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 281 毫秒
1.
Since we found certain structural features of triaziridine ( 1 ) obtained by MNDO calculations to be in qualitative agreement with those derived earlier from ab initio calculations, we used the MNDO method to derive properties of formyltriaziridine ( 2 ) and 1-formyl-2,3-diisopropyltriaziridine ( 3 ) as models for the preparatively known 2,3-dialkyl-triaziridine-1-carboxylates 4 and 5 . The main results are: (a) The triaziridine N-atoms with H, alkyl, or formyl as substituents (see 2 and 3 ) are pyramidal. N(1) carrying the formyl group is flatter than N(2) and N(3) with H or alkyl substitent. Bond lengths and angles at N(2) and N(3) are almost identical with those calculated for the N-atoms of 1 . (b) The MNDO inversion barriers at the H-substituted N(2) and N(3) of 2 are higher than those at the formyl-substituted N(1), but similar to the ab initio barriers at the N-atoms of 1 . (c) The MNDO inversion barriers at N(1) of 2 and 3 are 53 to 92 kJ/mol, whereas the rotation barriers around the N(1)–C(4) bond are 7 to 23 kJ/mol; thus, the previously observed dynamic NMR phenomena in trans-2,3-diisopropyltriaziridine-carboxylates ( 5 ) can now be assigned to the slowing down of N(1) inversion rather than N(1)–C(4) rotation.  相似文献   

2.
Stable Pyramidal configurations at the Nitrogen Atoms of Dialkyl-and Trialkyl-triaziridines Stereochemical features of the recently synthesized nine samples of di- and trialkyl-triaziridines, namely the 1,3-cyclopentylen-(series a ) and the two stereoisomers of the diisopropyl derivatives (series b and c ), containing as the third substituent an H-atom ( 2 ), a CH3 group ( 3 )or a CH2OH group ( 4 ), were elaborated on the basis of the 1H-, 13C-, and 15N-NMR spectra. The three N-atoms of the saturated N3-homocycle were found to be stable to pyramidal inversion in all cases. According to their NMR spectra, 2 – 4 of the series a and b possess twofold symmetry (Cs), while 2 – 4 of series c are asymmetric. Thus, series c has the trans-configuration at N(2)/N(3) and, consequently, the cis-configuration at N(1)/N(2), while series a and b have the cis-configuration at N(2)/N(3) and -since the all-cis-arrangement is excluded-the trans-configuration at N(1)/N(2). The asymmetry of the trans-configurated 2c turned into twofold symmetry (C2), when a little CF3COOH was added. The 1H- and 13C-NMR data of series b and c of our alkyl-triaziridines exhibit a shielding effect, according to which there are two types of i-Pr groups, i-Pr(a) and i-Pr(b). They differ in the NMR signals of the H- and the C-atoms of their CH groups: the H-atoms of i-Pr(a) are more deshielded by 0.75–1.111 ppm and its C-atoms are more shielded by 10.0–160.0 ppm as compared to the corresponding atoms of i-Pr(b). i-Pr(a) is cis (on the N3-homocycle) to a large substituent (such as i-Pr, Me, CH2OH) and to a lone pair, while i-Pr(b)is cis only to a small (H) or to no substituent and to one or two lone pairs. An analogous effect appears in the NMR signals of the CH3 and CH2OH groups at N(1) of 3 and 4 in the series b and c .  相似文献   

3.
The syntheses of the linear tetraamines H2N? (CH2)2? NH? (CH2)2? NH? (CH2)2? NH2 (n = 4 and 5) are described. The protonation of the homologous tetraamines for n = 2, 3, 4, 5, 6 and 8, as well as of N-methylethylenediamine, was investigated using potentiometric and calorimetric measurements. The results obtained are discussed taking into consideration the substituent effect on the basicity of the aminic N-atoms.  相似文献   

