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.
On irradiation in the presence of Fe(CO)5, the allenecarboxylates 1 afforded binuclear carbonyliron complexex 6 (Scheme 3), whereas the allenic lactone 7 under similar conditions gave a mixture of one binuclear and two mononuclear carbonyliron complexes ( 9 , 8 , and 10 ; Scheme 4). The structure of the complexes has been elucidated by X-ray crystallography. The structure of the binuclear complex 9 corresponds to that of 6 , while 8 has been shown to be a 1,3-butadiene(tricabonyl)iron complex. The unique structure of the 10 represents a new type of allenic complex. A stepwise formation of the complexes via intermediate allene(tetracarbonyl) iron complexes type 11 and 13 is suggested. Treatment of the binuclear complex 6b with FeCl3 led to the formation of the free ligand and a mixture of mononuclear complexes 13 and 14 (Scheme 5). On heating, the 1,3-diene complex 8 yielded the free ligand 15 , the prouduct of a (1,3) H shift in the allene 7 ; the complex 10 on the other hand liberates 7 on treatment with ethylenetracarbonitrile (TCNE) (Scheme 6).  相似文献   

2.
The allenecarboxylates 1a , b and allenic lactones 4a , b undergo thermally induced (1,3) H Shifts in the presence of Co2(CO)8. The non-isolated 1,3-dienes 2a , b react further affording the Diels-Alder Adducts 3a , b Scheme 1 in high yields. These adducts were not formed in the case of the 2-vinybutenolides 5a , b . On irradiationin the presence of Co2(CO)8 or Mn2(Co)10, the studied allenes reacted in a different manner, yielding either cyclization products 7 and 8 (Scheme 3) or products 9 and 10 , formed via H abstracton and solvent addition (Schemes 4 and 5).  相似文献   

3.
The thermal 4 : 1 equilibrium mixture of 1,3,5,6- and 1,3,5,10-tetramethylheptalene ( 13a and 13b , resp.) has been prepared, starting from the thermal equilibrium mixture of dimethyl 6,8,10-trimethylheptalene-1,2- and -4,5-dicarboxylate ( 6a and 6b , resp.; cf. Scheme 5). These heptalenes undergo double-bond shifts (DBS) even at ambient temperature. Treatment of the mixture 13a / 13b 4 : 1 with [Cr(CO)3(NH3)3] in boiling 1,2-dimethoxyethane resulted in the formation of all four possible mononuclear Cr(CO)3 complexes 19a – 19d of 13a and 13b , as well as two binuclear Cr(CO)3 complexes 20a and 20b , respectively, in a total yield of 87% (cf. Scheme 7). The mixture of complexes was separated by column chromatography, followed by preparative HPLC (cf. Fig. 2). The structures of all complexes were established by X-ray crystal-structure analyses (complex 19b and 20b ; cf. Figs. 6 – 8) and extensive 1H-NMR measurements (cf. Table 3). In 20b , the two Cr(CO)3 groups are linked in a `syn'-mode to the highly twisted heptalene π-skeleton. The correspondence of the 1H-NMR data of 20a with that of 20b indicates that the two Cr(CO)3 groups in 20a also have a `syn'-arrangement. The thermal behavior of the mononuclear complexes 19a – 19d has been studied at 85° in hexafluorobenzene (HFB). At this temperature, all four complexes undergo rearrangement to the same thermal equilibrium mixture (cf. Table 8). The rates for the thermal equilibration of each complex have been determined by 1H-NMR measurements (cf. Figs. 912) and analyzed by seven different kinetic schemes (Chapt. 2). The equilibration rates are in agreement with two different haptotropic rearrangements that take place, namely intra- and inter-ring shifts of the Cr(CO)3 group, whereby both rearrangements are accompanied by DBS of the heptalene π-skeleton (cf. Scheme 9). All individual kinetic steps possess similar ΔG values in the range of 29 – 31 kcal⋅mol−1 (cf. Table 8). The occurrence of inter-ring haptotropic migrations of Cr(CO)3 groups has already been established for anellated aromatic systems (cf. Scheme 10); however, it is the first time that these rearrangements have been unequivocally demonstrated for Cr(CO)3 complexes of non-planar bicyclic [4n]annulenes, such as heptalenes. The mechanism of migration may be similar to that proposed for aromatic systems (cf. Schemes 10 and 11).  相似文献   

