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 78 毫秒
1.
2.
Small noncoding HIV-1 leader exon 3 is defined by its splice sites A2 and D3. While 3′ splice site (3′ss) A2 needs to be activated for vpr mRNA formation, the location of the vpr start codon within downstream intron 3 requires silencing of splicing at 5′ss D3. Here we show that the inclusion of both HIV-1 exon 3 and vpr mRNA processing is promoted by an exonic splicing enhancer (ESEvpr) localized between exonic splicing silencer ESSV and 5′ss D3. The ESEvpr sequence was found to be bound by members of the Transformer 2 (Tra2) protein family. Coexpression of these proteins in provirus-transfected cells led to an increase in the levels of exon 3 inclusion, confirming that they act through ESEvpr. Further analyses revealed that ESEvpr supports the binding of U1 snRNA at 5′ss D3, allowing bridging interactions across the upstream exon with 3′ss A2. In line with this, an increase or decrease in the complementarity of 5′ss D3 to the 5′ end of U1 snRNA was accompanied by a higher or lower vpr expression level. Activation of 3′ss A2 through the proposed bridging interactions, however, was not dependent on the splicing competence of 5′ss D3 because rendering it splicing defective but still competent for efficient U1 snRNA binding maintained the enhancing function of D3. Therefore, we propose that splicing at 3′ss A2 occurs temporally between the binding of U1 snRNA and splicing at D3.  相似文献   

3.
Solid-phase synthesis has contributed to a significant portion of combinatorial libraries made to date. Among many advantages of solid-phase organic synthesis, one unique property is the pseudo-dilution effect. In reality, polymer support is a dynamic system and its structural mobility is controlled by many intrinsic and external factors such as cross-linking, load, solvent, temperature and incoming reagent concentrations. Site-site interaction can often occur and affect the direction of the intended reaction.  相似文献   

4.
5.
6.
7.
Abstract

The structure of 2′,3′-didehydro-2′,3′-dideoxyguanosine was determined by X-ray crystallographic analysis of the complex with pyridine. The two independent nucleoside molecules have similar, commonly observed glycosyl link (x = -102.3° and -94.2°) and 5′-hydroxyl (y = 54.0° and 47.6°) conformations. The five-membered rings are very planar with r.m.s. deviations from planarity of less than 0.015 A. 2′,3′-Didehydro-2′,3′-dideoxyadenosine has a similar glycosyl link conformation but a different 5′-hydroxyl group orientation and a slightly less planar 5-membered ring.  相似文献   

8.
9.
Abstract

The nucleoside constituents of nucleic acids prefer the anti conformation (1). When the sugar pucker is taken into account the nucleosides prefer the C2′endo-anti conformation. Of the nearly 300 nucleosides known, about 250 are in the anti conformation and 50 are in the syn-conformation, i.e., anti to syn conformation is 5:1. The nucleotide building blocks of nucleic acids show the same trend as nucleosides. Both the deoxy-guanosine and ribo- guanosine residues in nucleosides and nucleotides prefer the syn-C2′endo conformation with an intra-molecular hydrogen bond (for nucleosides) between the O5′- H and the N3 of the base and, a few syn-C3′endo conformations are also observed. Evidence is presented for the occurrence of the C3′endo-syn conformation for guanines in mis-paired double helical right-handed structures with the distorted sugar phosphate C4′-C5′ and P-O5′ bonds respectively, from g+ (gg) and g- to trans. Evidence is also provided for guanosine nucleotides in left-handed double-helical (Z-DNA) oligo and polynucleotides which has the same syn-C3′endo conformation and the distorted backbone sugar-phosphate bonds (C4′-C5′ and P- O5′) as in the earlier right-handed case.  相似文献   

10.
Abstract

A new approach to the synthesis of 2′,3′-dideoxyadenosine and 2′,3′-dideoxyinosine based on deoxygenation of 2′,3′-di-O-mesylnucleosides was developed.  相似文献   

