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 156 毫秒
1.
Three title compounds 4a—4c have been synthesized by the cyclodehydration of 1’-benzylidine-4’-(3β-substituted-5α-cholestane-6-yl)thiosemicarbazones 2a—2c with thioglycolic acid followed by the treatment with cold conc. H2SO4 in dioxane. The compounds 2a—2c were prepared by condensation of 3β-substituted-5α-cholestan- 6-one-thiosemicarbazones 1a—1c with benzaldehyde. These thiosemicarbazones 1a—1c were obtained by the reaction of corresponding 3β-substituted-5α-cholestan-6-ones with thiosemicarbazide in the presence of few drops of conc. HCl in methanol. The structures of the products have been established on the basis of their elemental, analytical and spectral data.  相似文献   

2.
Five novel pyrazole‐coupled glucosides, 1,5‐diaryl‐1H‐pyrazol‐3‐yl 2,3,4,6‐tetra‐O‐acetyl‐β‐D ‐glucopyranosides 5a – 5e , were synthesized by the phase‐transfer catalytic reaction of 1,5‐diaryl‐1H‐pyrazol‐3‐ols 4a – 4e with acetobromo‐α‐D ‐glucose in H2O/CHCl3 under alkaline conditions, using Bu4N+Br? as catalyst. Then, glucosides 5a – 5c were deacetylated in a solution of Na2CO3/MeOH to yield the 1,5‐diaryl‐3‐(β‐D ‐glucopyranosyloxy)‐1H‐pyrazoles 6a – 6c . Their structures were characterized by 1H,1H‐COSY, 1H‐, 13C‐, and 19F‐NMR spectroscopy, as well as elemental analysis. The structures of 5d and 6c were also determined by single‐crystal X‐ray diffraction analysis. A preliminary in vitro bioassay indicated that compounds 4e and 5d exhibited excellent‐to‐medium fungicidal activity against Sclerotinia sclerotiorum at the dosage of 10 μg/ml.  相似文献   

3.
The stereoselectivity of the Diels-Alder reaction of (E)-γ-oxo-α,β-unsaturated thioesters 3a-3d with cyclopentadiene is greatly enhanced in the presence of Lewis acids favoring the endo acyl isomers 4a-4d . In the absence of Lewis acid, Diels-Alder reaction of 3a-3d with cyclopentadiene at 25 °C gave two adducts 4a-4d and 5a-5d in a ratio of 1:1 respectively. In the presence of Lewis acids, Diels-Alder reaction of 3a-3d with cyclopentadiene gave 4a-4d and 5a-5d in ratios of 75-94:25-6 respectively. The stereoelectivity was enhanced to ratios of 95-98:5-2 with lowering the reaction temperature. The stereochemistry of the cycloadducts 4 and 5 was confirmed by iodocyclization. Reaction of the endo-thioester 5c with I2 in aqueous THF at 0 °C gave the novel methylthio group rearranged product 6c in 80% yield, the first example of iodo-lactonization of endo-thioesters. Reaction of the endo-acyl isomer 4b with I2 under the same reaction conditions gave an isomeric mixture of 7b and 8b in 1:2 ratio. The stereochemistry of the thioester group in 8b was proved by X-ray single-crystal analysis. The solvent effect on the endo selectivity of (Z)-γ-oxo-α,β-unsaturated thioester 2b was also examined.  相似文献   

4.
New Bi(cyclopropylidenes) by CuCl2-Induced ‘Carbene Dimerization’ of 1-Bromo-1-lithiocycIopropanes A series of so far unknown bi(cyclopropylidenes) 5 are prepared in a simple one-pot reaction by halogenolithio exchange between 1,1-dibromocyclopropanes 1a – c as well as 1e – i and BuLi at ?95°, to give 1-bromo-1-lithiocyclopropanes 2a – c as well as 2e – i , followed by treatment with CuCl2 at low temperature and a simple workup at room temperature (Table 1). The influence of reaction parameters on yields of 5 (Tables 2, 4, and 5) and diastereoselectivity of the reaction 2 → 5 (Table 3) are discussed. In view of an elucidation of the reaction mechanism, first kinetic experiments of the quantitative reaction 1c → 2c → 5c are reported.  相似文献   

