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.
The stereoselective direct transformation of N‐(propargylic)hydroxylamines into cis‐2‐acylaziridines was achieved by the combined use of AgBF4 and CuCl. Copper salts were found to promote the transformation of the intermediary 4‐isoxazolines into 2‐acylaziridines and both 3‐aryl‐ and 3‐alkyl‐substituted 2‐acylaziridines could be prepared by using this method. Furthermore, subsequent 1,3‐dipolar cycloaddition of azomethine ylides that were generated in situ from the intermediary 2‐acylaziridines with maleimides was achieved in a stereoselective one‐pot procedure to afford the corresponding 2‐acylpyrrolidines, which consisted of an octahydropyrrolo[3,4‐c]pyrrole skeleton.  相似文献   

2.
The first stable copper borohydride complex [(CAAC)CuBH4] [CAAC=cyclic(alkyl)(amino)carbene] bearing a single monodentate ligand was prepared by addition of NaBH4 or BH3NH3 to the corresponding [(CAAC)CuCl] complex. Both complexes are air‐stable and promote the catalytic hydrolytic dehydrogenation of ammonia borane. The amount of hydrogen released reaches 2.8 H2/BH3NH3 with a turnover frequency of 8400 mol molcat?1 h?1 at 25 °C. In a fifteen‐cycle experiment, the catalyst was reused without any loss of efficiency.  相似文献   

3.
A series of mono- and dicarbene gold(I) complexes of types Au(CAAC)(Cl) [CAAC = cyclic (alkyl)(amino)carbene] (1) and [Au(CAAC)2]+[X]? (X = Cl, AuCl2) (2) have been prepared through reaction of AuCl(SMe2) with free carbenes ae, and structurally characterized by single X-ray diffraction studies (1a, 1b, 2d, 2e). In addition two new free cyclic (alkyl)(amino)carbenes (c and e) have been synthesized.  相似文献   

4.
A stable cyclic (alkyl)(amino)carbene (CAAC) 1 inserts into the para‐CF bond of pentafluoropyridine, and after fluoride abstraction, the iminium‐pyridyl adduct [ 3 ]+ was isolated. A cyclic voltammetry study shows a reversible three‐state redox system involving [ 3 ]+, [ 3 ] ? , and [ 3 ] ? . The CAAC‐pyridyl radical [ 3 ] ? , obtained by reduction of [ 3 ]+ with magnesium, has been spectroscopically and crystallographically characterized. In contrast to the lack of π communication between the CAAC and the pyridine units in cation [ 3 ]+, the unpaired electron of [ 3 ] ? is delocalized over an extended π system involving both heterocycles.  相似文献   

5.
Reactions between [Mn(CO)5Br] and dpkbh in low boiling solvents in air gave fac-[MnI(CO)32-Npy,Nim-dpkbh)Br]·H2O, [MnIIBr23-Npy,Nim,O-dpkbh)], and [MnII3-Npy,Nim,O-dpkbh-H)2]·0.5H2O (Nim = imine nitrogen and Npy = pyridyl nitrogen). Crystallization of fac-[MnI(CO)32-Npy,Nim-dpkbh)Br]·H2O from dmso or CH3CN produced dark red crystals of [MnII3-Npy,Nim,O-dpkbh-H)2]·nX (X = dmso, n = 1 and X = H2O, n = 0.22). This is in contrast to the reaction of [Re(CO)5Cl] with dpkbh in refluxing toluene to form fac-[ReI(CO)32-,Npy,Npy-dpkbh)Cl] which can be crystallized from CH3CN, dmso or dmf to form fac-[ReI(CO)32-,Npy,Npy-dpkbh)Cl]·nX (X = CH3CN, n = 0 and solvate = dmso or dmf, n = 1). Infrared spectral measurements are consistent with keto coordination of dpkbh to Mn(I) in fac-[MnI(CO)32-Npy,Nim-dpkbh)Br]·H2O and Mn(II) in [MnIIBr23-Npy,Nim,O-dpkbh)] plus enol coordination of the amide-deprotonated dpkbh, to the Mn(II) center in [MnII3-Npy,Nim,O-dpkbh-H)2]·0.5H2O. Electronic absorption spectral measurements in non-aqueous solvents indicate sensitivity of fac-[MnI(CO)32-Npy,Nim-dpkbh)Br]·H2O and [MnII3-Npy,Nim,O-dpkbh-H)2]·0.5H2O to changes in their outer-shell environments. X-ray crystallographic analyses elucidated the identities of [MnIIBr23-Npy,Nim,O-dpkbh)] and [MnII3-Npy,Nim,O-dpkbh-H)2]·nX and divulged weaker coordination of [dpkbh] to Mn(II) in [MnIIBr23-Npy,Nim,O-dpkbh)] and stronger coordination of [dpkbh-H]? to Mn(II) in [MnII3-Npy,Nim,O-dpkbh-H)2]·0.22H2O. Low-temperature X-ray structural analyses were employed to account for the disorder in the structure of [MnII3-Npy,Nim,O-dpkbh-H)2] and the short NH bond distance observed in the structure of [MnIIBr23-Npy,Nim,O-dpkbh)]. A PLATON Squeeze treatment was invoked to account for the fractional occupancy of lattice water in the structure of [MnII3-Npy,Nim,O-dpkbh-H)2].  相似文献   