11.
Abstract

A simple procedure is described for the preparation of the title compounds 1, 8 and 9. 3′-3′ or 3′-5′ or 5′-5′ TpT was reacted with a twofold molar excess of TPS in anhydrous DMF, at room temperature, for 5 min, followed by a 1 min in situ treatment of the reaction mixture with excess 7.0 N NH4OH, at 0°C. The alkaline hydrolysis of 1, 8 and 9 proceeds without the assistance of 3′- and 5′-hydroxyl groups resulting in equimolar mixtures of thymidine (4) and thymidine 3′-phosphoramidate (6) (for the 3′-3′ isomer) or thymidine 5′-phosphoramidate (7) (for the 5′-5′ isomer) or 6 and 7 in equal quantities (for the 3′-5′ isomer).  相似文献   

12.
Many alleles of human disease genes have mutations within splicing consensus sequences that activate cryptic splice sites. In Caenorhabditis elegans, the unc-73(e936) allele has a G-to-U mutation at the first base of the intron downstream of exon 15, which results in an uncoordinated phenotype. This mutation triggers cryptic splicing at the −1 and +23 positions and retains some residual splicing at the mutated wild-type (wt) position. We previously demonstrated that a mutation in sup-39, a U1 snRNA gene, suppresses e936 by increasing splicing at the wt splice site. We report here the results of a suppressor screen in which we identify three proteins that function in cryptic splice site choice. Loss-of-function mutations in the nonessential splicing factor smu-2 suppress e936 uncoordination through changes in splicing. SMU-2 binds SMU-1, and smu-1(RNAi) also leads to suppression of e936. A dominant mutation in the conserved C-terminal domain of the C. elegans homolog of the human tri-snRNP 27K protein, which we have named SNRP-27, suppresses e936 uncoordination through changes in splicing. We propose that SMU-2, SMU-1, and SNRP-27 contribute to the fidelity of splice site choice after the initial identification of 5′ splice sites by U1 snRNP.PRE-mRNA splicing takes place in a large ribonucleoprotein complex called the spliceosome (Burge et al. 1999). Components of this splicing machinery assemble at conserved signal sequences within the pre-mRNA. The 5′ splice site consensus sequence M−3A−2G−1 | G+1U+2R+3A+4G+5U+6 and the 3′ splice site consensus sequence Y−3A−2G−1 | R+1 (M is either A or C; R is a purine, and Y is a pyrimidine) define the limits of the intron. Base-pairing interactions between the 5′ end of the U1 snRNA and the 5′ splice site consensus sequence occur early in spliceosome assembly. It is the nearly invariable GU dinucleotide at the first two positions of the 5′ end of the intron that defines the beginning of the intron. The 5′ consensus sequence is essential but insufficient for splice site selection, as 5′ splice sites with weaker consensus matches may require additional determinants for proper activation (Sanford et al. 2005).Mutations that disrupt the 5′ consensus splice signal can lead to genetic disease in humans (Nelson and Green 1990; Cohen et al. 1994). Approximately 15% of point mutations that cause genetic diseases affect pre-mRNA splicing consensus sequences (Krawczak et al. 1992). For some specific disease genes, as many as 50% of the known heritable alleles alter splicing (Teraoka et al. 1999; Ars et al. 2000; Roca et al. 2003; Pagenstecher et al. 2006). Among all the positions of the 5′ splice site consensus sequence, the highest proportion of human disease mutations occur at the +1G position (Buratti et al. 2007). The fidelity of pre-mRNA splice site choice is largely disrupted by this defect, since this mutation causes splicing at this site to be either abolished or outcompeted by the activation of nearby cryptic 5′ splice sites (Nelson and Green 1990; Cohen et al. 1994). Cryptic splice sites are used only when the wild-type splice donor is disrupted by mutation, as they tend to have very weak splice donor consensus sequences outside of a 5′-GU dinucleotide that defines the beginning of the intron (Roca et al. 2003). Suppression of mutations to the 5′ splice site consensus sequence in vivo has been achieved through the expression of U1 snRNAs containing compensatory base substitutions (Zhuang and Weiner 1986); however, suppression of mutations to the +1 position of the intron using reverse genetic approaches has not been successful (Newman et al. 1985; Nelson and Green 1990; Cohen et al. 1994).We have used a specific allele of the Caenorhabditis elegans unc-73 gene, e936, which contains a G-to-U mutation at the first nucleotide of intron 16 (Steven et al. 1998), as a model for studying cryptic splice site choice (Roller et al. 2000; Zahler et al. 2004). unc-73 encodes a RAC guanine nucleotide exchange factor that is expressed in neurons and is important for axon guidance (Steven et al. 1998). The e936 allele induces the use of three different cryptic 5′ splice sites (Figure 1A). Two of these 5′ splice sites, located at the −1 and +23 positions, define introns beginning with GU. The third 5′ splice site used is at the mutated wild-type (wt) position and is referred to as “wt” since splicing at this site still produces wild-type unc-73 mRNA and protein, even though the intron begins with UU (Roller et al. 2000). Use of either the −1 or the +23 cryptic site causes a shift in the reading frame and loss of gene function. In e936 animals, 90% of the stable messages of unc-73 are out-of-frame, yet the phenotype is not as severe as for other alleles in this gene. This indicates that the 10% of steady-state messages that are in frame have some functional role.Open in a separate windowFigure 1.—(A) Diagram of the unc-73 gene between exons 15 and 16. The positions of the −1 and +23 cryptic 5′ splice sites are indicated by arrows. The intronic e936 (+1G → U) point mutation is highlighted. (B) γ-32P-labeled RT–PCR results across the cryptic splicing region of unc-73(e936) for different strains. Lanes 1, 2, and 3 are loaded with RT–PCR reactions from wild type (N2), unc-73(e936);sup-39(je5), and unc-73(e936) RNA, respectively. The lines carrying the suppressor alleles and e936 follow in lanes 4–10 as indicated. (C) The unc-73 genomic sequence from exon 15 (uppercase letters) and intron 15 (lowercase letters). The locations of the az23 and e936 mutational substitutions are indicated below. The position of the −9 cryptic splice donor activated in e936az23 is indicated by an arrow above.In a previous genetic screen for extragenic suppressors of e936 movement defects, Way and colleagues identified sup-39 (Run et al. 1996). It was subsequently shown that mutations in sup-39 alter cryptic splice site choice of e936 (Roller et al. 2000). sup-39 encodes a U1 snRNA gene with a compensatory mutation at the position that normally base pairs with the +1G. This allows sup-39 to base pair with an intron with a +1U (Zahler et al. 2004). This dominant suppressor increases usage of the mutated splice site and improves the fraction of in-frame messages from e936 from 10 to 33%, with a dramatic improvement in coordination. A similar mutant U1 snRNA suppressor with a different compensatory substitution, sup-6(st19), was found to suppress the intronic +1G to A transition of unc-13(e309) to allow for splicing at the mutated wild-type site, even though the intron begins with AU instead of GU (Zahler et al. 2004).We are interested in identifying additional factors that play a role in cryptic 5′ splice site choice. To do this, we took advantage of unc-73(e936), in which modest increases in the use of the wt splice site lead to dramatic increases in coordination, as a sensitive screen for changes in cryptic splice site choice. In this article we report that the proteins SMU-1 and SMU-2, which are nonessential factors previously shown to have a role in alternative splicing (Spartz et al. 2004), have a role in selection of cryptic 5′ splice sites. We also report the identification of a new dominant suppressor of cryptic splicing, snrp-27, which encodes a C. elegans homolog of the human tri-snRNP 27K protein.  相似文献   