5.
Using the readily accessible chiral auxiliaries 1 – 3 the sulfonamide-shielded O-silylated esters 5 underwent π-face-selective α-acetoxylation on successive treatment with Pb(OAc)4 and NEt3 HF to give after recrystallization α-acetoxy ester 6 in 55–67% yields and in 95–100% d.e. Starting from conjugated enoates addition of RCu and subsequent acetoxylation 10 → 11 → 12 yielded α,β-bifunctionalized esters 12 with >95% configurational control at both Cα and Cβ. Nondestructive removal of the auxiliary ( 6 → 7 , 6 → 8 and 12 → 13 ) gave either α-hydroxycarboxylic acids or terminal α-glycols in high enantiomeric purity. The prepared glycols 8c and 13a are key intermediates for previously reported syntheses of the natural products 16 and 17 , respectively.  相似文献   

6.
Sodium perfluoroalkanesulfinate, RFSO2Na [RF?Cl(CF2)4, 1a; CF3(CF2)5, 1b; Cl(CF3)6, 1c] reacted with bromine in aqueous solution to give the corresponding sulfonyl bromide RFSO2Br (2a-2c) and in acetonitrile or acetic acid, to form perfluoroalkyl bromide RFBr (3a-3c). Heating in acetonitrile at 80°C, 2a-2c were converted smoothly into 3a-3c. However, reaction of sodium α,α-dichloropolyfluoroalkanesulfinate RCCl2SO2Na (R?CF3, Cl(CF2)n, n=2, 4, 6, 5a-5d) with bromine in aqueous solution gave directly the corresponding bromoalkanes 1-bromo-1,1-dichloropolyfluoroalkane RCCl2Br (6a-6d). In aqueous potassium iodide solution, 1a-1c, 5a and 5b also reacted with iodine to form the corresponding iodo-polyfluoroalkane 4a-4c, 7a and 7b directly. 6a and 7a underwent free radical addition to alkene readily in the presence of free radical initiator and reacted with Na2S2O4 in the usual way to form α,α-dichloropolyfluoroethane sulfinate (5a). 5a was stable in strong acid, but reacted with strong base to yield 10. 5a was oxidised by hydrogen peroxide to the sulfonate 11 and reduced by zinc in dilute acid to from the α-chloro sulfinate 12.  相似文献   

7.
Synthesis of 4(5)-Acyl-5(4)-alkylimidazoles from Symmetrical 1,3-Diones A new synthesis of 4(5)-acyl-5(4)-alkylimidazoles 1 is described. The symmetrical 1,3-diones 5a and 5b were reacted with N2O4 to give the nitro compounds 7a and 7b , respectively; 5c was treated with NaNO2 to give the nitroso compound 7c (Scheme 2). Hydrogenation of 7a , 7b and 7c over Pd/C in acetic acid/acetic formic anhydride yielded the formamides 9a , 9b and 9c , whose cyclization in formamide/formic acid afforded the 4(5)-acyl-5(4)-alkylimidazoles 1a, 1b and 1c , respectively. Oxazoles 11a and 11b were obtained from the corresponding formamides 9a and 9b with methanesulfonic acid/P2O5.  相似文献   

8.
Oxidation of 2H, 6H-thiin-3-ones 1a – c with 3-chloroperbenzoic acid affords the corresponding 1-oxides 2a – c . On irradiation (350 nm) in either benzene or MeCN, these cyclic sulfoxides 2 isomerize to 3H, 7H-1,2-oxathiepin-4-ones 3 . The tetramethyl derivative 3a is isolated by flash chromatography at ?10°, but, at higher temperatures, it undergoes ring contraction and H2O elimination to give 4,4-dimethyl-2(2-methylprop-2enylidene)thietan-3-one ( 4 ). Diemthyloxathiepinones 3b and 3c undergo ring contraction in MeOH to afford 1-(4-methylthiophen-2-yl)ethanone ( 5 ) and two diastereoisomeric 4,4-dimethyl-2-methoxy-2-(1-methoxyethyl)thietan-3-ones ( 6 and 7 , respectively).  相似文献   

