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.
A series of 7‐fluorinated 7‐deazapurine 2′‐deoxyribonucleosides related to 2′‐deoxyadenosine, 2′‐deoxyxanthosine, and 2′‐deoxyisoguanosine as well as intermediates 4b – 7b, 8, 9b, 10b , and 17b were synthesized. The 7‐fluoro substituent was introduced in 2,6‐dichloro‐7‐deaza‐9H‐purine ( 11a ) with Selectfluor (Scheme 1). Apart from 2,6‐dichloro‐7‐fluoro‐7‐deaza‐9H‐purine ( 11b ), the 7‐chloro compound 11c was formed as by‐product. The mixture 11b / 11c was used for the glycosylation reaction; the separation of the 7‐fluoro from the 7‐chloro compound was performed on the level of the unprotected nucleosides. Other halogen substituents were introduced with N‐halogenosuccinimides ( 11a → 11c – 11e ). Nucleobase‐anion glycosylation afforded the nucleoside intermediates 13a – 13e (Scheme 2). The 7‐fluoro‐ and the 7‐chloro‐7‐deaza‐2′‐deoxyxanthosines, 5b and 5c , respectively, were obtained from the corresponding MeO compounds 17b and 17c , or 18 (Scheme 6). The 2′‐deoxyisoguanosine derivative 4b was prepared from 2‐chloro‐7‐fluoro‐7‐deaza‐2′‐deoxyadenosine 6b via a photochemically induced nucleophilic displacement reaction (Scheme 5). The pKa values of the halogenated nucleosides were determined (Table 3). 13C‐NMR Chemical‐shift dependencies of C(7), C(5), and C(8) were related to the electronegativity of the 7‐halogen substituents (Fig. 3). In aqueous solution, 7‐halogenated 2′‐deoxyribonucleosides show an approximately 70% S population (Fig. 2 and Table 1).  相似文献   

2.
The preparation and the pairing properties of the new 3′‐deoxyribopyranose (4′→2′)‐oligonucleotide (=p‐DNA) pairing system, based on 3′‐deoxy‐β‐D ‐ribopyranose nucleosides is presented. D ‐Xylose was efficiently converted to the prefunctionalized 3‐deoxyribopyranose derivative 4‐O‐[(tert‐butyl)dimethylsilyl]‐3‐deoxy‐D ‐ribopyranose 1,2‐diacetate 8 (obtained as a 4 : 1 mixture of α‐ and β‐D ‐anomers; Scheme 1). From this sugar building block, the corresponding, appropriately protected thymine, guanine, 5‐methylcytosine, and purine‐2,6‐diamine nucleoside phosphoramidites 29 – 32 were prepared in a minimal number of steps (Schemes 2–4). These building blocks were assembled on a DNA synthesizer, and the corresponding p‐DNA oligonucleotides were obtained in good yields after a one‐step deprotection under standard conditions, followed by HPLC purification (Scheme 5 and Table 1). Qualitatively, p‐DNA shows the same pairing behavior as p‐RNA, forming antiparallel, exclusively Watson‐Crick‐paired duplexes that are much stronger than corresponding DNA duplexes. Duplex stabilities within the three related (i.e., based on ribopyranose nucleosides) oligonucleotide systems p‐RNA, p‐DNA, and 3′‐O‐Me‐p‐RNA were compared with each other (Table 2). Intrinsically, p‐RNA forms the strongest duplexes, followed by p‐DNA, and 3′‐O‐Me‐p‐RNA. However, by introducing the nucleobases purine‐2,6‐diamine (D) and 5‐methylcytosine (M) instead of adenine and cytosine, a substantial increase in stability of corresponding p‐DNA duplexes was observed.  相似文献   

