of 0.054. The conformation of the guanine base about the glycosyl bond is syn (χ1 = ?54°) and that of cytosine is anti (χ2 = 156°). The 5′ and 2′ and ribose moieties show C2′-endo and C3-endo mixed puckering just like in A2′p5′A, A2′p5′C, A2′p5U, and dC3′p5′G. Charge neutralization in G2′p5′C is accomplished through protonation of the cytosine base. An important feature of G2′p5′C is the stacking of guanine on ribose 04′ of cytosine similar to that seen in other 2′–5′ dimers. G2′p5′C, unlike its 3′–5′ isomer, does not form a miniature double helix with the Watson-Crick base-pairing pattern. Comparison of G2′p5′C with A2′p5′C reveals that they are isostructural. A branched trinucleotide model for the msDNA-RNA junction has been postulated. © 1994 John Wiley & Sons, Inc.  相似文献   

10.
Two crystals d(C2GCG2) and d(C5GCG5) have been studied under microscope by Fourier transform ir spectroscopy and Raman spectroscopy. The x-ray diffraction study of the latter crystal had shown that the d(C5GCG5) sequence is the first DNA dodecamer known to adopt a canonical A conformation [N. Verdaguer, J. Aymami, D. Fernandez-Forner, I. Fita, M. Coll, T. Huynh-Dinh, J. Igolen, and J. A. Subirana (1991) Journal of Molecular Biology, Vol. 221, pp. 623–635]. Characteristic ir marker bands and Raman marker peaks of the A conformation have thus been obtained and are compared with previously proposed assignments correlated to fiber diffraction x-ray results obtained on polymers. The d(C2GCG2) sequence crystal had previously been studied in an intermediate form between B and Z [L. Urpi, J. P. Ridoux, J. Liquier, N. Verdagner, I. Fita, J. A. Subirana, F. Iglesias, T. Huynh-Dinh, J. Igolen, and E. Taillandier (1989) Nucleic Acids Research, Vol. 17, pp. 6669–6679]. In this paper we present results obtained from a crystal with this oligonucleotide in Z conformation. The effect of the crystallization conditions on the geometry of the obtained oligomer helix is discussed. The influence of the addition, to the central tetramer CGCG, of dCn stretches (at the 5′ end) and dGn stretches (at the 3′ end) of different lengths, on the conformational flexibility of the nucleic acid, is considered. © 1993 John Wiley & Sons, Inc.  相似文献   

11.
A new series of pendant-type polymer-cobalt(III) complexes, [Co(LL)2(BPEI)Cl]2+, (where BPEI?=?branched polyethyleneimine, LL?=?dipyrido[3,2-a:2′,3′-c](6,7,8,9-tetrahydro)phenazine (dpqc), dipyrido[3,2-d:2′,3′-f]quinoxaline (dpq) and imidazo[4,5-f]1,10-phenanthroline (ip)) each with three different degrees of coordination have been synthesized and characterized. Studies to know the mode and strength of interaction between these polymer–metal complexes and calf thymus DNA have been performed by UV–Visible absorption and emission techniques. Among these series, each polymer metal complex having higher binding strength with DNA has been selected to test against human cancer/normal cell lines. On the basis of these spectral studies, it is proposed that our polymer–metal complexes bind with DNA mainly through intercalation along with some electrostatic binding. The order of binding strength for the complexes with ligand, dpqc?>?dpq?>?ip. The analysis of the results suggests that polymer–cobalt(III) complexes with higher degree of coordination effectively binds with DNA due to the presence of large number of positively charged cobalt(III) chelates in the polymer chain which cooperatively act to increase the overall binding strength. These polymer–cobalt(III) complexes with hydrophobic ligands around the cobalt(III) metal centre favour the base stacking interactions via intercalation. All the complexes show very good anticancer activities and increasing of binding strength results in higher inhibition value. The polymer–cobalt(III) complex with dpqc ligand possess two fold increased anticancer activity when compared to complexes with other ligands against MCF-7 cells. Besides, the complexes were insensitive towards the growth of normal cells (HEK-293) at the IC50 concentration.  相似文献   