9.
In a preceding communication [5] it was shown that 1, 5-dimethyl-6-methylene-tricyclo[3.2.1.02,7]oct-3-en-8-one ( 2 ) and related tricyclic ketones are converted by strong acids (CF3COOH, FSO3H) into polymethylated tropylium salts with loss of carbon monoxide, e.g. the 1, 2, 4-trimethyltropylium ion 4 from 2 (Scheme 1). Under the influence of neat formic acid at 20°, 2 gives rise to ring-methylated phenylacetic acids, i.e. 2, 4, 5-trimethylphenylacetic acid ( 5 , main product) as well as smaller amounts of 2, 4, 6-and 2, 3, 5-trimethylphenylacetic acids ( 6, 7 resp.; Scheme 2). –On rearrangement of 2 in HCOOD, ca. 2 D-atoms are incorporated (formula d2-5) into the 2, 4, 5-trimethylphenylacetic acid. The tricyclic 15 , containing 3 methyl groups, gives 2, 3, 5, 6-tetramethylphenylacetic acid ( 11 ; Scheme 4) with formic acid; the isomeric tricyclic 16 , 2, 3, 4, 5-tetramethylphenylacetic acid ( 12 ; Scheme 5). From 1, 2, 4, 5-tetramethyl-6-methylene-tricyclo[3.2.1.02,7]oct-3-en-8-one ( 17 ) one obtains pentamethylphenylacetic acid ( 14 ; Scheme 6). Similarly from 18 , a phenylacetic acid derivative, most probably 4-ethyl-2, 5-dimethyl-phenylacetic acid ( 19 ; Scheme 17), has been obtained. –In no case was the formation of α-phenylpropionic acid derivatives observed, not even from the tricyclic 23 containing six methyl groups. From the tricyclic ketone 2 in 70% formic acid a trimethyl-cyclohepta-2, 4, 6-triene-1-carboxyclic acid with partial formula 24 , besides 2, 4, 5-trimethylphenylacetic acid ( 5 ), is formed. 24 remained practically unchanged on standing in neat formic acid and thus does not represent an intermediate product arising by the rearrangement of 2 in that solvent. On standing in methanolic sulfuric acid, tricyclic 2 furnishes the two stereioisomeric methanol-addition products Z- 26 and E- 26 (Scheme 10); these are converted into the phenylacetic acids 5 , 6 and 7 by neat formic acid. The conversion of 2 and related compounds into ring-polymethylated phenylacetic acids, represents a novel and rather complicated reaction. In our opinion the reaction paths represented in Schemes 12 and 18 are responsible for the conversion of 2 into the trimethylphenylacetic acids, compound 40 representing a key intermediate. Analogous reaction paths can be assumed for the other tricyclic ketone transformations. The use of shift reagents in the NMR. spectroscopy and the high-resolution gas-chromatography of the corresponding methyl esters proved particularly important for the analysis of the reaction mixtures. The majority of the polymethylated phenylacetic acids were independently synthesised by means of the Willgerodt-Kindler reaction (chap. 3.2.), whose course is strongly influenced by methyl groups in the ortho-positions of the acetophenone derivatives employed.  相似文献   

10.
On ultraviolet irradiation of the cyclopent-2-enyl methyl ketones 1a – c at ?54° ? t ? 139°, photo-CIDNP. effects of the starting ketones, the 1,3-acetyl shifted isomers (2) , and radical disproportionation and combination products (4 – 7) were observed. These effects show a unique dependence of the polarization phase on temperature which is a novel feature in photo-CIDNP. studies. The results of the investigation, which also included experiments using triplet quenchers, triplet sensitizers and radical scavengers, are rationalized in terms of Schemes 2 and 3. α-Cleavage is a major excited-state reaction of 1a – c on direct irradiation. Temperature-activated α-cleavage (k(t)) to the radical pair R · · R ′1 and intersystem crossing (kisc) to the T2 state are among the competing S1 deactivation processes. The T2 state in turn cleaves (k) to R · · R ′3 A ‘low-temperature range’ with kisc ? k(t) and a ‘high-temperature range’ with k(t) ? kisc exhibiting preferential reactivity from the T2 and S1 states, respectively, can be defined for all three β,γ-unsaturated ketones 1a – c .  相似文献   