13.
The human immunodeficiency virus type 1 (HIV-1) accessory protein Vif is encoded by an incompletely spliced mRNA resulting from splicing of the major splice donor in the HIV-1 genome, 5′ splice site (5′ss) D1, to the first splice acceptor, 3′ss A1. We have shown previously that splicing of HIV-1 vif mRNA is tightly regulated by suboptimal 5′ss D2, which is 50 nucleotides downstream of 3′ss A1; a GGGG silencer motif proximal to 5′ss D2; and an SRp75-dependent exonic splicing enhancer (ESEVif). In agreement with the exon definition hypothesis, mutations within 5′ss D2 that are predicted to increase or decrease U1 snRNP binding affinity increase or decrease the usage of 3′ss A1 (D2-up and D2-down mutants, respectively). In this report, the importance of 5′ss D2 and ESEVif for avoiding restriction of HIV-1 by APOBEC3G (A3G) was determined by testing the infectivities of a panel of mutant viruses expressing different levels of Vif. The replication of D2-down and ESEVif mutants in permissive CEM-SS cells was not significantly different from that of wild-type HIV-1. Mutants that expressed Vif in 293T cells at levels greater than 10% of that of the wild type replicated similarly to the wild type in H9 cells, and Vif levels as low as 4% were affected only modestly in H9 cells. This is in contrast to Vif-deleted HIV-1, whose replication in H9 cells was completely inhibited. To test whether elevated levels of A3G inhibit replication of D2-down and ESEVif mutants relative to wild-type virus replication, a Tet-off Jurkat T-cell line that expressed approximately 15-fold-higher levels of A3G than control Tet-off cells was generated. Under these conditions, the fitness of all D2-down mutant viruses was reduced relative to that of wild-type HIV-1, and the extent of inhibition was correlated with the level of Vif expression. The replication of an ESEVif mutant was also inhibited only at higher levels of A3G. Thus, wild-type 5′ss D2 and ESEVif are required for production of sufficient Vif to allow efficient HIV-1 replication in cells expressing relatively high levels of A3G.Human immunodeficiency virus type 1 (HIV-1) Vif is a 23-kDa basic protein (4, 9) that is incorporated into virus particles during productive infection (8-10). Replication of HIV-1 in some T-cell lines is dependent on the expression of a functional Vif protein. Replication of Vif-deleted HIV-1 is restricted in these cells, which are termed nonpermissive, because of the presence of several host deaminases, the most important of which for HIV-1 replication is APOBEC3G (A3G) (25, 26). Human A3G is a single-stranded DNA deaminase that inhibits the replication of HIV-1 as well as other types of retroviruses and retrotransposons (5, 12, 17, 25, 32). HIV-1 Vif forms a complex with A3G and other cellular proteins to promote A3G ubiquitination, resulting in proteasomal degradation of A3G (1, 11, 14, 18, 26). Vif-deleted HIV-1 produced in the presence of A3G packages increased levels of A3G compared to those found in the wild type (WT) and has reduced infectivity in nonpermissive T-cell lines. This reduced infectivity in the absence of Vif has been correlated with the dC-to-dU hypermutation of newly synthesized minus-strand viral DNA by A3G (6, 13, 31, 32). However, other studies have shown that A3G is also able to restrict virus replication without hypermutating viral DNA (7, 19).It has previously been shown that the expression of Vif in infected cells is maintained at a relatively low level compared to levels of the other HIV-1 accessory proteins. One mechanism to explain this phenomenon is that Vif is degraded more rapidly than other accessory proteins by the proteasome (3). Another mechanism is that a relatively low level of vif mRNA is produced by alternative splicing (22). Alternative splicing of HIV-1 RNA results in the production of approximately 40 different mRNA species, which include three different mRNA size classes: 1.8-kb, completely spliced RNAs; 4-kb, incompletely spliced RNAs; and 9-kb, unspliced RNAs (Fig. (Fig.1A).1A). The 4-kb mRNA class encodes Vif, Vpr, Tat, Vpu, and Env, and the completely spliced, 1.8-kb mRNA class encodes Tat, Rev, and Nef. Unspliced viral RNA is both packaged into