6.
The reactivity of ZnII dialkyl species ZnMe2 with a cyclic(alkyl)(amino)carbene, 1-[2,6-bis(1-methylethyl)phenyl]-3,3,5,5-tetramethyl-2-pyrrolidinylidene (CAAC, 1 ), was studied and extended to the preparation of robust CAAC-supported ZnII Lewis acidic organocations. CAAC adduct of ZnMe2 ( 2 ), formed from a 1:1 mixture of 1 and ZnMe2, is unstable at room temperature and readily undergoes a CAAC carbene insertion into the Zn−Me bond to produce the ZnX2-type species (CAAC-Me)ZnMe ( 3 ), a reactivity further supported by DFT calculations. Despite its limited stability, adduct 2 was cleanly ionized to robust two-coordinate (CAAC)ZnMe+ cation ( 5+ ) and derived into (CAAC)ZnC6F5+ ( 7+ ), both isolated as B(C6F5)4 salts, showing the ability of CAAC for the stabilization of reactive [ZnMe]+ and [ZnC6F5]+ moieties. Due to the lability of the CAAC−ZnMe2 bond, the formation of bis(CAAC) adduct (CAAC)2ZnMe+ cation ( 6+ ) was also observed and the corresponding salt [ 6 ][B(C6F5)4] was structurally characterized. As estimated from experimental and calculations data, cations 5+ and 7+ are highly Lewis acidic species and the stronger Lewis acid 7+ effectively mediates alkene, alkyne and CO2 hydrosilylation catalysis. All supporting data hints at Lewis acid type activation–functionalization processes. Despite a lower energy LUMO in 5+ and 7+ , their observed reactivity is comparable to those of N-heterocyclic carbene (NHC) analogues, in line with charge-controlled reactions for carbene-stabilized ZnII organocations.  相似文献   

7.
The two‐coordinate [(CAAC)2Fe] complex [CAAC=cyclic (alkyl)(amino)carbene] binds dinitrogen at low temperature (T2 complex, [(CAAC)2Fe(N2)], was trapped by one‐electron reduction to its corresponding anion [(CAAC)2FeN2]? at low temperature. This complex was structurally characterized and features an activated dinitrogen unit which can be silylated at the β‐nitrogen atom. The redox‐linked complexes [(CAAC)2FeI][BArF4], [(CAAC)2Fe0], and [(CAAC)2Fe?IN2]? were all found to be active for the reduction of dinitrogen to ammonia upon treatment with KC8 and HBArF4?2 Et2O at ?95 °C [up to (3.4±1.0) equivalents of ammonia per Fe center]. The N2 reduction activity is highly temperature dependent, with significant N2 reduction to NH3 only occurring below ?78 °C. This reactivity profile tracks with the low temperatures needed for N2 binding and an otherwise unavailable electron‐transfer step to generate reactive [(CAAC)2FeN2]?.  相似文献   