11.
Oligodeoxyribonucleotides containing 1-deaza-2′-deoxyadenosine ( = 7-amino-3-(2-deoxy-β-D -erythro-pentofuranosyl)-3H-imidazo[4, 5-b]pyridine; 1b ) form Hoogsteen duplexes. Watson-Crick base pairs cannot be built up due to the absence of N(1). For these studies, oligonucleotide building blocks – the phosphonate 3a and the phosphoramidite 3b – were prepared from 1b via 4a and 5 , as well as the Fractosil-linked 6b , and used in solid-phase synthesis. The applicability of various N-protecting groups (see 4a – c ) was also studied. The Hoogsteen duplex d[(c1A)20] · d(T20) ( 11 · 13 ; Tm 15°) is less stable than d(A20) · d(T20) ( 12 · 13 ; Tm 60°). The block oligomers d([c1A)10–;T10] ( 14 ) and d[T10–(c1A)10] ( 15 ) containing purine and pyrimidine bases in the same strand are also able to form duplexes with each other. The chain polarity was found to be parallel.  相似文献   

12.
Nucleobase-anion glycosylation (KOH, tris[2-(2-methoxyethoxy)ethyl]amine (TDA-1), MeCN) of the pyrrolo[2,3-d]pyrimidines 4a – d with 5-O-[(1,1-dimethylethyl)dimethylsilyl]-2,3-O-(1-methylethylidene)-α-D -ribo-furanosyl chloride ( 5 ) gave the protected β-D -nucleosides 6a – d stereoselectively (Scheme 1). Contrary, the β-D -halogenose 8 yielded the corresponding α-D -nucleosides ( 9a and 9b ) apart from minor amounts of the β-D -anomers. The deprotected nucleosides 10a and 11a were converted into 4-substituted 2-aminopyrrolo[2,3-d]-pyrimidine β-D -ribofuranosides 1 . 10c , 12 , 14 , and 16 and into their α-D -anomers, respectively (Scheme 2). From the reaction of 4b with 5 , the glycosylation product 7 was isolated, containing two nucleobase moieties.  相似文献   

13.
We report herein that the reaction between a series of Hantzsch’s ester analogues 1 a – d with the Lewis acidic species B(C6F5)3 results in facile transfer of hydride to boron. The main products of this reaction are pyridinium borohydride salts 2 a – d , which are obtained in high to moderate yields. The N‐substituted substrates (N‐Me, N‐Ph) reacted in high yield 90–98 % and the connectivity of the products were confirmed by an X‐ray crystallographic analysis of the N‐Me borohydride salt 2 a . Unsubstituted Hanztsch’s ester 1 a reacted less effectively generating only 60 % of the corresponding borohydride salt, with the balance of the material sequestered as the ester‐bound Lewis acid–base adduct 3 a . Formation of the Lewis acid–base adduct could be minimized by increasing the steric bulk about the ester groups as in 1 d . The connectivity of the carbonyl‐bound adduct was confirmed by an X‐ray crystallographic analysis of 3 e the product of the reaction of methyl ketone 1 e with B(C6F5)3. We also explored the generation of these pyridinium salts by employing frustrated Lewis pair methodology. However, the reaction of mixtures of the corresponding pyridine and B(C6F5)3 with hydrogen gas only resulted in formation of trace amounts of the pyridinium borohydride, along with the Lewis acid–base adduct of the starting material and B(C6F5)3. The 1,2‐dihydropyridine adduct was the final product of this reaction. This was ascribed to the low basicity of the pyridine nitrogen and the complicating formation of an ester bound Lewis acid–base adduct.  相似文献   

14.
Photochemical behaviors of the pyrazinone derivatives 5,6,7,8-tetrahydroquinoxalin-2(1H)-ones 1a – c and 1,5,6,7,8,9-hexahydro-2H-cyclohepta[b]pyrazin-2-one 1d were investigated. Dye-sensitized photo-oxygenation of 1a-c gave the 1:1 adducts 5a – c of the corresponding 3,8a-epidioxy-3,5,6,7,8,8a-hexahydroquinoxalin-2(1H)-one 4 and H2O, whereas 1d gave 3,9a-epidioxy-1,3,5,6,7,8,9,9a-octahydro-2H-cyclohepta[b]pyrazin-2-one 4d (Scheme 2). The different kind of products was interpreted as being the result of the ring strain and steric hindrance of endoperoxides produced from 1a – d with singlet oxygen. Irradiation of 1a – b in the presence of alkenes gave tricyclic azetidine derivatives 9 by [2 + 2] cycloaddition of the C?N bond of 1 to the alkene.  相似文献   

