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 15 毫秒
1.
The Rh11-catalyzed carbenoid addition of diazoacetates to olefins was investigated with [Rh2{(4S)-phox}4] ( 1 ;phox = tetrakis[(4S)-tetrahydro-4-phenyloxazol-2-one]), [Rh2{(2S)-mepy}4] ( 2 ; mepy = tetrakis[methyl (2S)-tetrahydro-5-oxopyrrole-2-carboxylate]), and [Rh2(OAc)4] ( 3 ). While catalysis with 2 and 3 afford preferentially trans-cyclopropanecarboxylates, the cis-isomers are the major products with 1 . In general, the enantioselectivities achieved with 1 and 2 are comparable. Additions catalyzed by 1 are strongly sensitive to steric effects. Highly substituted olefins afford cyclopropanes in only poor yield. The preferential cis-selectivity observed in reactions catalyzed by 1 is attributed to dominant interactions between the ligand of the catalyst and the substituents of both olefin and diazoacetate, which overrule the steric interactions between olefin and diazoacetate in the transition state for carbene transfer.  相似文献   

2.
The intermediacy of metallocarbenes in decomposition reactions of iodonium ylides with [Rh2(OAc)4] was established by comparison with reactions of the corresponding diazo compounds. The sensitivity of the RhII-catalyzed intermolecular cyclopropane formation from substituted styrenes and bis(methoxycarbonyl)(phenyliodono)methanide ( 1a ) or dimethyl diazomalonate ( 1b ) is identical. The Hammett plot (with σ+) has a slope of ?0.47. Iodonium ylides and diazo compounds afford the same products in [Rh2(OAc)4]-catalyzed cyclopropane formations, cycloadditions, and intramolecular CH insertions, and exhibit the identical selectivity in intramolecular competitions for cyclopropane formation and insertion. The intramolecular CH insertion of the ylide 20c , when carried out in the presence of a chiral catalyst ([Rh2{(?)-(S)-ptpa}4]), results in formation of 21a having an ee of 67%, identical to the ee obtained with the diazo compound 20b .  相似文献   

3.
Enantioselective intramolecular amidation of aliphatic sulfonamides was achieved for the first time by means of chiral carboxylatodirhodium(II) catalysts in conjunction with PhI(OAc)2 and MgO in high yields and with enantioselectivities of up to 66% (Scheme 3, Table 1). The best results were obtained with [Rh2{(S)‐nttl)4] and [Rh2{(R)‐ntv)4] as catalysts ((S)‐nttl=(αS)‐α‐(tert‐butyl)‐1,3‐dioxo‐2H‐benz[de]isoquinoline‐2‐acetato, (R)‐nto=(αR)‐α‐isopropyl‐1,3‐dioxo‐2H‐benz[de] isoquinoline‐2‐acetato). In addition, these carboxylatodirhodium(II) catalysts were also efficient in intramolecular amidations of aliphatic sulfamates esters, although the enantioselectivity of these latter reactions was significantly lower (Scheme 4, Table 3).  相似文献   

4.
A stereoselective Rh‐catalyzed intermolecular amination of thioethers using a readily available chiral N‐mesyloxycarbamate to produce sulfilimines in excellent yields and diastereomeric ratio is described. A catalytic mixture of 4‐dimethylaminopyridine (DMAP) and bis(DMAP)CH2Cl2 proved pivotal in achieving high selectivity. The X‐ray crystal structure of the (DMAP)2?[Rh2{(S)‐nttl}4] complex was obtained and mechanistic studies suggested a RhII‐RhIII complex as the catalytically active species.  相似文献   