virions as genomic RNA and used as mRNA for Gag and Gag-Pol proteins (2, 27). As shown in Fig. Fig.1A,1A, four different 5′ splice donor sites (5′ss) and eight different 3′ splice acceptor sites (3′ss), which are highly conserved among group M HIV-1 strains, are used to produce alternatively spliced HIV-1 mRNAs at different levels in infected cells (22). The efficiencies with which these 5′ss and 3′ss are used are dependent on the presence of suboptimal cis splicing elements within the 5′ss and 3′ss themselves and more-distant elements, which include exonic splicing silencers, an intronic splicing silencer, and exonic splicing enhancers (ESE) (2, 15, 27).Open in a separate windowFIG. 1.HIV-1 splicing pattern and elements regulating vif mRNA splicing. (A) The conserved 5′ss (D1 to D4) and 3′ss (A1 to A7) located within the 9-kb HIV-1 genome are shown. Completely and incompletely spliced HIV-1 mRNAs (∼4 kb and ∼1.8 kb) are shown as open boxes. Spliced mRNAs are denoted by the translated open reading frame and by the exon content. The incompletely spliced mRNAs, denoted with an I, are differentiated from completely spliced mRNAs by inclusion of the intron between 5′ss D4 and 3′ss A7. Either one or both of the noncoding exons 2 and 3 (shown as gray-shaded exons) can be differentially included within all 1.8- and 4.0-kb mRNA species, with the exception of vif mRNA (1.2I) and vpr mRNA, which can include only exon 2 (1.[2].3I). LTR, long terminal repeat. (B) Three elements regulating vif mRNA splicing are shown: positively acting enhancer ESEVif, the 5′ss D2 (underlined), and a negatively acting G4 silencer motif. The locations of noncoding exon 2 and the start site for Vif protein synthesis are also shown. (C) HIV-1 5′ss D2-down mutants used in this study are shown. Sequences were aligned and compared with that of the consensus metazoan 5′ss. The sequence of the ESEVif mutant used in this study is also aligned and compared with the WT sequence. nt, nucleotides.HIV-1 Vif is translated from a low-abundance, incompletely spliced mRNA resulting from splicing of HIV-1 RNA between the major splice donor site (5′ss D1) and 3′ss A1. We have demonstrated that vif mRNA splicing is tightly regulated by the presence of multiple regulatory elements (Fig. (Fig.1B).1B). These include a highly conserved suboptimal 5′ss (5′ss D2) 50 nucleotides downstream from 3′ss A1, an SRp75-dependent ESE (ESEVif), and a GGGG silencer element proximal to 5′ss D2 (2). Mutations within the relatively weak 5′ss D2 that increase its homology to a consensus 5′ss result in increased inclusion of the noncoding 50-nucleotide exon defined by 3′ss A1 and 5′ss D2 (exon 2), increased single-spliced vif mRNA levels and Vif expression, and an excessive splicing phenotype in which virion production is reduced to 10 to 25% of that of the WT (16). Conversely, mutations that decrease the homology of 5′ss D2 to a consensus 5′ss inhibit splicing at 3′ss A1 and exon 2 inclusion into both incompletely and completely spliced HIV-1 mRNAs as well as decreased levels of vif mRNA. Virus production, however, is not significantly affected. Mutation of ESEVif resulted in a similar phenotype. We have shown previously that increased or decreased exon 2 inclusion into spliced mRNA does not affect the stability or expression of viral mRNAs (15). Based on these results, we hypothesized that the conserved suboptimal 5′ss D2, which together with 3′ss A1 defines exon 2, and ESEVif are necessary to maintain optimal levels of Gag and Gag-Pol required for HIV-1 replication while producing sufficient Vif to overcome the cellular restriction factor A3G (2). To further test this hypothesis, we examined a panel of HIV-1 mutants producing reduced levels of Vif under permissive and nonpermissive conditions. We also investigated the long-term replication capabilities of these mutant viruses in both permissive and nonpermissive A3G-expressing T-cell lines. Mutant viruses demonstrated increasing sensitivity to A3G, which is inversely proportional to their levels of Vif expression. Our results suggest that the reason 5′ss D2 and ESEVif exist in the HIV-1 genome is to regulate the levels of vif mRNA and Vif protein in infected cells.  相似文献   