15.
Ultraviolet irradiation of the aldehydes 6 – 11 in degassed solutions results exclusively in decarbonylation to the major products 34, 35 and 37 – 40 , and to small amounts of 2, 3-diphenyl-2, 3-dimethyl-butanes 36 from the phenyl aldehydes 6 and 7 . In the presence of tri-n-butylstannane, incorporation of stannane hydrogen competes, to substrate-specific limits, with the intramolecular deuterium transfer in 7 → 35 and 11 → 40 . The quantum yields for decarbonylation are Φ ~ 0.4–1.0 for the phenyl aldehydes 6 and 9 , and 0.02 for 8. Hammett correlations of Φ with resonance constants ( R ) for 6 (X = H, p-CH3, ? OCH3) and (? CF3) and with ωm+ values for the meta-substituted isomers are in agreement with the proposed α-cleavage to an associated radical pair with only moderate free radical character as the primary photochemical step. Φ for 10 (X = H) is 0.11, and for 10 (X = OCH3) 0.065. It is noteworthy that decarbonylation of 10 (X = OCH3) occurs also at 3340 Å (Φ? CO = 0.11) i.e., upon excitation in an absorption band which is presumably lower in energy than the n → π* transition and corresponds to the aromatic Lb transition of 2-methoxynaphthalene. Singlet multiplicity of the reactive excited states is probable on the basis of the fact that the decarbonylation of 6 (X = H) and 10 (X = H and OCH3) could be sensitised neither by acetone nor acetophenone, and could be quenched neither by naphthalene nor by cis-1, 3-pentadiene and nor by 1, 3-cyclohexadiene.  相似文献   

16.
A series of novel 1‐((1H‐1,2,3‐triazol‐4‐yl)methyl)‐2‐(1‐((1H‐1,2,3‐triazol‐4‐yl)methyl)‐5‐substituted‐1H‐indol‐3‐yl)‐6‐substituted‐1H‐benzo[d]imidazoles 5a – i have been prepared using click chemistry as an ideal strategy where [3 + 2] cycloaddition of azides with terminal alkynes has been developed as the target compounds. In route‐II, 5‐substituted‐1H‐indole‐3‐carbaldehydes 1a – c react with 5‐substituted orthophenylenediamine 8 to give desired products, that is, 6‐substituted‐2‐(5‐substituted‐1H‐indol‐3‐yl)‐1H‐benzo[d]imidazole 6a – i . Here, 6a – i react with 2 equiv of propargylbromide 7 to give novel 6‐substituted 2‐(5‐substituted‐1‐(prop‐2‐yn‐1‐yl)‐1H‐indol‐3‐yl)‐1‐(prop‐2‐yn‐1‐yl)‐1H‐benzo[d]imidazole 4a – i . 4a – i were reacted with 2 equiv of NaN3 in t‐butanol/water (1:2) and add catalytic amount of CuSO4.5H2O. Stir the reaction mixture at room temperature to get the target products 5a – i . Here, obtained products contain four rings, that is, one indole, two triazoles, and one benzimidazole. The main advantages of this method are short reaction times, easy workup, higher yields (88–92%), and no by‐products formation.  相似文献   

17.
Condensation of 3,4‐dimethoxybenzeneethanamine ( 3d ) and various benzeneacetic acids, i.e., 4a – e , via a practical and efficient one‐pot Bischler–Napieralski reaction, followed by NaBH4 reduction, produced a series of 1‐benzyl‐1,2,3,4‐tetrahydroisoquinolines, i.e., 5a – e , in satisfactory yields (Scheme 3). Oxidative coupling of the N‐acyl and N‐methyl derivatives 6a – e of the latter with hypervalent iodine ([IPh(CF3COO)2]) yielded products with two different skeletons (Scheme 4). The major products from N‐acyl derivatives 6a – c were (±)‐N‐acylneospirodienones 2a – c , while the minor was the 3,4‐dihydroisoquinoline 7 . (±)‐Glaucine ( 1 ), however, was the major product starting from N‐methyl derivative 6e . Possible reaction mechanisms for the formation of these two types of skeleton are proposed (Scheme 5).  相似文献   