3.
The base‐pairing properties of oligonucleotides containing the anomeric 5‐aza‐7‐deazaguanine 2′‐deoxyribonucleosides 1 and 5 are described. The oligonucleotides were prepared by solid‐phase synthesis, employing phosphoramidite or phosphonate chemistry. Stable `purine'⋅purine duplexes with antiparallel (aps) chain orientation are formed, when the α‐D ‐anomer 5 alternates with the β‐D ‐anomeric 2′‐deoxyguanosine ( 2 ) within the same oligonucleotide chain. Parallel (ps) oligonucleotide duplexes are observed, when the β‐D anomer 1 alternates with 2 . A renewed reversal of the chain orientation (ps→aps) occurs when compound 1 pairs with 2′‐deoxyisoguanosine ( 6 ). In all cases, it is unnecessary to change the orientation within a single strand when α‐D units alternate with their β‐D counterparts. Heterochiral base pairs of 5 (α‐D ) with 2′‐deoxyisoguanosine (β‐D ) are well accommodated in duplexes with random base composition and parallel chain orientation. Base pairs of 5 (α‐D ) with 2′‐deoxyguanosine (β‐D ) destabilize duplexes with antiparallel chains.  相似文献   

4.
A series of oligonucleotides containing (5′S)‐5′‐C‐butyl‐ and (5′S)‐5′‐C‐isopentyl‐substituted 2′‐deoxyribonucleosides were designed, prepared, and characterized with the intention to explore alkyl‐zipper formation between opposing alkyl chains across the minor groove of oligonucleotide duplexes as a means to modulate DNA‐duplex stability. From four possible arrangements of the alkyl groups that differ in the density of packing of the alkyl chains across the minor groove, three (duplex types I – III , Fig. 2) could experimentally be realized and their duplex‐forming properties analyzed by UV‐melting curves, CD spectroscopy, and isothermal titration calorimetry (ITC), as well as by molecular modeling. The results show that all arrangements of alkyl residues within the minor groove of DNA are thermally destabilizing by 1.5–3°/modification in Tm. We found that, within the proposed duplexes with more loosely packed alkyl groups (type‐ III duplexes), accommodation of alkyl residues without extended distorsion of the helical parameters of B‐DNA is possible but does not lead to higher thermodynamic stability. The more densely packed and more unevenly distributed arrangement (type‐ II duplexes) seems to suffer from ecliptic positioning of opposite alkyl groups, which might account for a systematic negative contribution to stability due to steric interactions. The decreased stability in the type‐ III duplexes described here may be due either to missing hydrophobic interactions of the alkyl groups (not bulky enough to make close contacts), or to an overcompensation of favorable alkyl‐zipper formation presumably by loss of structured H2O in the minor groove.  相似文献   

5.
A standard DNA sequencer was used as a novel and highly efficient tool to study the template‐controlled polymerization of RNA. When labeled with appropriate fluorescent dyes, primers and their extension products could be separated and quantified with excellent sensitivity, reproducibility, and speed. The new technique was applied to compare the template‐controlled incorporation of adenosine mononucleotide 2 and its purine‐2,6‐diamine analogue 3 , the latter being capable of forming three H‐bonds with thymidine or uridine residues. The rates and yields of incorporation are similar when only one thymidine unit is available for pairing in the template (see template 6 and Table 2). However, on template 7 with two consecutive thymidine residues, purine‐2,6‐diamine is clearly ahead of adenine (see Table 3). This advantage is most pronounced when the template contains stretches of three and four thymidine moieties (see templates 8 and 9 and Tables 4 and 5, resp.).  相似文献   

6.
The unusually N8‐glycosylated pyrazolo[3,4‐d]pyrimidine‐4,6‐diamine 2′‐deoxyribonucleoside ( 3 ) was synthesized and converted to the phosphoramidite 11 . Oligonucleotides were prepared by solid‐phase synthesis, and the base pairing of compound 3 was studied. In non‐self‐complementary duplexes containing compound 3 located opposite to the four canonical DNA constituents, strong base pairs are formed that show ambiguous pairing properties. The self‐complementary duplex d( 3 ‐T)6 ( 34 ⋅ 34 ) is significantly more stable than d(A‐T)6.  相似文献   