5.
Different classes of cyclopropanes derived from Meldrum's acid (=2,2‐dimethyl‐1,3‐dioxane‐4,6‐dione; 4 ), dimethyl malonate ( 5 ), 2‐diazo‐3‐(silyloxy)but‐3‐enoate 16 , 2‐diazo‐3,3,3‐trifluoropropanoate 18 , diazo(triethylsilyl)acetate 24a , and diazo(dimethylphenylsilyl)acetate 24b were prepared via dirhodium(II)‐catalyzed intermolecular cyclopropanation of a set of olefins 3 (Schemes 1 and 46). The reactions proceeded with either diazo‐free phenyliodonium ylides or diazo compounds affording the desired cyclopropane derivatives in either racemic or enantiomer‐enriched forms. The intramolecular cyclopropanation of allyl diazo(triethylsilyl)acetates 28, 30 , and 33 were carried out in the presence of the chiral dirhodium(II) catalyst [Rh2{(S)‐nttl)4}] ( 9 ) in toluene to afford the corresponding cyclopropane derivatives 29, 31 and 34 with up to 37% ee (Scheme 7). An efficient enantioselective chiral separation method based on enantioselective GC and HPLC was developed. The method provides information about the chemical yields of the cyclopropane derivatives, enantioselectivity, substrate specifity, and catalytic activity of the chiral catalysts used in the inter‐ and intramolecular cyclopropanation reactions and avoids time‐consuming workup procedures.  相似文献   

6.
The two dinuclear IrI complexes [Ir2(μ‐Cl)2 {(R)‐(S)‐PPF‐PPh2}2] ( 1 ; (R)‐(S)‐PPF‐PPh2=(S)‐1‐(diphenylphosphino)‐2‐[(R)‐1‐(diphenylphosphino)ethyl]ferrocene and [Ir2(μ‐Cl)2{(R)‐binap}2] ( 3 ; (R)‐binap=(R)‐[1,1′‐binaphthalene]‐2,2′‐diylbis[diphenylphosphine]) smoothly react with 4 equiv. of the lithium salt of aniline to afford the new bis(anilido)iridate(I) (=bis(benzenaminato)iridate(1‐)) complexes Li[Ir(NHPh)2{(R)‐(S)‐PPF‐PPh2}] ( 4 ) and Li[Ir(NHPh)2{(R)‐binap}] ( 5 ), respectively. The anionic complexes 4 and 5 react upon protonolysis to give the dinuclear aminato‐bridged derivatives [Ir2(μ‐NHPh)2{(R)‐(S)‐PPF‐PPh2}2] ( 6 ) and [Ir2(μ‐NHPh)2{(R)‐binap}2] ( 7 ), which were characterized by X‐ray crystallography. None of the new complexes 4 – 7 shows catalytic activity in the hydroamination of olefins.  相似文献   

7.
For the first time, the stereochemical course of an asymmetric cyclopropanation can be discussed on the basis of experimental structural information on a pertinent chiral dirhodium carbene intermediate. Key to success was the formation of racemic single crystals of a heterochiral [Rh2{(S*)‐PTTL}4{=C(Ar)COOMe}][Rh2{(R*)‐PTTL}4] (Ar=MeOC6H4; PTTL=N‐phthaloyl‐tert‐leucinate) capsule, which has been characterized by X‐ray diffraction. NMR spectroscopic data confirm that the obtained structural portrait is also relevant in solution and provide additional information about the dynamics of this species. The chiral binding pocket is primarily defined by the conformational preferences of the N‐phthaloyl‐protected amino acid ligands and reinforced by a network of weak interligand interactions that get stronger when chlorinated phthalimide residues are used.  相似文献   

8.
Parallel effort : Stereodivergent parallel kinetic resolution of a racemic mixture of dienes using Davies' [Rh2{(S)‐dosp}4] or [Rh2{(R)‐dosp}4] catalysts promotes a tandem vinyl diazoacetate cyclopropanation/Cope rearrangement sequence to afford two diastereomeric, enantioenriched cycloheptadienes, which correspond to the natural antipodes of the title diterpenoids (see scheme).