4.
The desulfurization of several N,2-diaryl-5-(arylimino)-2,5-dihydro-4-nitroisothiazol-3-amines 5 with Ph3P led to complex mixtures of products in low yields. For instance, quinoxaline-2-carboxamide 1-oxides of type 6 (Scheme 2) and, in some cases, also 3-nitroquinolines of type 7 (Scheme 5) were isolated. By the desulfurization of the substituted derivatives 5b – e , a rearrangement of the intermediates yielded 6 and 7 with a different substitution pattern from that expected from the starting materials (Scheme 3). The additional formation of two isomeric 1,2,5-oxadiazole-3-carboxamides 8 was observed only in the case of 5d (R1=R2=F) (Scheme 6). Under the same reaction conditions, the major product of the desulfurization of 5c was the quinoxaline-2-carboxamide 1-oxide 9 (Scheme 7). Reaction mechanisms involving intermediate ketene imines and O transfer from the NO2 group to the neighboring ketene imine are proposed. The structures of 6a , 6e , 6k , 7b , and 8d were established by X-ray crystallography, while the structure of 9 was elucidated by 2D-NMR spectroscopy and corroborated by X-ray crystallography.  相似文献   

5.
Diphenyl phosphorazidate (DPPA) was used as the azide source in a one-pot synthesis of 2,2-disubstituted 3-amino-2H-azirines 1 (Scheme 1). The reaction with lithium enolates of amides of type 2 , bearing two substituents at C(2), proceeded smoothly in THF at 0°; keteniminium azides C and azidoenamines D are likely intermediates. Under analogous reaction conditions, DPPA and amides of type 3 with only one substituent at C(2) gave 2-diazoamides 5 in fair-to-good yield (Scheme 2). The corresponding 2-diazo derivatives 6–8 were formed in low yield by treatment of the lithium enolates of N,N-dimethyl-2-phenylacetamide, methyl 2-phenylacetate, and benzyl phenyl ketone, respectively, with DPPA. Thermolysis of 2-diazo-N-methyl-N-phenylcarboxamides 5a and 5b yielded 3-substituted 1,3-dihydro-N-methyl-2H-indol-2-ones 9a and 9b , respectively (Scheme 3). The diazo compounds 5–8 reacted with 1,3-thiazole-5 (4H)-thiones 10 and thiobenzophenone ( 13 ) to give 6-oxa-1,9-dithia-3-azaspiro[4.4]nona-2,7-dienes 11 (Scheme 4) and thiirane-2-carboxylic acid derivatives 14 (Scheme 5), respectively. In analogy to previously described reactions, a mechanism via 1,3-dipolar cycloaddition, leading to 2,5-dihydro-1,3,4-thiadiazoles, and elimination of N2 to give the ‘thiocarbonyl ylides’ of type H or K is proposed. These dipolar intermediates with a conjugated C?O group then undergo either a 1,5-dipolar electrocyclization to give spirohetrocycles 11 or a 1,3-dipolar electrocyclization to thiiranes 14 .  相似文献   