7.
Oligonucleotides containing the 5‐substituted 2′‐deoxyuridines 1b or 1d bearing side chains with terminal C?C bonds are described, and their duplex stability is compared with oligonucleotides containing the 5‐alkynyl compounds 1a or 1c with only one nonterminal C?C bond in the side chain. For this, 5‐iodo‐2′‐deoxyuridine ( 3 ) and diynes or alkynes were employed as starting materials in the Sonogashira cross‐coupling reaction (Scheme 1). Phosphoramidites 2b – d were prepared (Scheme 3) and used as building blocks in solid‐phase synthesis. Tm Measurements demonstrated that DNA duplexes containing the octa‐1,7‐diynyl side chain or a diprop‐2‐ynyl ether residue, i.e., containing 1b or 1d , are more stable than those containing only one triple bond, i.e., 1a or 1c (Table 3). The diyne‐modified nucleosides were employed in further functionalization reactions by using the protocol of the CuI‐catalyzed Huisgen–Meldal–Sharpless [2+3] cycloaddition (‘click chemistry’) (Scheme 2). An aliphatic azide, i. e., 3′‐azido‐3′‐deoxythymidine (AZT; 4 ), as well as the aromatic azido compound 5 were linked to the terminal alkyne group resulting in 1H‐1,2,3‐triazole‐modified derivatives 6 and 7 , respectively (Scheme 2), of which 6 forms a stable duplex DNA (Table 3). The Husigen–Meldal–Sharpless cycloaddition was also performed with oligonucleotides (Schemes 4 and 5).  相似文献   

8.
The synthesis of the 7‐halogenated derivatives 1b (7‐bromo) and 1c (7‐iodo) of 7‐deaza‐2′‐deoxyxanthosine ( 1a ) is described. A partial Br→I exchange was observed when the demethylation of 6‐methoxy precursor compound 4b was performed with Me3SiCl/NaI. This reaction is circumvented by the nucleophilic displacement of the MeO group under strong alkaline conditions. The halogenated 7‐deaza‐2′‐deoxyxanthosine derivatives 1b , c show a decreased S‐conformer population of the sugar moiety compared to the nonhalogenated 1a . They are expected to form stronger triplexes when they replace 1a in the 1 ?dA?dT base triplet.  相似文献   

9.
The title compound, C14H16N4O4, adopts the anti conformation at the gly­cosylic bond [χ−117.1 (5)°]. The sugar pucker of the 2′‐deoxy­ribo­furan­osyl moiety is C2′‐endo–C3′‐exo, 2T3 (S‐type). The orientation of the exocyclic C4′—C5′ bond is +sc (gauche). The propynyl group is linear and coplanar with the nucleobase moiety. The structure of the compound is stabilized by several hydrogen bonds (N—H⋯O and O—H⋯O), leading to the formation of a multi‐layered network. The nucleobases, as well as the propynyl groups, are stacked. This stacking might cause the extraordinary stability of DNA duplexes containing this compound.  相似文献   

10.
The novel coumarin‐3‐carboxamides (=2‐oxo‐2H‐1‐benzopyran‐3‐carboxamides) 5a – 5g containing lipophilic spacers were synthesized through the Ugi‐four‐component reaction (Scheme 1). The reactions of aromatic aldehydes 1 , 4,4′‐oxybis[benzenamine] or 4,4′‐methylenebis[benzenamine] as diamine 2 , coumarin‐3‐carboxylic acid (=2‐oxo‐2H‐benzopyran‐3‐carboxylic acid; 3 ), and alkyl isocyanides 4 lead to the desired substituted coumarin‐3‐carboxamides 5a – 5g at room temperature with high bond‐forming efficiency. These novel coumarin derivatives exhibit brilliant fluorescence at 544 nm in CHCl3.  相似文献   