  相似文献   


9.
Synthesis and Dynamic Behaviour of [Rh2(μ-H)3H2(PiPr3)4]+. Contributions to the Reactivity of the Tetrahydridodirhodium Complex [Rh2H4(PiPr3)4] An improved synthesis of [Rh2H4(PiPr3)4] ( 2 ) from [Rh(η3-C3H5)(PiPr3)2] ( 1 ) or [Rh(η3-CH2C6H5)(PiPr3)2] ( 3 ) and H2 is described. Compound 2 reacts with CO or CH3OH to give trans-[RhH(CO)(PiPr3)2] ( 4 ) and with ethene/acetone to yield a mixture of 4 and trans-[RhCH3(CO)(PiPr3)2] ( 5 ). The carbonyl(methyl) complex 5 has also been prepared from trans-[RhCl(CO)(PiPr3)2] ( 6 ) and CH3MgI. Whereas the reaction of 2 with two parts of CF3CO2H leads to [RhH22-O2CCF3) · (PiPr3)2] ( 8 ), treatment of 2 with one equivalent of CF3CO2H in presence of NH4PF6 gives the dinuclear compound [Rh2H5(PiPr3)4]PF6 ( 9a ). The reactions of 2 with HBF4 and [NO]BF4 afford the complexes [Rh2H5(PiPr3)4]BF4 ( 9b ) and trans-[RhF(NO)(PiPr3)2]BF4 ( 11 ), respectively. In solution, the cation [Rh2(μ-H)3H2(PiPr3)4]+ of the compounds 9a and 9b undergoes an intramolecular rearrangement in which the bridging hydrido and the phosphane ligands are involved.  相似文献   

10.
The cyclopropenation of diethoxypropyne ( 1 ) with methyl diazoacetate in the presence of [Rh2{(2S)-mepy}4] (mepy=methyl 5-oxopyrrolidine-2-carboxylate) proceeds with >95% ee. The resulting cyclopropenecarboxylate 2 underwent stereoselective hydrogenation to the cis-cyclopropane 3 . Hydrolysis of the acetal function of 3 liberated the formyl cyclopropenecarboxylate 4 , which was transformed by Wittig reaction with the phosphonate 5 to afford dehydroamino acid 6 as a mixture of (Z)- and (E)-isomers in various proportions. The (Z)-isomer 6a was hydrolyzed, and the structure and the absolute configuration of the (Z)-dicarboxylic acid 7a were established by X-ray crystallography. The cis-divinylcyclopropane 11 (ee>95%), in turn, was synthesized from 4 via Wittig reaction to afford 8 , which was transformed to the aldehyde 10 and subjected to a second Wittig reaction. Thermolysis of 11 afforded (+)-dictyopterene C′ ( 12 ) in quantitative yield.  相似文献   

11.
Two enantiomerically pure trinuclear compounds of formula (P)-[Mo3S4{(R,R)-Me–BPE}3Br3]Br and (P)-[Mo3Se4{(R,R)-Me–BPE}3Cl3]Cl, (P)-1b.Br and (P)-1c.Cl, respectively, have been synthesized in a good yield and a stereospecific manner by excision of polymeric [Mo3Q7X4]n (Q = S or Se, X = Cl or Br) phases with (R,R)-Me–BPE{1,2-bis[(2R,5R)-2,5-(dimethylphospholan-1-yl)]ethane}. They have been transformed into chiral hetereo cuboidal compounds [Mo3S4{(R,R)-Me–BPE}3Br3]PF6, (P)-2b.PF6, and [Mo3Se4{(R,R)-Me–BPE}3Cl3]PF6, (P)-2c.PF6, by reaction with copper salts. All these compounds have been characterized by 31P NMR, IR, UV–Vis, mass spectrometry, elemental analysis, and chiral dichroism. The catalytic potential of tetranuclear cuboidal compounds has been assessed in the paradigm intermolecular cyclopropanation reaction of styrene with ethyl diazoacetate. Results are compared with those obtained for the analogue [Mo3S4{(R,R)-Me–BPE}3Cl3]PF6, (P)-2a.PF6. The catalytic data demonstrate that the Se derivative (P)-2c.PF6 is less reactive than the S analogues, but it leads to a similar product distribution as the sulfide analogue (P)-2a.PF6. By contrast, exchange of chlorine by the bulky bromine gives rise to a catalyst which makes the carbene dimerization more competitive. These data agree with temporal breaking of one of the Cu–Q bonds to generate an active catalytic species.  相似文献   