8.
A stereoselective method for the synthesis of 2′,5′-disubstituted N-arylpyrrolofullerenes from anilines, alkyl glyoxylates, alkyl diazoacetates, and fullerene C60 was proposed. The key step of the synthesis is the 1,3-dipolar cycloaddition reaction of fullerene with azomethine ylides generated by heating of dialkyl aziridinedicarboxylates. The thermal opening of the aziridine ring to azomethine ylide and the cycloaddition of the latter to C60 at 100 °C are nearly completely stereoselective: only trans-adducts are formed from cis-aziridines, whereas trans-aziridines give exclusively cis-adducts.  相似文献   

9.
The synthesis, structure, and ligand substitution mechanism of a new five-coordinate trigonal-bipyramidal copper(II) complex, [CuII(py tBuMe2N3)Cl2] (1), with a sterically constrained py tBuMe2N3 chelate ligand, py tBuMe2N3?=?2,6-bis-(ketimino)pyridyl, are reported. The kinetics and mechanism of chloride substitution by thiourea, as a function of nucleophile concentration, temperature, and pressure, were studied in detail and compared with an earlier study reported for the analogous complex [CuII(py tBuN3)Cl2] (2) [py tBuN3?=?2,6-bis-(aldimino)pyridyl]. Catalysis of the oxidation of 3,5-di-tert-butylcatechol to 3,5-di-tert-butylquinone by 1 and 2 was studied. Correlations between the reactivity, chloride substitution behavior, and reduction potentials of both complexes were made. These show that the rate of oxidation is independent of the rate of chloride substitution, indicating that the substitution of chloride by catechol as substrate occurs in a fast step. Spectral data show a non-linear relationship between the ability of the complexes to oxidize 3,5-DTBC and the Lewis acidity of their copper(II) centers. Electrochemical data demonstrate that the most effective complex 1 has a E 0 value that approaches the E 0 value of the natural tyrosinase enzyme.  相似文献   

10.
Reaction of CuI with 1 or 2 equivalent(s) N,N′‐Bis(diphenylphosphino)‐2,6‐diaminopyridine (BDDP) gives two different complexes, [Cu(I)μ‐(BDDP‐κP,Npy)]2 ( 1 ) and [Cu(BDDP‐κP,Npy)2]I ( 2 ), in high yields. The determination of the molecular structure show that both CuI atoms are tetrahedrally coordinated, rather than a square‐planar geometry reported for Cr0, NiII‐BDDP complexes before, which contains a planar tridentate chelate ring system. The introduction of AuCl(tht) (tht = tetrahydrothiophene) into [Cu(BDDP‐κP,Npy)2]I leads unexpectedly to the formation of a digold complex 2,6‐[(ClAuPh2P)HN]2C5H3N and dimeric [Cu(I)μ‐(BDDP‐κP,Npy)]2.  相似文献   

11.
In neutral and acidic solution, homonuclear s-block metal complexes [Sr2(1,3-pdta)(H2O)3.5] n (1), {[Ba2(1,3-pdta)(H2O)6]·H2O} n (2), {[Sr(1,3-H2pdta)]·(H2O)} n (3), and [Ba(1,3-H2pdta)(H2O)3] n (4) {1,3-H4pdta=1,3-propanediamine-N,N,N′,N′-tetraacetic acid, CH2[CH2N(CH2CO2H)2]2} were isolated. In 1 and 2, hexadentate 1,3-pdta joins two metal ions via the diamine chain. In 3 and 4, the nitrogens of 1,3-H2pdta were protonated and show no coordination. There is no coordinated water in 3, unusual coordination for strontium. In 4, the coordination number is nine and there are three coordinated waters for each barium. One carboxy group of pdta is free without coordination. Thermal decomposition shows that temperatures of ligand elimination start at 408, 423, 298, and 250?°C for 1–4, respectively. Acidic condition is favorable for preparation of metal oxide at low temperature.  相似文献   

12.
In the title compound, [RuCl2(C2H3N)(C27H31N3)]·CH2Cl2, the RuII ion is six‐coordinated in a distorted octahedral arrangement, with the two Cl atoms located in the apical positions, and the pyridine (py) N atom, the two imino N atoms and the acetonitrile N atom located in the basal plane. The two equatorial Ru—Nimino distances are almost equal (mean 2.087 Å) and are substantially longer than the equatorial Ru—Npy bond [1.921 (4) Å]. It is observed that the NiminoM—Npy angle for the five‐membered chelate rings of pyridine‐2,6‐diimine complexes is inversely related to the magnitude of the M—Npy bond. The title structure is stabilized by intra‐ and intermolecular C—H...Cl hydrogen bonds, as well as by van der Waals interactions.  相似文献   