6.
Synthesis of Trifluoromethyl-Substituted Sulfur Heterocycles Using 3,3,3-Trifluoropyruvic-Acid Derivatives The reaction of methyl 3,3,3-trifluoropyruvate ( 1 ) with 2,5-dihydro-1,3,4-thiadiazoles 4a, b in benzene at 45° yielded the corresponding methyl 5-(trifluoromethyl)-1,3-oxathiolane-5-carboxylates 5a, b (Scheme 1) via a regioselective 1,3-dipolar cycloaddition of an intermediate ‘thiocarbonyl ylide’ of type 3 . With methyl pyruvate, 4a reacted similarly to give 6 in good yield. Methyl 2-diazo-3,3,3-trifluoropropanoate ( 2 ) and thiobenzophenone ( 7a ) in toluene underwent a reaction at 50°; the only product detected in the reaction mixture was thiirane 8a (Scheme 2). With the less reactive thiocarbonyl compounds 9H-xanthene-9-thione ( 7b ) and 9H-thioxanthene-9-thione ( 7c ) as well as with 1,3-thiazole-5(4H)-thione 12 , diazo compound 2 reacted only in the presence of catalytic amounts of Rh2(OAc)4. In the cases of 7a and 7b , thiiranes 8b and 8c , respectively, were the sole products (Scheme 3). The crystal struture of 8c has been established by X-ray crystallography (Fig.). In the reaction with 12 , desulfurization of the primarily formed thiirane 14 gave the methyl 3,3,3-trifluoro-2-(4,5-dihydro-1,3-thiazol-5-ylidene)propanoates (E)-and (Z)- 15 (Scheme 4). A mechanism of the Rh-catalyzed reaction via a carbene addition to the thiocarbonyl S-atom is proposed in Scheme 5.  相似文献   

7.
Boron-Trifluoride-Catalyzed Reactions of 3-Amino-2H-azirines with Amino-acid Esters and Amines After activation by protonation or complexation with BF3, 3-amino-2H-azirines 1 react with the amino group of α-amino-acid esters 3 to give 3,6-dihydro-5-aminopyrazin-2(1H)-ones 4 by ring enlargement (Scheme 2, Table 1). The configuration of 3 is retained in the products 4 . With unsymmetrically substituted 1 (R1 ≠ R2), two diastereoisomers of 4 (cis and trans) are formed in a ratio of 1:1 to 2:1. With β-amino-acid esters 5 and 7 , only openchain α-amino-imidamides 6 and 8 , respectively, are formed, but none of the seven-membered heterocycle (Scheme 3). Primary amines also react with BF3-complexed 1 to yield α-amino-imidamides of type 9 (Scheme 4, Table 2). Compound 9b is characterized chemically by its transformation into crystalline derivatives 10 and 12 with 4-nitrobenzoyl chloride and phenyl isothiocyanate, respectively (Scheme 5). The structure of 12 is established by X-ray crystallography. Mechanisms for the reaction of activated 1 with amino groups are proposed in Schemes 6 and 7.  相似文献   

8.
Synthesis of (6R, all-E)-Neoxanthin and Related Allenic Carotenoids We present the first synthesis of enantiomerically pure neoxanthin ( 1 ) by a Wittig-Horner condensation between the ylide from the novel diethyl 12′-apo-15, 15′-didehydroviolaxanthin-12′-phosphonate ( 35 ) and the allenic C15-aldehyde 31 (Scheme 4) via the crystalline 15, 15′-didehydroneoxanthin ( 36 ; 70% yield). After partial hydrogenation of the triple bond of 36 and isomerisation of the (15Z)-intermediate 37 , neoxanthin ( 1 ) was obtained in good yield. Similar syntheses gave (15Z, 9′Z)-neoxanthin ( 45 ; Scheme 5) and (9Z)-15, 15′-didehydroneoxanthin ( 47 ; Scheme 6). Comparison of the physical data of synthetic 1 with those of a freshly isolated sample of neoxanthin from the flowers of Trollius europaeus confirmed their identity. The unusually low melting point of 1 is caused by a very easy thermal isomerisation into a mixture of the neochromes 4 and 5 (Scheme 1). Such a thermal rearrangement is not observed with 15, 15′-didehydroneoxanthin ( 36 ). To explain this, we assume a zwitterionic excited state of the allenic group that induces the rearrangement of the violaxanthin end group into the furanoid epoxide (Scheme 7).  相似文献   