12.
《Tetrahedron: Asymmetry》2005,16(21):3484-3487
The dirhodium-catalyzed aziridination of olefins with chiral sulfonimidamides as iminoiodane precursors has been investigated under stoichiometric conditions. Diastereoisomeric excesses of up to 82% have been achieved using [Rh2{(S)-nttl}4] as catalyst. Matching and mismatching effects were observed upon use of chiral rhodium(II) carboxylate catalysts.  相似文献   

13.
A new method for the synthesis of (1R,4S,5S)-4-hydroxymethyl-3-oxabicyclo[3.1.0]hexan2-one, the cyclopropane analog of (S)-5-hydroxypent-2-en-4-olide, has been suggested based on oxidation of (1S,2S,4R,6R)-7,9-dioxatricyclo[4.2.1.02,4]nonan-5-one. Oxidation of cyclobutanones, spirojoined with the fragments of 6,8-dioxabicyclo[3.2.1]oct-2-ene, 6,8-dioxabicyclo[3.2.1]octane (at position 4), or 7,9-dioxatricyclo[4.2.1.02,4]nonane (at position 5), upon the action of m-chloroperoxybenzoic acid or the KMnO4-H2SO4-H2O system leads to the corresponding spirojoined butanolides in 73–85% yields. The same cyclobutanones easily undergo the four-membered ring opening upon the action of dilute H2SO4 at 50–90 °C to form 6,8-dioxabicyclo[3.2.1]octane-4- or 7,9-dioxatricyclo[4.2.1.02,4]nonane-5-propionic acid.  相似文献   

14.
The decomposition of cyclohexyl diazoacetate ( 5a ) in the presence of the chiral [Rh2{(2S)-mepy}4] catalyst leads to a 3:1 cis/trans mixture of bicyclic lactone 6a with an enantiomeric excess of 95–97% (cis) and 90% (trans). The conformationally rigid tert-butyl derivatives 5b and 5c afford, in the presence of the same catalyst, 6b and 6c , respectively, via insertion into the equatorial C? H bonds exclusively, with ee's of ca. 95%. A remarkable degree of induction (92–95%) results in the lactone 6g upon decomposition of 1-isopropyl-2-methylpropyl diazoacetate ( 5g ). The diazoacetates derived from 1-methylcyclohexanol, cyclopentanol and 1-methylcyclopentanol ( 5d–f ) afford under similar conditions insertion products with higher diastereoselectivity, but significantly lower enantioselectivity. Other dirhodium catalysts are less efficient.  相似文献   