12.
Polarized Raman scattering measurements have been made of a single crystal of uridylyl(3′–5′)adenosine (UpA) by the use of a Raman microscope with 488.0 nm excitation. The UpA crystal belongs to space group P21 (monoclinic), and Raman intensities Iaa, Ibb, and Ic′c′, have been determined for each Raman band. These intensities correspond to the aa, bb, and c′c′ components of the crystal Raman tensor, where c′ is defined as an axis perpendicular to the crystallographic a axis in the ac plane. From these experimental data, and by taking the known crystal structure into account, anisotropic and isotropic molecular Raman tensors have been calculated for the following 11 normal modes: ring stretching modes of the adenine residue (protonated) at 1560, 1516, 1330, and 715 cm−1; ring stretching modes of the uracil residue at 1696, 1657, 1615, 1228, and 790 cm−1; PO2 symmetric stretching mode at 1080 cm−1; P(—)O single bond stretching mode at 801 cm−1. These pieces of information of the Raman tensors are considered to be useful for estimating the orientations of the DNA and RNA strands in a biological complex from a polarized Raman spectroscopic measurement of such a complex. © 1998 John Wiley & Sons, Inc. Biopoly 45: 135–147, 1998  相似文献   

13.
Net photosynthetic rate of yellow upper leaves (UL) of Ligustrum vicaryi was slightly, but not significantly higher than that of green lower leaves (LL). Diurnally, maximum photochemical efficiency of photosystem 2, PS2 (Fv/Fm) of LL did not significantly decline but the UL showed fairly great daily variations. Yield of PS2 of UL showed an enantiomorphous variation to the photosynthetically active radiation and was significantly lower than in the LL. Unlike Fv/Fm, the efficiency of energy conversion in PS2 and both non-photosynthetic and photosynthetic quenching did not differ in UL and LL. Significant differences between UL and LL were found in contents of chlorophyll (Chl) a, b, and carotenoids (Car) and ratios of Chl a/b, Chl b/Chl (a+b), and Car/Chl (a+b). Leaf colour dichotocarpism in L. vicaryi was mainly caused by different photon utilization; sunflecks affected the LL.  相似文献   

14.
This paper provided a novel approach for evaluating phase stability and elastic properties in metastable Ti–Mo alloys with low Mo content by first-principles combined with cluster structure. In 54-atom body-centered-cubic supercell by substituting Ti atoms with 2–7 Mo atoms (7.1–23.0?wt% Mo), individual cluster structure of β-phase was constructed by ‘-Mo-Ti-Mo-’ cluster unit having the lowest cohesive energy. The distorted supercell was more stable than undistorted one at a low Mo content. With increasing Mo content, the density of state at Fermi level decreased, and bonding electron number increased, indicating β-phase stability was gradually promoted. Tetragonal shear elastic constant (C′?=?(C11?–?C12)/2), shear modulus (G111) and anisotropy factor (A?=?C44/C′) exhibited a fluctuation with Mo addition, while the change trend of A was opposite to C′ and G111. Calculated Young’s modulus exhibited similar changing trend to the C′, implying that the softening of C′ resulted in low Young’s modulus of β-phase. Measured Young’s modulus exhibited significant difference from calculated one, which was mainly caused by formation of α″-martensite and ω-phase. The values of C′, G111 and A were considered to associate with not only elastic properties of β-phase itself but also transition from β-phase to α″-martensite and/or ω-phase.  相似文献   