9.
Two new chiral bidentate (phosphinophenyl)benzoxazine P,N-ligands 2a and 2b were synthesized from highly enantiomer-enriched 2-(1-aminoalkyl)phenols 4 . Ligand rac- 2a was obtained on refluxing the t-Bu-substituted (aminomethyl)phenol 4a with 2-(diphenylphosphino)benzonitrile in chlorobenzene in the presence of anhydrous ZnCl2 followed by decomplexation (Scheme 2). This reaction, when carried out with (+)-(S)- 4a , was accompanied by racemization at the stereogenic center of the alkyl side chain. The enantiomerically pure ligands (+)-(R)- 2a and (−)-(S)- 2a were obtained using a stepwise procedure via the amides (−)-(R)- and (+)-(S)- 5b , respectively, followed by cyclization to benzoxazines (+)-(R)- and (−)-(S)- 7b , respectively, with triflic anhydride and by F-atom substitution by diphenylphosphide (Schemes 3 and 5). In the case of the i-Pr analogue 2b , this last step resulted in racemization (Scheme 6). This was overcome by preparing the bromo derivative and introducing the diphenylphosphine group via Br/Li exchange and reaction with chlorodiphenylphosphine (Scheme 7). The first application of (+)-(R)- 2a in an asymmetric Heck reaction showed high enantioselectivity (91%) (Scheme 8).  相似文献   

10.
Synthesis of Enantiomerically Pure Mimulaxanthin and of Its (9Z,9′Z)- and (15Z)Isomers We present the details of a synthesis of optically active, enantiomerically pure stereoisomers of mimulaxanthin (=(3s,5R,6R,3′S,5′R,6′R)-6,7,6′,7′-tetradehydro-5,6,5′,6′-tetrahydro-β,β-carotin-3,5,3′,5′-tetrol) either as free alcohols 1a and 24a or as their crystalline (t-Bu)Me2Si ethers 1b and 24b . Grasshopper ketone 2a , a presumed synthon, unexpectedly showed a very sluggish reaction with Wittig-Horner reagents. Upon heating with the ylide of ester phosphonates, an addition across the allenic bond occurred. On the contrary, a slow but normal 1,2-addition took place with the ylide from (cyanomethyl)phosphonate but, unexpectedly, with concomitant inversion at the chiral axis. So a mixture of(6R,6S,9E,9Z)-isomers 6 – 9 was produced {(Scheme 1). However, a fast and very clean 1,2-addition occurred with the ethynyl ketone 12 to yield the esters 13 and 14 (Scheme 2). DIBAH reduction of the separated stereoisomers gave the allenic alcohols 15 and 16 in high yield. Mild oxidation to the aldehydes 17 and 18 followed by their condensation with the acetylenic C10-bis-ylide 19 led to the stereoisomeric 15,15′-didehydromimulaxanthins 20 and 22 , respectively (Schemes 3 and 4). Mimulaxanthins 1 and 24 were prepared by partial hydrogenation of 20 and 22 followed by a thermal (Z/E)-isomerization. As expected, the mimulaxanthins exhibit very weak CD curves, obviously caused by the allenic bond that insulates the chiral centers in the end group from the chromophor. On the contrary, some of the C15-allenic synthons showed not only fairly strong CD effects but also a split CD curve which, in our interpretation, results from an exciton coupling between the allene and the C(9)?C(10) bond. We postulate a rotation around the C(8)? C(9) bond, presumably caused by an intramolecular H-bond in 16 or by a dipol interaction between the polarized double bonds in 6 , 7 , 8 , and 17 .  相似文献   

11.
The [Co2(CO)8]-mediated retro-Diels-Alder reaction of the annelated barrelenes 1 afforded the 1H-indol-2(3H)-one derivatives 3 (Scheme 1), while the hydrobarrelene 4a , under the same conditions, was converted to the anilide 6 (Scheme 2); 4b remained unaffected. The direct irradiation of 1 led to the annelated cyclooctatetraenes 7 (Scheme 3). On irradiation in the presence of excess of [Fe(CO)5], 1a , 1b , and 4a gave the tricarbonyliron complexes 8 , 9 , and 11 , respectively (Schemes 3 and 4); under these conditions, 4b was inert.  相似文献   