4.
By LiAlH4 (Cl3Si)2CH2, (Cl2Si? CH2)2SiCl2 are reduced to (H3Si)2CH2 (a), (H3Si? CH2)2SiH2 (b) and (H2Si? CH2)3(c). However with the compounds (Cl3Si)2CCl2, (Cl3Si? CCl202SiCl2 and (Cl2Si? CCl2)3 cleavages of the Si? C-bond and reduction of the CCl-groups occur apart from the normal reduction of the Si-Cl-groups to (H3Si)2CCl2 (d), (H3SiCCl2)2SiH2 (e) and (H2Si? CCl2)3. Excess LiAlH4 favours this cleavage, the exact amount of a quarter of a mole LiAlH4 per SiCl-group allows the formation of (d), (e), (f). The cleavage of (e) is in accordance with: (1), (2),(3). Therefore SiH34 and (H3Si)2CCl2 are the main-reaction-products and CH3SiH3 is formed acc. to equ. (3). Because of the cleavage of (H2Si? CCl2)3 with LiAlH4 H3Si? CCl2? SiH2? CH3and H3Si? CH2? SiH2? CH2? SiH2? CH3 are preferentially formed after the hydrolysis. The CH2-containing compounds (a), (b), (c) cannot be cleaved in an analogous reaction.  相似文献   

5.
Building upon previous studies on the synthesis of bis(sigma)borate and agostic complexes of ruthenium, the chemistry of nido‐[(Cp*Ru)2B3H9] ( 1 ) with other ligand systems was explored. In this regard, mild thermolysis of nido‐ 1 with 2‐mercaptobenzothiazole (2‐mbzt), 2‐mercaptobenzoxazole (2‐mbzo) and 2‐mercaptobenzimidazole (2‐mbzi) ligands were performed which led to the isolation of bis(sigma)borate complexes [Cp*RuBH3L] ( 2 a – c ) and β‐agostic complexes [Cp*RuBH2L2] ( 3 a – c ; 2 a , 3 a : L=C7H4NS2; 2 b , 3 b : L=C7H4NSO; 2 c , 3 c : L=C7H5N2S). Further, the chemistry of these novel complexes towards various diphosphine ligands was investigated. Room temperature treatment of 3 a with [PPh2(CH2)nPPh2] (n=1–3) yielded [Cp*Ru(PPh2(CH2)nPPh2)‐BH2(L2)] ( 4 a – c ; 4 a : n=1; 4 b : n=2; 4 c : n=3; L=C7H4NS2). Mild thermolysis of 2 a with [PPh2(CH2)nPPh2] (n=1–3) led to the isolation of [Cp*Ru(PPh2(CH2)nPPh2)(L)] (L=C7H4NS2 5 a – c ; 5 a : n=1; 5 b : n=2; 5 c : n=3). Treatment of 4 a with terminal alkynes causes a hydroboration reaction to generate vinylborane complexes [Cp*Ru(R?C?CH2)BH(L2)] ( 6 and 7 ; 6 : R=Ph; 7 : R=COOCH3; L=C7H4NS2). Complexes 6 and 7 can also be viewed as η‐alkene complexes of ruthenium that feature a dative bond to the ruthenium centre from the vinylinic double bond. In addition, DFT computations were performed to shed light on the bonding and electronic structures of the new compounds.  相似文献   

6.
Triaziridines. Synthesis of cis-2,3-Diisopropyltriaziridine-1-carboxylic Esters Irradiation of the (Z)-azimines 1a , b in Et2O with a Hg high pressure lamp through Corex yielded (besides 30% of the previously described trans-triaziridines 3a , b ) 15% of the new cis-triaziridines 4a , b . The same irradiation of the (E)-azimines 2a , b afforded only 15–18% of 3a , b but 20–23% of 4a , b . Thus, these azimine photocyclizations show some stereospecificity. The triaziridines 3a , b and 4a , b formed in this way were always accompanied by the same three types of by-products, namely 10–15% of the ‘triazones’ 5a , b , 11–20% of the carbamic esters 6a , b , and 5–10% of the ether/nitrene insertion products 7a , b . The constitution and configuration of the new cis-triaziridines 4 followed from their spectral properties. Of particular interest are the symmetry properties of 4 derived from the 1H-, 13C-, and 15N-NMR spectra: The stereoisomers 3 and 4 differ only in that the isochronicity of the two constitutionally equivalent molecular halves is temperature dependent in 3 but independent in 4 . Both triaziridines 3 and 4 exhibit the IR CO band at (for carbamates) remarkably high frequency. The results confirm that the alkyl-substituted N-atoms of triaziridines are pyramidally stable, that the corresponding acyl-substituted N-atoms (N(1)) are also pyramidal, but can invert more readily, and that rotation around the N(1), C(?O) bond is rapid. Thus, there can be only little amide-type delocalization between a triaziridine N-atom and an acyl substituent of the carbamate type attached to it.  相似文献   