11.
Oligonucleotides containing 7‐deaza‐2′‐deoxyinosine derivatives bearing 7‐halogen substituents or 7‐alkynyl groups were prepared. For this, the phosphoramidites 2b – 2g containing 7‐substituted 7‐deaza‐2′‐deoxyinosine analogues 1b – 1g were synthesized (Scheme 2). Hybridization experiments with modified oligonucleotides demonstrate that all 2′‐deoxyinosine derivatives show ambiguous base pairing, as 2′‐deoxyinosine does. The duplex stability decreases in the order Cd>Ad>Td>Gd when 2b – 2g pair with these canonical nucleosides (Table 6). The self‐complementary duplexes 5′‐d(F7c7I‐C)6, d(Br7c7I‐C)6, and d(I7c7I‐C)6 are more stable than the parent duplex d(c7I‐C)6 (Table 7). An oligonucleotide containing the octa‐1,7‐diyn‐1‐yl derivative 1g , i.e., 27 , was functionalized with the nonfluorescent 3‐azido‐7‐hydroxycoumarin ( 28 ) by the Huisgen–Sharpless–Meldal cycloaddition ‘click’ reaction to afford the highly fluorescent oligonucleotide conjugate 29 (Scheme 3). Consequently, oligonucleotides incorporating the derivative 1g bearing a terminal C?C bond show a number of favorable properties: i) it is possible to activate them by labeling with reporter molecules employing the ‘click’ chemistry. ii) Space demanding residues introduced in the 7‐position of the 7‐deazapurine base does not interfere with duplex structure and stability (Table 8). iii) The ambiguous pairing character of the nucleobase makes them universal probes for numerous applications in oligonucleotide chemistry, molecular biology, and nanobiotechnology.  相似文献   

12.
The automated on‐line synthesis of DNA‐3′‐PNA chimeras 1 – 4 and (2′‐O‐methyl‐RNA)‐3′‐PNA chimeras 5 – 8 is described, in which the 3′‐terminal part of the oligonucleotide is linked to the N‐terminal part of the PNA via N‐(ω‐hydroxyalkyl)‐N‐[(thymin‐1‐yl)acetyl]glycine units (alkyl=Et, Ph, Bu, and pentyl). By means of UV thermal denaturation, the binding affinities of all chimeras were directly compared by determining their Tm values in the duplex with complementary DNA and RNA. All investigated DNA‐3′‐PNA chimeras and (2′‐O‐methyl‐RNA)‐3′‐PNA chimeras form more‐stable duplexes with complementary DNA and RNA than the corresponding unmodified DNA. Interestingly, a N‐(3‐hydroxypropyl)glycine linker resulted in the highest binding affinity for DNA‐3′‐PNA chimeras, whereas the (2′‐O‐methyl‐RNA)‐3′‐PNA chimeras showed optimal binding with the homologous N‐(4‐hydroxybutyl)glycine linker. The duplexes of (2′‐O‐methyl‐RNA)‐3′‐PNA chimeras and RNA were significantly more stable than those containing the corresponding DNA‐3′‐PNA chimeras. Surprisingly, we found that the charged (2′‐O‐methyl‐RNA)‐3′‐PNA chimera with a N‐(4‐hydroxybutyl)glycine‐based unit at the junction to the PNA part shows the same binding affinity to RNA as uncharged PNA. Potential applications of (2′‐O‐methyl‐RNA)‐3′‐PNA chimeras include their use as antisense agents acting by a RNase‐independent mechanism of action, a prerequisite for antisense‐oligonucleotide‐mediated correction of aberrant splicing of pre‐mRNA.  相似文献   

13.
Oligonucleotides containing an alkylene intrastrand cross‐link (IaCL) between the O6‐atoms of two consecutive 2′‐deoxyguanosines (dG) were prepared by solid‐phase synthesis. UV thermal denaturation studies of duplexes containing butylene and heptylene IaCL revealed a 20 °C reduction in stability compared to the unmodified duplexes. Circular dichroism profiles of these IaCL DNA duplexes exhibited signatures consistent with B‐form DNA. Human O6‐alkylguanine DNA alkyltransferase (hAGT) was capable of repairing both IaCL containing duplexes with slightly greater efficiency towards the heptylene analog. Interestingly, repair efficiencies of hAGT towards these IaCL were lower compared to O6‐alkylene linked IaCL lacking the 5′‐3′‐phosphodiester linkage between the connected 2′‐deoxyguanosine residues. These results demonstrate that the proficiency of hAGT activity towards IaCL at the O6‐atom of dG is influenced by the backbone phosphodiester linkage between the cross‐linked residues.  相似文献   