15.
To study the conversion from a meso form to a racemic form of tetrahydrofurantetracarboxylic acid (H4L), seven novel coordination polymers were synthesized by the hydrothermal reaction of Zn(NO3)2 ? 6 H2O with (2S,3S,4R,5R)‐H4L in the presence of 1,10‐phenanthroline (phen), 2,2′‐bipyridine (2,2′‐bpy), or 4,4′‐bipyridine (4,4′‐bpy): [Zn2{(2S,3S,4R,5R)‐L}(phen)2(H2O)] ? 2 H2O ( 1 ), [Zn4{(2S,3R,4R,5R)‐L}{(2S,3S,4S,5R)‐L}(phen)2(H2O)2] ( 2 ), [Zn2{(2S,3S,4R,5R)‐L}(H2O)2] ? H2O ( 3 ), [Zn4{(2S,3R,4R,5R)‐L}{(2S,3S,4S,5R)‐L} (2,2′‐bpy)2(H2O)2] ? 2 H2O ( 4 ), [Zn2 {(2S,3S,4R,5R)‐L}(2,2′‐bpy)(H2O)] ( 5 ), [Zn4{(2S,3R,4R,5R)‐L}{(2S,3S,4S,5R)‐L} (4,4′‐bpy)2(H2O)2] ( 6 ), and [Zn2 {(2S,3S,4R,5R)‐L}(4,4′‐bpy)(H2O)] ? 2 H2O ( 7 ). These complexes were obtained by control of the pH values of reaction mixtures, with an initial of pH 2.0 for 1 , 2.5 for 2 , 4 , and 6 , and 4.5 for 3 , 5 , and 7 , respectively. The expected configuration conversion has been successfully realized during the formation of 2 , 4 , and 6 , and the enantiomers of L, (2S,3R,4R,5R)‐L and (2S,3S,4S,5R)‐L, are trapped in them, whereas L ligands in the other four complexes retain the original meso form, which indicates that such a conversion is possibly pH controlled. Acid‐catalyzed enol–keto tautomerism has been introduced to explain the mechanism of this conversion. Complex 1 features a simple 1D metal–L chain that is extended into a 3D supramolecular structure by π–π packing interactions between phen ligands and hydrogen bonds. Complex 2 has 2D racemic layers that consist of centrosymmetric bimetallic units, and a final 3D supramolecular framework is formed by the interlinking of these layers through π–π packing interactions of phen. Complex 3 is a 3D metal–organic framework (MOF) involving meso‐L ligands, which can be regarded as (4,6)‐connected nets with vertex symbol (45.6)(47.68). Complexes 4 and 5 contain 2D racemic layers and (6,3)‐honeycomb layers, respectively, both of which are combined into 3D supramolecular structures through π–π packing interactions of 2,2′‐bpy. The structure of complex 6 is a 2D network formed by 4,4′‐bpy bridging 1D tubes, which consist of metal atoms and enantiomers of L. These layers are connected through hydrogen bonds to give the final 3D porous supramolecular framework of 6 . Complex 7 is a 3D MOF with novel (3,4,5)‐connected (63)(42.64)(42.66.82) topology. The thermal stability of these compounds was also investigated.  相似文献   

16.
Monocyclometalated compound [Rh2{(C8H4S)P(C8H5S)2}(CH3CO2H)2(O2CCH3)3] ( 1 a ) and bis‐cyclometalated compound [Rh2{(C8H4S)P(C8H5S)2}2(CH3CO2H)2(O2CCH3)2] ( 2 a ) have been isolated from the reaction of dirhodium tetraacetate and tris(2‐benzo[b]thienyl)phosphine ( 2 BTP ) using low acidic solutions. By contrast, in pure acetic acid the reaction of Rh2(O2CCH3)4 with 2 BTP and tris(2‐thienyl)phosphine ( 2 TP ), followed by replacement of the axial acetate ligands by chlorides, led to [Rh2{(2‐C8H5 S )P(2‐C8H5S)2}2Cl2(O2CCH3)2] ( 3 b ) and [Rh2{(2‐C4H3 S )P(C4H3S)2}2Cl2(O2CCH3)2] ( 5 b ), respectively. These new dirhodium(II) compounds possess equatorial bridging ligands in a phosphorous–sulfur (P,S) coordination mode. The reversible switching between the P,C and P,S bonding mode of the phosphine has been studied in the monocyclometalated [Rh2{(C4H2S)P(C4H3S)2}(CH3CO2H)2(O2CCH3)3] ( 6 a ), which was selectively transformed into compound [Rh2{(2‐C4H3 S )P(C4H3S)2}(CF3SO3)(CH3CO2H)(O2CCH3)3] ( 7 c ) in triflic acid media. Remarkably, compound 7 c reverts to the starting compound 6 a upon treatment with sodium acetate. Theoretical DFT calculations for both the P,C/P,S rearrangement and the base‐promoted reversion have been performed to explain the experimental findings. Data suggest the P,C/P,S rearrangement occurs by means of a “concerted protonation–demetalation mechanism” followed by η2 coordination of the thienyl ring and subsequent isomerization to the S‐η1‐coordination mode. In the reversion reaction, the base coordinated at the axial position would promote a concerted metalation–deprotonation mechanism.  相似文献   