7.
The aldol adducts 1a – 13a of R,R-2(tertbutyl)-6-methyl-1,3-dioxan-4-one (from 3-hydroxybutanoic acid) to aldehydes, single diastereoisomers obtained as described previously, are acetylated or benzoylated to the corresponding esters 1b – 5b and 6c – 13c , respectively, which in turn are reduced with LiAlH4 to the title compounds 14 – 24 . The enantiomerically pure triols thus available may be useful as chiral building blocks, as auxiliaries for enantioselective reactions, and as center pieces for chiral dendrimers.  相似文献   

8.
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   

9.
Contributions to the Chemistry of Boron, 134. Adducts of (Dimethylamino)boranes with Aluminium and Gallium Halides2) The (dimethylamino)boranes (CH3)2BN(CH3)2 ( 1 ) and RB[N(CH3)2]2 ( 2, 3 ) form 1 : 1 adducts 1a – c and 2a , b , 3a , b with AlCl3, AlBr3, and GaCl3, respectively, in contrast to B[N(CH3)2]3 ( 4 ) which reacts with GaCl3 to produce a 1:2 adduct 4a . NMR and IR data are in accord with the presence of a simple N – Al or N – Ga coordinative bound for the first two classes of compounds. However, 4a is to be regarded as a tris(dimethylamino)borane-dichlorogallium(1+) tetrachlorogallate containing a bidentate aminoborane component.  相似文献   

10.
Triaziridines. Ring Openings of Triaziridines Eleven triaziridine derivatives were heated at 60° in CDCl3 to obtain information on the tendency towards, resp. the resistance to, ring opening of the N3-homocycle by thermolysis. Among these triaziridines, there are three which contain, as one of the substituents, a methoxycarbonyl group (ester derivatives 1 , 5 and 16 ), three a methyl group (methyl derivatives 18 , 24 , and 26 ), three an H-atom ( 14 , 27 , and 30 ), and two a negative charge ( 31 and 32 ). The other two substituents in each of these four classes of triaziridines are trans-located i-Pr groups ( 1 , 18 , 27 , and 31 ), cis-located i-Pr groups ( 5 , 24 , 14 , and 32 ), and a 1,3-cis-cyclopentylidene group ( 16 , 26 , and 30 ). As major products these mild thermolyses, we isolated : from the trans-ester 1 and from the annellated ester derivative 16 , the 1-acyl-azimines 2 and 17 , respectively, from the cis-ester 5 , the 3-acyl-triazene 4 , from the trans-methyl derivative 18 , the (E)-diazene 19 , and hexamine 21 , from the cis-methyl derivative 24 the 2-methylazimine 25 , both from the trans- and cis-H-derivatives 27 , and 14 , respectively, the H- triazene 13 and, finally, both from the trans-and cis-anion 31 and 32 , respectively – after protonation the H-triazene 13 and – after methylation – the methyl-triazene 33 . The same thermolysis of the annellated methyl and H-derivatives 26 and 30 , respestively, resulted only in decomposition. These results can be uniformly interpreted with a primary opening of the triaziridine ring by rupture of one of the two types of N? N bonds lending to azimines or triazenide anions. Some of the azimines were isolable, namely 2 , 17 , and 25 , and one was spectroscopically observable as an intermediate, namely 11 on the way to the triazene 4 . The other azimines are plausible intermediates to the isolated products, namely 15 on the way to 13 , and 22 on the way to 19 and 21 . The triazenide anion 28 is the evident intermediate on the way to 13 or to 33 . The annellated azimines are assumed not be formed from 26 and 30 , or then to be be decomposed under the conditions of their formation. We conclude that the triaziridine derivatives 1, 16 , and 18 underwent thermal ring opening between N(1) and N(2), while the derivatives 5 , 14 , 24 , 27 , 31 , and 32 were ruptured between N(2) and N(3); no conclusion was possible on the ring opening of the derivatives 26 and 30 . The predominant formation of the (Z)-azimine 2 from the trans-triaziridine 1 , and of the (E)-isomer 3 – among the two azimines – from the cis-triaziridine 5 suggests a stereospecificity in the triaziridine ring openings. This would, however, not be expected to be observable in the products from the other triaziridines, since both N? N bonds of the azimine 25 and of the anion 28 probably rotate rapidly and since the secondary trans formations of the other primary products are not able to retain configurational information.  相似文献   