14.
The structure and stability of a 14‐mer DNA duplex containing a nucleotide analog with a hydroxymethyl substituent at the C(8) of 2′‐deoxyadenosine has been investigated by molecular‐dynamics simulation. The DNA duplex studied has the sequence 5′‐d(CGTAAGCTCGATAG)‐3′⋅5′‐d(CTATCGA*GCTTACG)‐3′, where the O(3′) of the dG6 nucleotide in the second strand is linked through a phosphinato group with the O(10) of the dA 2′‐deoxyadenosine‐derived nucleotide. Previous experimental results showed that the stability of this duplex in aqueous solution of 0.1M NaCl at pH 7 and room temperature is significantly lower than that of the corresponding unmodified DNA duplex. Comparison of molecular‐dynamics trajectories of the unmodified and modified B‐DNA duplexes in aqueous solution, at similar conditions than the experiment, shows that the substitution of the dA nucleotide by the dA* nucleotide in the second strand induces stretching of the double helix, which results in opening of the grooves and consequent exposure of the double‐helix core to the solvent.  相似文献   

15.
By automated synthesis, we prepared hybrid oligonucleotides consisting of covalently linked RNA and p‐DNA sequences (p‐DNA=3′‐deoxyribopyranose (4′→2′)‐oligonucleotides) (see Table 1). The pairing properties of corresponding hybrid duplexes, formed from fully complementary single strands were investigated. An uninterrupted ππ‐stacking at the p‐DNA/RNA interface and cooperative pairing between the two systems was achieved by connecting them via a 4′‐p‐DNA‐2′→5′‐RNA‐3′ and 5′‐RNA‐2′→4′‐p‐DNA‐2′ phosphodiester linkage, respectively (see Fig. 4). The RNA 2′‐phosphoramidites 9 – 12 , required for the formation of the RNA‐2′→4′‐p‐DNA phosphodiester linkage were synthesized from the corresponding, 3′‐O‐tom‐protected ribonucleosides (tom=[(triisopropylsilyl)oxy]methyl; Scheme 1). Analogues of the flavin mononucleotide (=FMN) binding aptamer 22 and the hammerhead ribozyme 25 were prepared. Each of these analogues consisted of two p‐DNA/RNA hybrid single strands with complementary p‐DNA sequences, designed to substitute stem/loop and stem motifs within the parent compounds. By comparative binding and cleavage studies, it was found that mixing of the two complementary p‐DNA/RNA hybrid sequences resulted in the formation of the fully functional analogues 23 ⋅ 24 and 27 ⋅ 28 of the FMN‐binding aptamer and of the hammerhead ribozyme, respectively.  相似文献   

16.
Oligonucleotides incorporating 7‐deaza‐2′‐deoxyxanthosine ( 3 ) and 2′‐deoxyxanthosine ( 1 ) were prepared by solid‐phase synthesis using the phosphoramidites 6 – 9 and 16 which were protected with allyl, diphenylcarbamoyl, or 2‐(4‐nitrophenyl)ethyl groups. Among the various groups, only the 2‐(4‐nitrophenyl)ethyl group was applicable to 7‐deazaxanthine protection being removed with 1,8‐diazabicyclo[5.4.0]undec‐7‐ene (DBU) by β‐elimination, while the deprotection of the allyl residue with Pd0 catalyst or the diphenylcarbamoyl group with ammonia failed. Contrarily, the allyl group was found to be an excellent protecting group for 2′‐deoxyxanthosine ( 1 ). The base pairing of nucleoside 3 with the four canonical DNA constituents as well as with 3‐bromo‐1‐(2‐deoxy‐β‐D ‐erythro‐pentofuranosyl)‐1H‐pyrazolo[3,4‐d]pyrimidine‐4,6‐diamine ( 4 ) within the 12‐mer duplexes was studied, showing that 7‐deaza‐2′‐deoxyxanthosine ( 3 ) has the same universal base‐pairing properties as 2′‐deoxyxanthosine ( 1 ). Contrary to the latter, it is extremely stable at the N‐glycosylic bond, while compound 1 is easily hydrolyzed under slightly acidic conditions. Due to the pKa values 5.7 ( 1 ) and 6.7 ( 3 ), both compounds form monoanions under neutral conditions (95% for 1 ; 65% for 3 ). Although both compounds form monoanions at pH 7.0, pH‐dependent Tm measurements showed that the base‐pair stability of 7‐deaza‐2′‐deoxyxanthosine ( 3 ) with dT is pH‐independent. This indicates that the 2‐oxo group is not involved in base‐pair formation.  相似文献   