14.
Summary The degree of methylation at the 5 and 3 CCGG sequences flanking the variable number of tandem repeat (VNTR) region of the DXS255 locus at Xp11.22 was analysed separately in several haematopoietic cell lineages. The 5 CCGG site on active chromosomes was found to be completely methylated in B and T lymphocytes and granulocytes. Methylation of the 5 site on inactive X chromosomes differed between females (0%–60%), but was consistent in different cell lineages obtained from individual females. In contrast, methylation at the 3 CCGG site on active chromosomes was found to vary in B lymphocytes (40%–100%), whereas complete methylation was found in T lymphocytes and granulocytes. The extent of methylation on inactive X chromosomes was found to differ significantly between B lymphocytes (17%), T lymphocytes (54%) and granulocytes (82%). Thus, methylation at the 5 CCGG site seems to be primarily related to the status of X chromosome inactivation, whereas methylation at the 3 CCGG site is mainly subject to cell-lineage-specific influences.  相似文献   

15.
Abstract

The 5′-O-(4,4′-dimethoxytrityl) and 5′-O-(tert-butyldimethylsilyl) derivatives of 2′-,3′-O-thiocarbonyl-6-azauridine and 2′,3′-O-thiocarbonyl-5-chlorouridine were synthesized from the parent nucleosides by reaction with 4, 4′-dimethoxytrityl chloride and tert-butyldimethylsilyl chloride, respectively, followed by treatment with 1,1′-thiocarbonyldiimidazole. Introduction of a 2′-,3′-double bond into the sugar ring by reaction of the 5′-protected 2′-,3′-O-thionocarbonates with 1, 3-dimethyl-2-phenyl-1, 3, 2-diazaphospholidiine was unsuccessful, but could be accomplished satisfactorily with trimethyl phosphite. Reactions were generally more successful with the 5′-silylated than with the 5′-tritylated nucleosides. Formation of 2′-,3′-O-thiocarbonyl derivatives proceeded in higher yield with 5′-protected 6-azauridines than with the corresponding 5-chlorouridines because of the propensity of the latter to form 2,2′-anhydro derivatives. In the reaction of 5′-O-(tert-butyldimethylsilyl)-2′-,3′-O-thiocarbonyl-6-azauridine with trimethyl phosphite, introduction of the double bond was accompanied by N3-methylation. However this side reaction was not a problem with 5′-O-(tert-butyldimethylsilyl)-2′-, 3′-O-thioarbonyl-5-chlorouridine. Treatment of 5′-O-(tert-butyldimethylsilyl)-2′-, 3′-didehydro-2′-,3′-dideoxy-6-azauridine with tetrabutylammonium fluoride followed by hydrogenation afforded 2′-,3′-dideoxy-6-azauridine. Deprotection of 5′-O-(tert-butyldimethylsilyl)-2′-, 3′-didehydro-2′-,3′-dideoxy-5-chlorouridine yielded 2′-,3′-didehydro-2′-,3′-dide-oxy-5-chlorouridine.  相似文献   