11.
The portions of the N3H3 singlet potential energy surface corresponding to triaziridines ( 1 ), azimines ( 2 ) and triazenes ( 3 ) have been calculated by ab initio SCF using 3-21G, 6-31G, and 6-31G** basis sets. Minima and transition states were located by force gradient geometry optimization. The most important computation results are: (1) Triaziridines ( 1 ): The configuration at the 3 N-atoms is pyramidal. There are 2 stereoisomers, 1a and 1b . The c,t-isomer 1a has less energy than the c,c-isomer 1b . The 2 stereoisomerizations by N-inversion hve rather high activation energies. The N,N bonds in 1 are longer and weaker (STO-3G estimation) than in hydrazine. The N-homocycle 1 exhibits less ring strain than the C-homocycle cyclopropane or three-membered heterocycles. (2) Azimine ( 2 ): All 6 Atoms are in the same plane. There are 3 stereoisomers, 2a, 2b , and 2c . The order of ground state energies is (Z,Z) < (E,Z) ? (E,E). The 2 N,N bond lengths correspond to multiplicity 1½. The electronic structure of 2 corresponds to a 1,3-dipole with almost equal delocalization of the 4 π-electrons over all 3 N-atoms. The negative net charge at the central N-atom is much less than that at the terminal N-atoms. Azimines should behave as π-donors in complexation with transition metals (3) Triazene ( 3 ): All 6 atoms are in the same plane. There are 2 stereoisomers, 3a and 3b . The order of ground-state energies is (E) < (Z). The stereoisomerization proceeds as pure N-inversion. N-Inversion has a high energy barrier inversion at N(1) is faster than at N(2). One of the N,N bond lengths is typical for a double, the other for a single bond. The electronic structure of triazene 3 entails rather localized π- and p-electron pairs at N(1),N(2) and at N(3). Triazenes should behave as p-donors in complexation with transition metals. (4) -N3H3-Isomers: The order of ground-state energies is 3 < 2 < 1 . The energy differences between these constitutional isomers are much larger than between the stereoisomers of each. The [1,2]-H shifts for conversions of 2 to 3 and the [1,3]-H shift for tautomerization of 3 have relatively high activation energies; both shifts can be excluded as modes of thermal, unimolecular transformations.  相似文献   

12.
CF3SO2N?SCl2 reagiert mit (CH3)2S[NSi(CH3)3]2, (C4H8)S[NSi(CH3)3]2 oder (C5H10)S[NSi(CH3)3]2 unter Trimethylchlorsilanabspaltung zu den achtgliedrigen S4N4-Derivaten S4N4(NSO2CF3)2(CH3)4 3 , S4N4(NSO2CF3)2(C4H8)2 4a und S4N4(NSO2CF3)2(C5H1 0)2 4b . In den achtgliedrigen SN-Ringen haben die Schwefelatome die Koordinationszahl 3 und 4. Die Röntgenstrukturanalyse von 4a ergab eine Sessel-Konformation. 4a kristallisiert orthorhombisch in der Raumgruppe Pna21 mit a = 17,641(4), b = 6,406(2), c = 19,130(4) Å, dx = 1,815 g cm?3 und Z = 4. Die mittleren S? N-Abstände betragen an den vierfach koordinierten Schwefelatomen 1,597 Å und an den Schwefelatomen mit der Koordinationszahl 3 1,650 Å. CF3SO2N? SCl2 reagiert mit trimethylzinnhaltigen S? N-Verbindungen zum bekannten CF3SO2N[Sn(CH3)3]S(CH3)NSO2CF3 und Dimethylzinndichlorid. Synthesis and X-Ray Structure Analysis of S4N4-Derivatives with Threefold and Fourfold Coordinated Sulfur Atoms CF3SO2N?SCl2 reacts with (CH3)2S[NSi(CH3)3]2, (C4H8)S[NSi(CH3)3]2 or (C5H10S[NSi(CH3)2]2 under elimination of (CH3)3SiCl to yield the eight-membered S4N4 derivatives S4N4?NSO2CF3)2(CH3)4, 3 , S4N4(NSO2CF3)2(C4H8)2 4a und S4N4(NSO2CF3)2(C5H1 0)2 4b . In the eight-membered SN-rings the sulfur atoms have the coordination number 3 and 4. The X-ray structure analysis of 4a revealed a chair conformation. 4a crystallizes in the orthorhombic space group Pna21 with a = 17.641(4), b = 6.406(2), c = 19.130(4) Å, dx = 1.815 g cm?3, and Z = 4. The average S? N distance was found to be 1.597 Å at fourfold coordinated sulfur atoms and 1.650 Å at sulfur with coordination number 3. CF3SO2N=SCl2 reacts with trimethyl tin-containing S? N compounds to the known CF3SO2N[Sn(CH3)3]S(CH3)NSO2CF3 and dimethyl tin dichloride.  相似文献   