17.
Photolysis of [Me2SiSiMe2)[C5H4Fe(CO2)]2with a series of bis(phosphine)ligands Ph2P(CH2)n PPh2(n=1-4) leads to the formation of the corresponding diiron complexes with intramolecular and intermolecular bis(phosphine) substitution.When these complexes were heated in refluxing xylene.only in the complexes with intermolecular bis(phosphine )substitution the thermal rearrangement reaction occurred.  相似文献   

18.
The molecular structure of the title tricyclic compound, C17H21NO4, which is the immediate precursor of a potent synthetic inhibitor {Lek157: sodium (8S,9R)‐10‐[(E)‐ethyl­idene]‐4‐methoxy‐11‐oxo‐1‐aza­tri­cyclo­[7.2.0.03,8]­undec‐2‐ene‐2‐carboxyl­ate} with remarkable potency, provides experimental evidence for the previously modelled relative position of the fused cyclo­hexyl ring and the carbonyl group of the β‐lactam ring, which takes part in the formation of the initial tetrahedral acyl–enzyme complex. In this hydro­phobic mol­ecule, the overall geometry is influenced by C—H?O intramolecular hydrogen bonds [3.046 (4) and 3.538 (6) Å, with corresponding normalized H?O distances of 2.30 and 2.46 Å], whereas the mol­ecules are interconnected through intermolecular C—H?O hydrogen bonds [3.335 (4)–3.575 (5) Å].  相似文献   

19.
Synthesis and Crystal Structures of (PPh4)2[TeS3] · 2 CH3CN and (PPh4)2[Te(S5)2] (PPh4)2[TeS3] · 2 CH3CN was obtained by the reaction of PPh4Cl, Na2S4 and Te in acetonitrile. With sulfur it reacts yielding (PPh4)2[Te(S5)2]. The crystal structures of both products were determined by X-ray diffraction. (PPh4)2[TeS3] · 2 CH3CN: triclinic, space group P1 , Z = 2, R = 0.041 for 4 629 reflexions; it contains trigonal-pyramidal [TeS3]2? ions with an average Te? S bond length of 233 pm. (PPh3)2[Te(S5)2]: monoclinic, P21/n, Z = 2, R = 0.037 for 2 341 reflexions. In the [Te(S5)2]2? ion the tellurium atom has a nearly square coordination by four S atoms. Along with the Te atoms each of the two S5 groups forms a ring with chair conformation.  相似文献   

20.
Zinc enolates derived from 1-R-2,2-dibromoethanone reacted with 3-(2-oxo-4a,8a-dihydro-2H-chromene-3-carbonyl)chromen-2-one to give the corresponding 1-R-1a-{(1-R-2-oxo-1,7b-dihydrocyclopropa[c]chromen-1(2H)-yl)carbonyl}-1a,7b-dihydrocyclopropa[c]chromen-2(1H]-ones as a single diastereoisomer with cis arrangement of the hydrogen atoms with respect to the cyclopropane ring plane. Reactions of the same electrophilic substrate with zinc enolates obtained from 1-aryl-2,2-dibromoalkanones led to the formation of 1-alkyl-1a-{(1-alkyl-1-aroyl-2-oxo-1,7b-dihydrocyclopropa[c]chromen-1(2H)-yl)carbonyl}-1-aroyl-1a,7b-dihydrocyclopropa[c]chromen-2(1H)-ones as a single diastereoisomer with trans arrangement of the alkyl group and hydrogen atom with respect to the cyclopropane ring plane.  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