15.
A strain of Lasiodiplodia mediterranea, a fungus associated with grapevine decline in Sicily, produced several metabolites in liquid medium. Two new dimeric γ‐lactols, lasiolactols A and B ( 1 and 2 ), were characterized as (2S*,3S*,4R*,5R*,2′S*,3′S*,4′R*,5′R*)‐ and (2R*,3S*,4R*,5R*,2′R*,3′S*,4′R*,5′R*)‐(5‐(4‐hydroxymethyl‐3,5‐dimethyl‐tetrahydro‐furan‐2‐yloxy)‐2,4‐dimethyl‐tetrahydro‐furan‐3‐yl]‐methanols by IR, 1D‐ and 2D‐NMR, and HR‐ESI‐MS. Other four metabolites were identified as botryosphaeriodiplodin, (5R)‐5‐hydroxylasiodiplodin, (–)‐(1R,2R)‐jasmonic acid, and (–)‐(3S,4R,5R)‐4‐hydroxymethyl‐3,5‐dimethyldihydro‐2‐furanone ( 3  –  6 , resp.). The absolute configuration (R) at hydroxylated secondary C‐atom C(7) was also established for compound 3 . The compounds 1  –  3 , 5, and 6 , tested for their phytotoxic activities to grapevine cv. Inzolia leaves at different concentrations (0.125, 0.25, 0.5, and 1 mg/ml) were phytotoxic and compound 5 showed the highest toxicity. All metabolites did not show in vitro antifungal activity against four plant pathogens.  相似文献   

16.
Abstract

The oxidation of methyl 5–0-benzyl-3-deoxy-3-fluoro-α-D-arabi-nofuranoside (1) with DMSO/Ac2o afforded a ~ 2:1 mixture of 2-keto derivatives with erythro and threo configuration resulting from isomerization at C3. Successive treatment of the above mixture with MeONH2, LiA1H4, and S-ethyl trifluoroacetate followed by silica gel chromatography afforded methyl 5–0-benzyl-2, 3-dideoxy-3-fluoro-2-(trifluoroacetamido)-α-D-ribofuranoside (6b) and its lyxo isomer 7b in a total yield of 25% and 5%, respectively. The arabino analogue 25 was prepared from 6b. Compounds 6b, 7b and 25 were converted to the corresponding 5–0-benzoyl derivatives 8a, 9 and 26. A series of 2′-amino-2′, 3′-dideoxy-3′-fluoro-β-D-ribo- and-α-D-lyxofuranosides of natural heterocyclic bases have been synthesized starting from 8a and 9. None of the test compounds had any antiviral activity. 3′-Fluoro-2′-amino-2′, 3′-dideoxycytidine (16) was the only compound showing inhibition of murine L1210 and human Molt/4F cell proliferation (50% effective concentration: 39–42μg/m1).  相似文献   

17.
Edward A. Berry  Dong-Woo Lee  Kazuo Nagai 《BBA》2010,1797(3):360-7281
Ascochlorin is an isoprenoid antibiotic that is produced by the phytopathogenic fungus Ascochyta viciae. Similar to ascofuranone, which specifically inhibits trypanosome alternative oxidase by acting at the ubiquinol binding domain, ascochlorin is also structurally related to ubiquinol. When added to the mitochondrial preparations isolated from rat liver, or the yeast Pichia (Hansenula) anomala, ascochlorin inhibited the electron transport via CoQ in a fashion comparable to antimycin A and stigmatellin, indicating that this antibiotic acted on the cytochrome bc1 complex. In contrast to ascochlorin, ascofuranone had much less inhibition on the same activities. On the one hand, like the Qi site inhibitors antimycin A and funiculosin, ascochlorin induced in H. anomala the expression of nuclear-encoded alternative oxidase gene much more strongly than the Qo site inhibitors tested. On the other hand, it suppressed the reduction of cytochrome b and the generation of superoxide anion in the presence of antimycin A3 in a fashion similar to the Qo site inhibitor myxothiazol. These results suggested that ascochlorin might act at both the Qi and the Qo sites of the fungal cytochrome bc1 complex. Indeed, the altered electron paramagnetic resonance (EPR) lineshape of the Rieske iron-sulfur protein, and the light-induced, time-resolved cytochrome b and c reduction kinetics of Rhodobacter capsulatus cytochrome bc1 complex in the presence of ascochlorin demonstrated that this inhibitor can bind to both the Qo and Qi sites of the bacterial enzyme. Additional experiments using purified bovine cytochrome bc1 complex showed that ascochlorin inhibits reduction of cytochrome b by ubiquinone through both Qi and Qo sites. Moreover, crystal structure of chicken cytochrome bc1 complex treated with excess ascochlorin revealed clear electron densities that could be attributed to ascochlorin bound at both the Qi and Qo sites. Overall findings clearly show that ascochlorin is an unusual cytochrome bc1 inhibitor that acts at both of the active sites of this enzyme.  相似文献   