13.
The preparation of ylides of the general structure is described. Thermolysis of 14a (R = CH3, R' = H, Ar = C6H5) gave dimethylamine and 2,4-dimethyl-6-phenyl-s-triazine. Thermolysis of ylides 14b (R = C6H5; R' = CH3, Ar = C6H5) and 14c (R = C6H5, R' = CH3, Ar = p-tolyl) gave dimethylamine, ArCH = NCH3 and 1-methyl-2-Ar-4,6-diphenyl-1,2-dihydro-s-triazines ( 19a,b ). Triazines 19a and 19b were also prepared by condensation of N-methylbenzamidine with benzaldehyde and p-tolualdehyde, respectively. Thermolysis of 14d (R = C6H5, R1 = CH2C6H5,Ar = C6H5) gave 1-benzyl-2,4,6-triphenyl-1,2-dihydro-s-triazine ( 19c ) and N-benzylidenebenzylamine. Mechanistic aspects of these reactions are discussed.  相似文献   

14.
1,2,5-Thiadiazolidin-3-one 1,1-dioxide derivatives (±)- 1a – d and (±)- 2 were designed by molecular modeling as MHC (major histocompatibility complex) class-II inhibitors. They were prepared from the unsymmetrically N,N′-disubstituted acyclic sulfamides (±)- 4a – d (Scheme 1) and (±)- 11 (Scheme 2). These N-alkyl-N′-arylsulfamide precursors were synthesized by nucleophilic substitution of either a sulfamoyl-chloride or a N-sulfamoyloxazolidinone. Extension of base-induced cyclization methods from aliphatic to aromatic sulfamides gave access to the desired target molecules. The N-alkyl-1,2,5-thiadiazolidin-3-one 1,1-dioxide derivatives (±)- 3a – c were also prepared by the oxazolidinone route (Scheme 4) for coupling to a tetrapeptide fragment. The X-ray crystal structure of 1,2,5-thiadiazolidin-3-one 1,1-dioxide (±)- 21a was solved, and the directionality of the H-bond donor (N−H) and acceptor (SO2) groups of the cyclic scaffold determined (Figs. 1 and 2). The pKa value of the N−H group in (±)- 21a was determined by 1H-NMR titration as 11.9 (Fig. 3). Compounds (±)- 1a – d were shown to inhibit competition peptide binding to HLA-DR4 molecules in the single-digit millimolar concentration range.  相似文献   

15.
Lithiation of O-functionalized alkyl phenyl sulfides PhSCH2CH2CH2OR (R = Me, 1a; i-Pr, 1b; t-Bu, 1c; CPh3, 1d) with n-BuLi/tmeda in n-pentane resulted in the formation of α- and ortho-lithiated compounds [Li{CH(SPh)CH2CH2OR}(tmeda)] (α-2ad) and [Li{o-C6H4SCH2CH2CH2OR)(tmeda)] (o-2ad), respectively, which has been proved by subsequent reaction with n-Bu3SnCl yielding the requisite stannylated γ-OR-functionalized propyl phenyl sulfides n-Bu3SnCH(SPh)CH2CH2OR (α-3ad) and n-Bu3Sn(o-C6H4SCH2CH2CH2OR) (o-3ad). The α/ortho ratios were found to be dependent on the sterical demand of the substituent R. Stannylated alkyl phenyl sulfides α-3ac were found to react with n-BuLi/tmeda and n-BuLi yielding the pure α-lithiated compounds α-2ac and [Li{CH(SPh)CH2CH2OR}] (α-4ab), respectively, as white to yellowish powders. Single-crystal X-ray diffraction analysis of [Li{CH(SPh)CH2CH2Ot-Bu}(tmeda)] (α-2c) exhibited a distorted tetrahedral coordination of lithium having a chelating tmeda ligand and a C,O coordinated organyl ligand. Thus, α-2c is a typical organolithium inner complex.Lithiation of O-functionalized alkyl phenyl sulfones PhSO2CH2CH2CH2OR (R = Me, 5a; i-Pr, 5b; CPh3, 5c) with n-BuLi resulted in the exclusive formation of the α-lithiated products Li[CH(SO2Ph)CH2CH2OR] (6ac) that were found to react with n-Bu3SnCl yielding the requisite α-stannylated compounds n-Bu3SnCH(SO2Ph)CH2CH2OR (7ac). The identities of all lithium and tin compounds have been unambiguously proved by NMR spectroscopy (1H, 13C, 119Sn).  相似文献   