13.
The CAAC [CAAC=cyclic (alkyl)(amino)carbene] family of carbene ligands have shown promise in stabilizing unusually low‐coordination number transition‐metal complexes in low formal oxidation states. Here we extend this narrative by demonstrating their utility in affording access to the first examples of two‐coordinate formal Fe0 and Co0 [(CAAC)2M] complexes, prepared by reduction of their corresponding two‐coordinate cationic FeI and CoI precursors. The stability of these species arises from the strong σ‐donating and π‐accepting properties of the supporting CAAC ligands, in addition to steric protection.  相似文献   

14.
The general strategies to stabilize a boryl radical involve single electron delocalization by π-system and the steric hinderance from bulky groups. Herein, a new class of boryl radicals is reported, with intramolecular mixed-valent B(III)Br-B(II) adducts ligated by a cyclic (alkyl)(amino)carbene (CAAC). The radicals feature a large spin density on the boron center, which is ascertained by EPR spectroscopy and DFT calculations. Structural and computational analyses revealed that the stability of radical species was assisted by the CAAC ligand and a weak but significant B(III)Br-B(II) interaction, suggesting a cooperative avenue for stabilization of boryl radicals. Two-electron reduction of these new boryl radicals provides C−H insertion products via a borylene intermediate.  相似文献   

15.
An improved synthetic route to homoleptic complex [Pt(CAACMe)2] (CAAC=cyclic (alkyl)(amino)carbenes) and convenient routes to new heteroleptic complexes of the form [Pt(CAACMe)(PR3)] are presented. Although the homoleptic complex was found to be inert to many reagents, oxidative addition and metal‐only Lewis pair (MOLP) formation was observed from one of the heteroleptic complexes. The spectroscopic, structural, and electrochemical properties of the zero‐valent complexes were explored in concert with density functional theory (DFT) and time‐dependent density functional theory (TD‐DFT) calculations. The homoleptic [Pt(CAAC)2] and heteroleptic [Pt(CAAC)(PR3)] complexes were found to be similar in their spectroscopic and structural properties, but their electrochemical behavior and reactivity differ greatly. The unusually strong color of the CAAC‐containing Pt0 complexes was investigated by TD‐DFT calculations and attributed to excitations into the LUMOs of the complexes, which are predominantly composed of bonding π interactions between Pt and the CAAC carbon atoms.  相似文献   

16.
Kinetics and mechanism of the reactions of methyl diazoacetate, dimethyl diazomalonate, 4-nitrophenyldiazomethane, and diphenyldiazomethane with sulfonium ylides and enamines were investigated by UV-Vis and NMR spectroscopy. Ordinary alkenes undergo 1,3-dipolar cycloadditions with these diazo compounds. In contrast, sulfonium ylides and enamines attack at the terminal nitrogen of the diazo alkanes to give zwitterions, which undergo various subsequent reactions. As only one new bond is formed in the rate-determining step of these reactions, the correlation lg k2(20 °C)=sN(N+E) could be used to determine the one-bond electrophilicities E of the diazo compounds from the measured second-order rate constants and the known reactivity indices N and sN of the sulfonium ylides and enamines. The resulting electrophilicity parameters (−21<E<−18), which are 11–14 orders of magnitude smaller than that of the benzenediazonium ion, are used to define the scope of one-bond nucleophiles which may react with these diazoalkanes.  相似文献   

17.
In the title compound, [RuCl2(C2H3N)(C27H31N3)]·0.5CH2Cl2, the RuII ion is six‐coordinated in a distorted octa­hedral arrangement, with the two Cl atoms located in the apical positions, and the pyridine (py) N atom, the two imino N atoms and the acetonitrile N atom located in the basal plane. The dichloromethane solvent mol­ecule lies on a twofold axis. The two equatorial Ru—Nimino distances are almost equal (mean 2.089 Å) and are substantially longer than the equatorial Ru—Npy bond [1.914 (4) Å]. It is observed that the NiminoM—Npy bond angle for the five‐membered chelate rings of pyridine‐2,6‐diimine complexes is inversely related to the magnitude of the M—Npy bond. The title structure is stabilized by intra‐ and inter­molecular C—H⋯Cl hydrogen bonds. The inter­molecular hydrogen bonds form an R66(24) ring and a chain of edge‐fused rings running parallel to the [001] direction.  相似文献   