18.
Abstract

Moffatt oxidation of 2′,3′-O-isopropylidenearisteromycin (1a) and treatment of the 5′-carboxaldehyde with [(p-tolylsulfonyl)methylene]triphenylphosphorane gave the homologated vinylsulfone 2. Treatment of 2 with tributylstannane/AIBN gave the (E/Z)-vinylstannanes which were converted into the E and Z fluoro- and iodovinyl analogs. Chain extension via the 5′-cyano-5′-deoxy derivative 10a gave the 6′-carboxaldehyde of homoaristeromycin. S-Adenosyl-L-homocysteine hydrolase was strongly inhibited by the fluorovinyl, 5b, and iodovinyl, 4b and 7b, compounds, and time-dependent kinetics were observed [1–2 μM (Ki) and 0.1–0.2 min?1 (k inact)]. The mechanism of inactivation was shown to involve addition of water at the vinyl 5′ or 6′ carbons with elimination of halide.

  相似文献   

19.
Molecular-mechanics calculations have been carried out on the base-paired hexanucleoside pentaphosphates d(TATATA)2, d(ATATAT)2, d(A6)·d(T6), d(CGCGCG)2, d(GCGCGC)2, and d(C6)·d(G6) in both A- and B-DNA geometries. The calculated relative energies of these polymers are consistent with the relative stabilities of the polymers found experimentally. In particular, the results of our calculations support the observation that the homopolymer d(A)n·d(T)n is more stable in a B-DNA conformation, while the homopolymer d(G)n·d(C)n is more stable in an A-DNA conformation. The molecular interactions responsible for these differential stabilities include both inter- and intrastrand base stacking, as well as base–phosphate interactions. While definitive experiments on the heteropolymer stabilities have not yet been carried out, the results of our calculations also suggest a greater stability of the purine-3′,5′-pyrimidine sequence over the pyrimidine-3′,5′-purine sequence in both the A- and B-conformations. The reason for this greater stability lies in the importance of the inherent directionality (5′ → 3′ vs 3′ → 5′) of phosphate–base and base–base interactions. The largest conformation change observed on energy refinement is sugar repuckering, which occurs mainly on pyrimidine-attched sugars and only in the B-DNA geometry. We suggest a molecular mechanism, specifically, differential base–sugar steric interactions involving neighboring sugars, to explain why this repuckering occurs more with d(A6)·d(T6) than with other isomers.  相似文献   

20.
Abstract

Synthetic oligoribonucleotides (RNA) are efficiently prepared with 2′-O-tert-butyldimethylsilyl nucleoside 3′-O-phosphoramidites with labile base-protection; Admf or APac, Gdmf, Cibu, U. After cleavage from the polystyrene support, the exocyclic amine protecting groups are removed with conc. NH4OH: ethanol/3:1 by heating at 55°C for 3–5 h. The 2′-O- silyl protecting groups are removed with tetra-n-butylammonium fluoride in THF or more conveniently with neat triethylamine trihydrofluoride. To gain the advantages of increased capacity on reverse phase HPLC and the convenience of cartridge based purification (OPC, Oligonucleotide Purification Cartridge), the 5′ trityl was left on the RNA as the final protecting group to be removed. The mild conditions which are effective for trityl removal are shown to preserve 3′-5′ phosphate linkage integrity in RNA. The absence of phosphate migration is demonstrated by model studies, utilizing N4 -isobutyryl-5′-O-DMT-3′-O-TBDMS-2′-O-(2-cyanoethyl-N,N-diisopropylphosphoramidite) as a control monomer and digestion by 3′-5′ selective P1 nuclease and alkaline phosphatase and HPLC analysis. Oligoribonucleotides were analyzed by Microgel capillary electrophoresis, anion-exchange HPLC, and the enzymatic digest/HPLC method.

  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