16.
7‐Oxabenzonorbornadienes derivatives 1 a – d underwent reductive coupling with alkyl propiolates CH3C?CCO2CH3 ( 2 a ), PhC?CCO2Et ( 2 b ), CH3(CH2)3C?CCO2CH3 ( 2 c ), CH3(CH2)4C?CCO2CH3 ( 2 d ), TMSC?CCO2Et ( 2 e ), (CH3)3C?CCO2CH3 ( 2 f ) and HC?CCO2Et ( 2 g ) in the presence of [NiBr2(dppe)] (dppe=Ph2PCH2CH2PPh2), H2O and zinc powder in acetonitrile at room temperature to afford the corresponding 2alkenyl‐1,2‐dihydronapthalen‐1‐ol derivatives 3 a – n with remarkable regio‐ and diastereoselectivity in good to excellent yields. Similarly, the reaction of 7azabenzonorbornadienes derivative 1 e with propiolates 2 a, b and d proceeded smoothly to afford reductive coupling products 2alkenyl‐1,2‐dihydronapthalene carbamates 3 o – p in good yields with high regio‐ and stereoselectivity. This nickel‐catalyzed reductive coupling can be further extended to the reaction of 7oxabenzonorbornene derivatives. Thus, 5,6‐di(methoxymethyl)‐7‐oxabicyclo[2.2.1]hept‐2‐ene ( 4 ) reacted with 2 a and 2 d to furnish cyclohexenol derivatives bearing four cis substituents 5 a and b in 81 and 84 % yield, respectively. In contrast to the results of 4 with 2 , the reaction of dimethyl 7oxabicyclo[2.2.1]hept‐5‐ene‐2,3‐dicarboxylate ( 6 ) with propiolates 2 a – d afforded the corresponding reductive coupling/cyclization products, bicyclo[3.2.1]γ‐lactones 7 a – d in good yields. The reaction provides a convenient one‐pot synthesis of γ‐lactones with remarkably high regio‐ and stereoselectivity.  相似文献   

17.
Second‐order rate constants (k1) have been measured spectrophotometrically for reactions of 2‐methoxy‐3‐X‐5‐nitrothiophene 1a‐c (X = NO2, CN, and COCH3) with secondary cyclic amines (pyrrolidine 2a , piperidine 2b , and morpholine 2 c ) in CH3CN and 91:9 (v/v) CH3OH/CH3CN at 20°C. The experimental data show that the rate constants (k1) values exhibit good correlation with the parameters of nucphilicity (N) of the amines 2a‐c and are consistent with the Mayr's relationship log k (20°C) = s(E + N). We have shown that the electrophilicity parameters E derived for 1a–c and those reported previously for the thiophenes 1d‐g (X = SO2CH3, CO2CH3, CONH2, and H) are linearly related to the pKa values for their gem‐dimethoxy complexes in methanol. Using this correlation, we successfully evaluated the electrophilicity E values of 12 structurally diverse electrophiles in methanol for the first time. In addition, a satisfactory linear correlation (r2 = 0.9726) between the experimental (log kexp) and the calculated (log kcalcd) values for the σ‐complexation reactions of these 12 electrophiles with methoxide ion in methanol has been observed and discussed.  相似文献   