18.
The kinetics of the reactions of ethenesulfonyl fluoride (ESF) with sulfonium and pyridinium ylides were measured photometrically to determine the electrophilicity parameter of ESF according to the correlation lg k20 °C=sN(N+E). With E=?12.09, ESF is among the strongest Michael acceptors in our comprehensive electrophilicity scale, which explains its excellent performance in reactions with many nucleophiles. Its predicted usability as a reagent in electrophilic aromatic substitutions with electron‐rich arenes was confirmed by uncatalyzed reactions with alkyl‐substituted pyrroles.  相似文献   

19.
The four isomers 2,4‐, (I), 2,5‐, (II), 3,4‐, (III), and 3,5‐difluoro‐N‐(3‐pyridyl)benzamide, (IV), all with formula C12H8F2N2O, display molecular similarity, with interplanar angles between the C6/C5N rings ranging from 2.94 (11)° in (IV) to 4.48 (18)° in (I), although the amide group is twisted from either plane by 18.0 (2)–27.3 (3)°. Compounds (I) and (II) are isostructural but are not isomorphous. Intermolecular N—H...O=C interactions form one‐dimensional C(4) chains along [010]. The only other significant interaction is C—H...F. The pyridyl (py) N atom does not participate in hydrogen bonding; the closest H...Npy contact is 2.71 Å in (I) and 2.69 Å in (II). Packing of pairs of one‐dimensional chains in a herring‐bone fashion occurs viaπ‐stacking interactions. Compounds (III) and (IV) are essentially isomorphous (their a and b unit‐cell lengths differ by 9%, due mainly to 3,4‐F2 and 3,5‐F2 substitution patterns in the arene ring) and are quasi‐isostructural. In (III), benzene rotational disorder is present, with the meta F atom occupying both 3‐ and 5‐F positions with site occupancies of 0.809 (4) and 0.191 (4), respectively. The N—H...Npy intermolecular interactions dominate as C(5) chains in tandem with C—H...Npy interactions. C—H...O=C interactions form R22(8) rings about inversion centres, and there are π–π stacks about inversion centres, all combining to form a three‐dimensional network. By contrast, (IV) has no strong hydrogen bonds; the N—H...Npy interaction is 0.3 Å longer than in (III). The carbonyl O atom participates only in weak interactions and is surrounded in a square‐pyramidal contact geometry with two intramolecular and three intermolecular C—H...O=C interactions. Compounds (III) and (IV) are interesting examples of two isomers with similar unit‐cell parameters and gross packing but which display quite different intermolecular interactions at the primary level due to subtle packing differences at the atom/group/ring level arising from differences in the peripheral ring‐substitution patterns.  相似文献   

20.
Two iron(II) complexes, [FeII(pytBuN3)2](FeCl4) (1) and [FeII(pytBuMe2N3)Cl2] (2), with sterically constrained pytBuN3 and pytBuMe2N3 chelate ligands (pytBuN3 = 2,6-bis-(aldiimino)pyridyl; pytBuMe2N3 = 2,6-bis-(ketimino)pyridyl), have been synthesized and characterized by elemental analysis, IR, UV–vis spectra, and preliminary X-ray single-crystal diffraction. The latter revealed that Fe(II) in 1 is six-coordinate by six nitrogen donors from two bisiminopyridines in a distorted octahedron. Complex 2 reacts with thiourea with a second-order rate constant k2 = (2.50 ± 0.05) × 10?3 M?1 s?1 at 296 K, and the reaction seemed to be slow. In a similar way, the interaction of 2 and DNA was studied by fluorescence and absorption spectroscopy. The results revealed that 2 caused fluorescence quenching of DNA through a dynamic quenching procedure. The binding constants KA, Kapp, and KSV as well as the number of binding sites between 2 and DNA were determined.  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