16.
The RNA genome of the hepatitis C virus (HCV) contains multiple conserved structural cis domains that direct protein synthesis, replication, and infectivity. The untranslatable regions (UTRs) play essential roles in the HCV cycle. Uncapped viral RNAs are translated via an internal ribosome entry site (IRES) located at the 5′ UTR, which acts as a scaffold for recruiting multiple protein factors. Replication of the viral genome is initiated at the 3′ UTR. Bioinformatics methods have identified other structural RNA elements thought to be involved in the HCV cycle. The 5BSL3.2 motif, which is embedded in a cruciform structure at the 3′ end of the NS5B coding sequence, contributes to the three-dimensional folding of the entire 3′ end of the genome. It is essential in the initiation of replication. This paper reports the identification of a novel, strand-specific, long-range RNA–RNA interaction between the 5′ and 3′ ends of the genome, which involves 5BSL3.2 and IRES motifs. Mutants harboring substitutions in the apical loop of domain IIId or in the internal loop of 5BSL3.2 disrupt the complex, indicating these regions are essential in initiating the kissing interaction. No complex was formed when the UTRs of the related foot and mouth disease virus were used in binding assays, suggesting this interaction is specific for HCV sequences. The present data firmly suggest the existence of a higher-order structure that may mediate a protein-independent circularization of the HCV genome. The 5′–3′ end bridge may have a role in viral translation modulation and in the switch from protein synthesis to RNA replication.  相似文献   

17.
Summary The Spec1 and Spec2 mRNAs (Strongylocentrotus purpuratus ectoderm mRNAs) represent a small gene family that encodes 10–12 members of the troponin C superfamily of calcium-binding proteins. These mRNAs and proteins accumulate in the aboral (dorsal) ectoderm of sea urchin embryos and larvae. Using genomic and cDNA clones, we have compared the sequences of four Spec mRNAs: Spec1, Spec2a, Spec2c, and Spec2d. The mRNAs all have at least 120 bases of 5 untranslated leader, approximately 450 bases of open reading frame, and 900 bases (Spec1) or 1250 bases (Spec2a, 2c, 2d) of 3 untranslated trailer. Unexpectedly, when long stretches of 5 untranslated regions or 3 untranslated regions are compared to one another, they are found to be less divergent than the protein-coding regions. Comparing Spec2d, the most divergent member of the family, with the other Spec mRNAs shows that while the protein-coding regions are 60–62% matched, the untranslated regions are greater than 80% matched. Comparisons among Spec1, Spec2a, and Spec2c demonstrate similar but less dramatic conservation of untranslated regions. Our data imply that the Spec gene family has evolved differently from most gene families, with mutations accumulating most rapidly in intron regions, less rapidly in protein-conding regions, and least rapidly in 5 and 3 untranslated regions.  相似文献   

18.
19.

Since the discovery of 3′-azido-3′-deoxythymidine (AZT) and 2′,3′-didehydro-2′,3′-dideoxythymidine (d4T) as potent and selective inhibitors of the replication of human immunodeficiency virus (HIV), there has been a growing interest for the synthesis of 2′,3′-didehydro-2′,3′-dideoxynucleosides with electron withdrawing groups on the sugar moiety. Here we described an efficient method for the synthesis of such nucleoside analogs bearing structural features of both AZT and d4T. The key intermediate, 3-azido-1,2-bis-O-acetyl-5-O-benzoyl-3-deoxy-D-ribofuranose, 5 was synthesized from commercially available D-xylose in five steps, from which a series of pyrimidine and purine nucleosides were synthesized in high yields. The resultant protected nucleosides were converted to target nucleosides using appropriate chemical modifications. The final nucleosides were evaluated as potential anti-HIV agents.  相似文献   

20.
Abstract

The title compound was prepared by reduction of the oxime of the 3′-ketouridine. Condensation with aldehydes gave a series of nitrones whose reduction afforded “second generation” hydroxylamines, some of which showing antiviral activity. The nitroxide free radicals formed upon oxidation of hydroxylamines gave good e.s.r. spectra useful for configurational and conformational assignments.  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