18.
1. Photochlorination in CCl4 of the Si-chlorinated carbosilanes (Cl3Si? CH2)2SiCl2 and (Cl2Si? CH2)3 leads to totally chlorinated compounds, e. g. (Cl3Si? CCl2)2SiCl2. After chlorination has started at one CH2 group, formation of a CCl2 group is preferred before another CH2 group is involved into the reaction. Thus preparation of compounds a, b, c is possible. Cl3Si? CCl2? SiCl2? CH2? SiCl3 (a) for (b) and (c) (see “Inhaltsübersicht”). SO2Cl2 (benzoyl peroxide) as chlorinating agent reacts more slowly, and opens an access to carbosilanes containing CHCl groups such as (d), Cl3Si-CHCl? SiCl2? CH2? SiCl3 (e). Reactions of compounds (a) to (d) with LiAlH4 yields carbosilanes with SiH groups, and partially chlorinated C atoms. 2. By the high reactivity of Si? CCl2? Si groups an exchange of Cl atoms of CCl groups in perchlorinated carbosilanes is possible for H atoms of Si? H groups in perhydrogenated carbosilanes, thus allowing the preparation of compounds containing CHCl and SiHCl groups, e. g. according to Gl.(1) (Inhaltsübersicht). Further reactions, formulated as the last equations in Inhaltsübersicht, are reported as well as the rearrangement of H3Si? CHCl? SiH3.  相似文献   

19.
Treatment of the (butadiene)ML2 complexes 1 [ML2 = Cp2Zr ( a ), Cp2Hf ( b ), and (.-C5H4CH3)2Zr ( c )] with B(C6F5)3 gives the 1:1 addition products (CH2CHCHCH2–B(C6F5_3)ML2 ( 3a – c ). At –40°C the betaine complex 3a inserts one equivalent of methylenecyclopropane to give the regioisomeric insertion products 5a and 6a in a 60:40 ratio. These products exhibit the cyclopropylidene moiety in the α- and β-positions, respectively, relative to zirconium. The corresponding hafnocene complexes 5b and 6b are obtained in a 70:30 ratio starting from 3b . The reaction of 3 ( a – c ) with allene gives a single insertion product ( 7a – c ) in each case where the exo-methylene group is in the α-position to the metal center ([2,1]-insertion). The complexes 5 – 7 are chiral. They all exhibit a pronounced ·-interaction of the internal –C4H=C5H double bond of the s̀-ligand chain with the metal center in addition to a metallocene/–C6H2–[B] ion pair interaction. The relative contributions of the cationic metallocene end of the dipolar complexes 5 – 7 are quite dependent on the steric and electronic properties of the respective metallocene units involved. This is revealed by a comparison to typical 13C-NMR parameters of the complexes 5 – 7 with a pair of suitable model complexes, namely the ethylene insertion product 4 into the betaine system 3a and its THF adduct 4 .THF.  相似文献   

20.
Thermolysis of [Cp*Ru(PPh2(CH2)PPh2)BH2(L2)] 1 (Cp*=η5‐C5Me5; L=C7H4NS2), with terminal alkynes led to the formation of η4‐σ,π‐borataallyl complexes [Cp*Ru(μ‐H)B{R‐C=CH2}(L)2] ( 2 a – c ) and η2‐vinylborane complexes [Cp*Ru(R‐C=CH2)BH(L)2] ( 3 a – c ) ( 2 a , 3 a : R=Ph; 2 b , 3 b : R=COOCH3; 2 c , 3 c : R=p‐CH3‐C6H4; L=C7H4NS2) through hydroboration reaction. Ruthenium and the HBCC unit of the vinylborane moiety in 2 a – c are linked by a unique η4‐interaction. Conversions of 1 into 3 a – c proceed through the formation of intermediates 2 a – c . Furthermore, in an attempt to expand the library of these novel complexes, chemistry of σ‐borane complex [Cp*RuCO(μ‐H)BH2L] 4 (L=C7H4NS2) was investigated with both internal and terminal alkynes. Interestingly, under photolytic conditions, 4 reacts with methyl propiolate to generate the η4‐σ,π‐borataallyl complexes [Cp*Ru(μ‐H)BH{R‐C=CH2}(L)] 5 and [Cp*Ru(μ‐H)BH{HC=CH‐R}(L)] 6 (R=COOCH3; L=C7H4NS2) by Markovnikov and anti‐Markovnikov hydroboration. In an extension, photolysis of 4 in the presence of dimethyl acetylenedicarboxylate yielded η4‐σ,π‐borataallyl complex [Cp*Ru(μ‐H)BH{R‐C=CH‐R}(L)] 7 (R=COOCH3; L=C7H4NS2). An agostic interaction was also found to be present in 2 a – c and 5 – 7 , which is rare among the borataallyl complexes. All the new compounds have been characterized in solution by IR, 1H, 11B, 13C NMR spectroscopy, mass spectrometry and the structural types were unequivocally established by crystallographic analysis of 2 b , 3 a – c and 5 – 7 . DFT calculations were performed to evaluate possible bonding and electronic structures of the new compounds.  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