17.
Oligonucleotides with alternating 8-aza-7-deaza-2′-deoxyadenosine (= c7z8Ad2) and dT residues (see 11, 14 and 16 ) or 4-aminopyrazolo [3,4-d] pyrimidine N2-(β-D -2′-deoxyribofuranoside) (= c7z8A′d1); ( 3 ) and dT residues (see 12 ) have been prepared by solid-phase synthesis using P(III) chemistry, Additionally, palindromic oligomers derived from d(C-T-G-G-A-T-C-C-A-G) but containing 2 or 3 instead of dA (see 18 – 22 ) have been synthesized. Benzoylation of 2 or 3 , followed by 4,4′-dimethoxytritylation and subsequent phosphitylation yielded the methyl or the cyanoethyl phosphoramidites 8a,b and 9 . They were employed in automated. DNA synthesis. Alternating oligomers containing 2 or 3 showed increase dTm values compared to those with dA, in particular 12 with an unusual N2-glycosylic bond. The palindromic oligomers 18 - 22 containing 2 or 3 instead of dA outside of the enzymic recognition side reduced the hydrolysis rate, replacement within d(G-A-T-C) abolished phosphodiester hydrolysis.  相似文献   

18.
New complexes [(η6p‐cymene)Ru(C5H4N‐2‐CH=N–Ar)X]PF6 [X = Br ( 1 ), I ( 2 ); Ar = 4‐fluorophenyl ( a ), 4‐chlorophenyl ( b ), 4‐bromophenyl ( c ), 4‐iodophenyl ( d ), 2,5‐dichlorophenyl ( e )] were prepared, as well as 3a – 3e (X = Cl) and the new complexes [(η6‐arene)RuCl(N‐N)]PF6 (arene = C6H5OCH2CH2OH, N‐N = 2,2′‐bipyridine ( 4 ), 2,6‐(dimethylphenyl)‐pyridin‐2‐yl‐methylene amine ( 5 ), 2,6‐(diisopropylphenyl)‐pyridin‐2‐yl‐methylene amine ( 6 ); arene = p‐cymene, N‐N = 4‐(aminophenyl)‐pyridin‐2‐yl‐methylene amine ( 7 )]. X‐ray diffraction studies were performed for 1a , 1b , 1c , 1d , 2b , 5 , and 7 . Cytotoxicities of 1a – 1d and 2 were established versus human cancer cells epithelial colorectal adenocarcinoma (Caco‐2) (IC50: 35.8–631.0 μM), breast adenocarcinoma (MCF7) (IC50: 36.3–128.8.0 μM), and hepatocellular carcinoma (HepG2) (IC50: 60.6–439.8 μM), 3a – 3e were tested against HepG2 and Caco‐2, and 4 – 7 were tested against Caco‐2. 1 – 7 were tested against non‐cancerous human epithelial kidney cells. 1 and 2 were more selective towards tumor cells than the anticancer drug 5‐fluorouracil (5‐FU), but 3a – 3e (X = Cl) were not selective. 1 and 2 had good activity against MCF7, some with lower IC50 than 5‐FU. Complexes with X = Br or I had moderate activity against Caco‐2 and HepG2, but those with Cl were inactive. Antibacterial activities of 1a , 2b , 3a , and 7 were tested against antibacterial susceptible and resistant Gram‐negative and ‐positive bacteria. 1a , 2b , and 3a showed activity against methicillin‐resistant S. aureus (MIC = 31–2000 μg · mL–1).  相似文献   