12.
Reaction of the thiosemicarbazone ligands C4H4NC(H)=NN(H)C(S)NHR (R = Me, a ; Et, b ) with Li2[PdCl4] gave the dinuclear complexes [Pd{C4H4NC(H)=NNC(S)NHR}(μ‐Cl)]2 (R = Me, 1a ; Et, 1b ) with a central Pd2Cl2 core and with deprotonation of the thiosemicarbazones at the hydrazinic nitrogen atom. Treatment of 1a and 1b with triphenylphosphine gave the mononuclear compounds [Pd{C4H4C(H)=NNC(S)NHR}(Cl)(PPh3)] (R = Me, 2a ; Et, 2b ), whereas reaction of 1a and 1b with tertiary diphosphines gave mono‐ and dinuclear compounds, as appropriate, with the corresponding diphosphine acting as a monodentate ( 6b ), chelating ( 3a ) and bridging ligand ( 4a, 5a , 4b, 5b ). Treatment of 1a and 1b with (Ph2PCH2CH2PPh2)W(CO)5 gave the new heterobimetallic complexes 7a and 7b . The crystal structures of complexes 3a and 4a are described.  相似文献   

13.
The polymerization of ethylene was studied by using a series of polysiloxane-bridged dinuclear zirconocenes [(SiMe2O)nSiMe2(C5H4)2][(C9H7)ZrCl2]2 ( 7 , n = 1 ; 8 , n = 2; 9 , n = 3), the corresponding mononuclear zirconocene (C5H5)(C9H7)ZrCl2, 10 , and the pentamethylene-bridged dinuclear zirconocene [(CH2)5(C5H4)2][(C9H7)-ZrCl2]2, 13 . From the polymerization studies using these catalysts it was found that (i) activities of the polysiloxane dinuclear zirconocenes 7–9 wre lower than that of the corresponding mononuclear zirconocene 10 , (ii) molecular weights of polyethylenes produced by the dinuclear metallocenes are greater than that of polyethylene produced by the mononuclear metallocene, (iii) the complex 9 holding the longest bridging ligand exhibited the highest activity but produced a polymer having the smallest molecular weight among the polysiloxane-bridged dinuclear zirconocenes, and (iv) the pentamethylene-bridged dinuclear metallocene 13 showed higher activity than the complexes 7–9 and the mononuclear zirconocene 10 . The formation of the lowest molecular weight of polyethylene by 9 was attributed to the influence of electron withdrawal caused by the Lewis acid–base interaction between the acidic aluminum of the cocatalyst and the basic oxygen at the polysiloxane linkage as well as the lack of a steric problem. An increase in steric congestion around the metal center led to not only a decrease in catalytic activity due to preventing facile monomer access to the active site but also an increase in the molecular weight of polyethylenes due to supressing β-H elimination. © 1997 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 35: 3717–3728, 1997  相似文献   

14.
4-Alkoxy-1,3-oxazol-5(2H)-ones of type 4 and 7 were synthesized by two different methods: oxidation of the 4-(phenylthio)-1,3-oxazol-5(2H)-one 2a with m-chloroperbenzoic acid in the presence of an alcohol gave the corresponding 4-alkoxy derivatives 4 , presumably via nucleophilic substitution of an intermediate sulfoxide (Scheme 2). The second approach is the BF3-catalyzed condensation of imino-acetates of type 6 and ketones (Scheme 3). The yields of this more straightforward method were modest due to the competitive formation of 1,3,5-triazine tricarboxylate 8. At 155°, 1,3-oxazol-5(2H)-one 7b underwent decarboxylation leading to an alkoxy-substituted nitrile ylide which was trapped in a 1,3-dipolar cycloaddition by trifluoro-acetophenone to give the dihydro-oxazoles cis- and trans- 9 (Scheme 4). In the absence of a dipolarophile, 1,5-dipolar cyclization of the intermediate nitrile ylide yielded isoindole derivatives 10 (Schemes 4 and 5).  相似文献   