18.
A series of cyclohexane‐1,2‐diamine ( 3a – 3d ) and benzene‐1,2‐diamine derivatives ( 3e – 3h ) were pre‐ pared. Followed by hydrolysis, the reaction of 3a – 3c with PCl3 successfully led to the formation of cor‐ responding metastable saturated heteroatom‐substituted secondary phosphine oxides (HASPO 4a – 4c ), a tautomer of the saturated heteroatom‐substituted phosphinous acid (HAPA). Whereas ambient‐stable diamine‐coordinated palladium complexes were obtained, HAPA‐coordinated palladium complexes were not successfully synthesized. The molecular structures of HASPO 4c , Pd(OAc)2(3a) , PdBr2(3b) and Pd(OAc)2(3c) and [Cu(NO3)(3d)+][NO3 ? ] were determined by single‐crystal X‐ray diffraction method. Catalysis of in‐situ Suzuki‐Miyaura cross‐coupling reactions for aryl bromides and phenylboronic acid using diamine 3a as ancillary ligand showed that the optimized reaction condition at 60 °C is the combination of 2 mmol % 3a /3.0 mmol KOH/1.0 mL 1,4‐dioxane/1 mmol % Pd(OAc)2. Moreover, moderate reactivity was observed when using aryl chlorides as substrates (supporting infor‐ mation). When diamine 3d was employed in Heck reaction, good tolerance of functional groups of aryl bromides were observed while using 4‐bromoanisole and styrene as substrates. The optimized condi‐ tion for Heck reaction at 100 °C is 3 mmol % 3d /3.0 mmol CsF/1.0 mL toluene/3 mmol % Pd(OAc)2. In general, cyclohexane‐1,2‐diamine derivatives exhibited better catalytic properties than those of benzene‐1,2‐diamines.  相似文献   

19.
In a search for inhibitors of platelet aggregation, a number of α-methylidene-γ-butyrolactones 5 and 6 bearing flavone or xanthone moieties, respectively, were synthesized and evaluated for their antiplatelet activity against thrombin(Thr)-, arachidonic-acid(AA)-, collagen(Col)?, and platelet-activating-factor(PAF)-induced aggregation in washed rabbit platelets. These compounds were synthesized from 7-hydroxyflavone ( 1 ) or 3-hydroxyxanthone ( 2 ) via O-alkylation (→ 3 and 4 , resp.) and Reformatsky-type condensation (Scheme). Most of the flavone-containing α-methylidene-γ-butyrolactones 5a – d showed potent antiplatelet effects on AA- and Col-induced aggregation, while xanthone derivatives 6c – e were found to have the same pharmacological profile than aspirin in which only AA-induced aggregation was inhibited (Table 1). However, 6c – e were approximately three to ten times more potent than aspirin (Table 2). For the vasorelaxing effects, 5a was the only compound which exhibited significant inhibitory activity on the high-K+ medium, Ca2+-induced vasoconstriction (Table3). Both 5a and 6a , with an aliphatic Me substituent at C(γ) of the lactone, were active against norepinephrine-induced phasic and tonic constrictions while their γ-aryl-substituted counterparts 5b – f and 6b – f were inactive.  相似文献   

20.
In-situ-generated phthalimidonitrene was added to five α, β-unsaturated esters containing a chiral secondary O-function at C(γ). The additions were fully suprafacial, inasmuch as the (E)-isomers 1 afforded only the trans-aziridines 2 and 3 (J(β, γ) = 4.8?5.1 Hz) and the (Z)-isomers 4 only the cis-aziridines 5 and 6 (8.2?8.5 Hz). The products 2 , 3 , 5 , and 6 where shown to possess the arabino-, xylo-, ribo-, and lyxo- configuration, respectively, by X-ray structure analysis of 2b , 2d , and 6a . The diastereoface selectivity of the nitrene additions, induced by the chiral substructure around C(γ), resulted in more 2 than 3 from 1 , but more 6 than 5 from 4 , which means that the preference of attack at the double bond switches from one side to the other depending on the C=C configuration. The preferences were higher at lower temperature. The aziridines 2a , 2d , and 3d exhibit 1H-NMR-visible isomerism at the ring N-atom; the major (78?95 %)invertomer A is always the one with the phthalimido group in trans-position to the (larger) substructure around C(γ). The other aziridines only show 1H-NMR signals of one invertomer, which – by steric reasoning - ought to be A ; this is confirmed by a 1H-NMR argument for 3a , 5a , 6a , 5c , and 6c .  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