19.
Selected 2,6‐(disubstituted)purine 2′,3′‐didehydro‐2′,3′‐dideoxynucleosides and 2′,3′‐dideoxynucleosides were prepared and evaluated. Treatment of 5′‐protected ribonucleosides with phenoxythiocarbonyl chloride and 4‐(dimethylamino)pyridine, or under Schotten‐Baumann conditions, gave high yields of 2′,3′‐O‐thiono‐carbonates that underwent Corey‐Winter elimination. Treatment of unprotected ribonucleosides with α‐ace‐toxyisobutyryl bromide in “moist” acetonitrile gave trans 2′,3′‐bromohydrin acetate mixtures that underwent reductive elimination with zinc‐copper couple or zinc/acetic acid. Catalytic hydrogenation of the resulting 2′,3′‐enes gave 2′,3′‐dideoxynucleosides. Treatment of the 2‐amino‐6‐chloropurine and 6‐amino‐2‐fluoro‐purine derivatives with nucleophiles gave 2,6‐(disubstituted)purine 2′,3′‐dideoxynucleosides. 2′,3′‐Dideoxyguanosine and the 2‐amino‐6‐[amino ( 16d ), methoxy ( 16b ), ethoxy ( 16c ), and methylamino ( 16j )]purine 2′,3′‐dideoxynucleosides showed good anti‐hepatitis B activity with infected primary duck hepatocytes. Cytotoxic effects with selected analogues were evaluated in human T‐lymphoblastic and promyelocytic leukemia cell lines. The 2‐amino‐6‐fluoro derivative 16m was the most cytotoxic of the 2‐amino‐6‐(substituted)purine 2′,3′‐dideoxynucleosides, and 2‐fluoro‐2′,3′‐dideoxyadenosine ( 21a ) was the most cytotoxic compound. The order of efficiency of hydrolysis of the 6‐substituent from 2‐amino‐6‐(sub‐stituted)purine 2′,3′‐dideoxynucleosides (Vmax/Km) with adenosine deaminase from calf intestine was: 2‐amino‐6‐[amino ( 16d ) > methoxy ( 16b ) > ethoxy ( 16c )], all of which were ≤3% of the efficiency with adenosine. The 6‐methylamino derivative 16j , as well as 16b , 16c , and 16d were readily converted into 2′,3′‐dideoxyguanosine by duck cell supernatants.  相似文献   

20.
Three 2′‐phenanthrenyl‐C‐deoxyribonucleosides with donor (phenNH2), acceptor (phenNO2), or no (phenH) substitution on the phenanthrenyl core were synthesized and incorporated into oligodeoxyribonucleotides. Duplexes containing either one or three consecutive phenR residues, which were located opposite each other, were formed. Within these residues, the phenR residues are expected to recognize each other through interstrand stacking interactions, in much the same way as described previously for biphenyl DNA. The thermal, thermodynamic, and fluorescence properties of such duplexes were determined by UV melting analysis and fluorescence spectroscopy. Depending on the nature of the substituent, the thermal stability of single‐modified duplexes can vary between ?2.7 to +11.3 °C in Tm and that of triple‐modified duplexes from +7.8 to +11.1 °C. Van′t Hoff analysis suggested that the observed higher thermodynamic stability in phenH‐ and phenNO2‐containing duplexes is of enthalpic origin. A single phenH or phenNO2 residue in a bulge position also stabilizes a corresponding duplex. If a phenNO2 residue is placed in a bulge position next to a base mismatch this can lead, in a sequence‐dependent manner, to duplex destabilization. The phenNO2 residue was found to be a highly efficient (10–100‐fold) quencher of phenH and phenNH2 fluorescence if placed in the opposite position to the fluorophores. When phenH and phenNH2 residues were placed opposite each other, efficient quenching of phenH and enhancement of phenNH2 fluorescence was found, which is an indicator for electron‐ or energy‐transfer processes between the aromatic units.  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