15.
The 3,3‐dichloro‐2,2,4,4‐tetramethylcyclobutanethione ( 4b ) was prepared from the parent diketone by successive reaction with PCl5 and Lawesson reagent in pyridine. This new thioketone 4b was transformed into 1‐chlorocyclobutanesulfanyl chloride 5 and chloro 1‐chlorocyclobutyl disulfide 9 by treatment with PCl5 and SCl2, respectively, in chlorinated solvents (Schemes 1 and 2). These products reacted with S‐ and P‐nucleophiles by substitution of Cl at the S‐atom; e.g., the reaction with 4b yielded the di‐ and trisulfides 6b and 11 , respectively. Surprisingly, only pentasulfide 12 was formed in the reaction of 9 with thiobenzophenone (Scheme 3). In contrast to 5 and 9 , the corresponding chloro 1‐chlorocyclobutyl trisulfide 13 could not be detected, but reacted immediately with the starting thioketone 4b to give the tetrasulfide 14 (Scheme 4). Oxidation of 4b with 3‐chloroperbenzoic acid (mCPBA) yielded the corresponding thione oxides (= sulfine) 15 , which underwent 1,3‐dipolar cycloadditions with thioketones 3a and 4b (Scheme 5). Furthermore, 4b was shown to be a good dipolarophile in reactions with thiocarbonylium methanides (Scheme 6) and iminium ylides (= azomethine ylides; Scheme 7). In the case of phenyl azide, the reaction with 4b gave the symmetrical trithiolane 25 (Scheme 8).  相似文献   

16.
Phase-transfer alkylation of the 2-oxocycloalkane-l-carbonitriles 1a and 1b with ethyl 4-bromo-3-methoxy-2-butenoate ( 2 ), followed by deprotection and base-catalyzed cyclization gave the annulated cyclopentenones 5a and 5b , respectively, in high overall yields (Scheme 1). Stereoselective catalytic hydrogenation of 5b followed by de-ethoxycarbonylation afforded 14-oxo-cis-bicyclo[10.3.0]pentadecane-l-carbonitrile ( 7 ). Treatment of 7 with LiN(i-Pr)2 in THF gave the known synthetic muscone precursor 8 (Scheme 2). The tricyclo[10.4.0.01,15]hexadecan-14-one ( 14 ) was prepared from 7 in 5 steps by a reaction sequence proceeding without affecting the chiral centres (Scheme 2). The structure of 14 was established by X-ray structure analysis (Figure).  相似文献   

17.
The C? C coupling of the two bicyclic, unsaturated dicarboximides 5 and 6 with aryl and heteroaryl halides gave, under reductive Heck conditions, the C‐aryl‐N‐phenyl‐substituted oxabicyclic imides 7a – c and 8a – c (Scheme 3). Domino‐Heck C? C coupling reactions of 5, 6 , and 1b with aryl or heteroaryl iodides and phenyl‐ or (trimethylsilyl)acetylene also proved feasible giving 8, 9 , and 10a – c , respectively (Scheme 4). Reduction of 1b with LiAlH4 (→ 11 ) followed by Heck arylation and reduction of 5 with NaBH4 (→ 13 ) followed by Heck arylation open a new access to the bridged perhydroisoindole derivatives 12a , b and 14a , b with prospective pharmaceutical activity (Schemes 5 and 6).  相似文献   

18.
The reactions of 1,1,3,3-tetramethyl-8-thia-5,6-diazaspirol[3.4]oct-5-en-2-one ( 1a ) with imidazole-2-thiones 3 and pyrimidine-2(1H)-thione ( 6 ) in CHCl3 at 40 – 50° yield 2,2,4,4-tetramethylcyclobutanone dithioacetals of type 4 and 7 , respectively, by interception of the intermediate thiocarbonyl ylide 2a (Scheme 2). Thiirane 5 is formed as a minor product by 1,3-dipolar electrocyclization of 2a . When thioacetamide ( 8a ) and thiobenzamide ( 8b ) are used as trapping reagents, the primary adduct 10 undergoes a spontaneous cyclization by intramolecular nucleophilic addition of the imino group at the carbonyl group to yield bicyclic products of type 9 . The structure of 9a has been established by X-ray crystallography.  相似文献   

19.
On the Reactivity of Titanocene Complexes [Ti(Cp′)22‐Me3SiC≡CSiMe3)] (Cp′ = Cp, Cp*) towards Benzenedicarboxylic Acids Titanocene complexes [Ti(Cp′)2(BTMSA)] ( 1a , Cp′ = Cp = η5‐C5H5; 1b , Cp′ = Cp* = η5‐C5Me5; BTMSA = Me3SiC≡CSiMe3) were found to react with iodine and methyl iodide yielding [Ti(Cp′)2(μ‐I)2] ( 2a / b ; a refers to Cp′ = Cp and b to Cp′ = Cp*), [Ti(Cp′)2I2] ( 3a / b ) and [Ti(Cp′)2(Me)I] ( 4a / b ), respectively. In contrast to 2a , complex 2b proved to be highly moisture sensitive yielding with cleavage of HCp* [{Ti(Cp*)I}2(μ‐O)] ( 7 ). The corresponding reactions of 1a / b with p‐cresol and thiophenol resulted in the formation of [Ti(Cp′)2{O(p‐Tol)}2] ( 5a / b ) and [Ti(Cp′)2(SPh)2] ( 6a / b ), respectively. Reactions of 1a and 1b with 1,n‐benzenedicarboxylic acids (n = 2–4) resulted in the formation of dinuclear titanium(III) complexes of the type [{Ti(Cp′)2}2{μ‐1,n‐(O2C)2C6H4}] (n = 2, 8a / b ; n = 3, 9a / b ; n = 4, 10a / b ). All complexes were fully characterized analytically and spectroscopically. Furthermore, complexes 7 , 8b , 9a ·THF, 10a / b were also be characterized by single‐crystal X‐ray diffraction analyses.  相似文献   

20.
A series of racemic 6,6’-[(6,6’-dimethyl-[1,1’-biphenyl]-2,2’-diyl)bis(nitrylomethilidyne)]-bis(2-R1-4-R2-phenol) proligands ( L1H2 , R1=R2=Me; L2H2 , R1=tBu, R2=Me; L3H2 , R1=R2=cumyl; L4H2 , R1=CPh3, R2=Me) were reacted with {Mg[N(SiMe3)2]2}2 to provide mononuclear and dinuclear magnesium complexes [ L1 2Mg2] ( 1 ), L2–4 Mg ( 2 – 4 ), { L1 – 3 [MgN(SiMe3)2]2} ( 5 – 7 ). Complexes 3* and 4* in which each metal center is coordinated with a THF molecule were obtained when the corresponding crude complexes were recrystallized with a mixture of THF and n-hexane. Similarly, the formation of THF coordinated structure 7* of the heteroleptic dinuclear complex 7 was identified. The molecular structures of complexes 3* , 6 and 7* were established by X-ray single crystal diffraction studies, which show that mononuclear complex 3* possesses a five-coordinated metal center adopting a distorted square pyramid configuration, the two metal centers of the dinuclear complex 6 are bridged by two phenoxy oxygen atoms and each has a four-coordinated distorted tetrahedral configuration, and each metal center of the dinuclear complex 7* is still four-coordinated upon the coordination of THF but without bridging to each other. All complexes were investigated for the ring-opening polymerization (ROP) of rac-lactide (rac-LA) at 60 °C in toluene or tetrahydrofuran. Compared with the mononuclear counterparts, the dinuclear magnesium silylamido complexes showed significantly higher activities.  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