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 447 毫秒
1.
Reaction of cyclooctatetraene (COT) iron(II) tricarbonyl, [Fe(cot)(CO)3], with one equivalent of K4Ge9 in ethylenediamine (en) yielded the cluster anion [Ge8Fe(CO)3]3? which was crystallographically‐characterized as a [K(2,2,2‐crypt)]+ salt in [K(2,2,2‐crypt)]3[Ge8Fe(CO)3]. The chemically‐reduced organometallic species [Fe(η3‐C8H8)(CO)3]? was also isolated as a side‐product from this reaction as [K(2,2,2‐crypt)][Fe(η3‐C8H8)(CO)3]. Both species were further characterized by EPR and IR spectroscopy and electrospray mass spectrometry. The [Ge8Fe(CO)3]3? cluster anion represents an unprecedented functionalized germanium Zintl anion in which the nine‐atom precursor cluster has lost a vertex, which has been replaced by a transition‐metal moiety.  相似文献   

2.
The endohedral stannaspherene cluster anion [Ir@Sn12]3? was synthesized in two steps. The reaction of K4Sn9 with [IrCl(cod)]2 (cod: 1,5‐cyclooctadienyl) in ethylenediamine (en) solution first yielded the [K(2,2,2‐crypt)]+ salt (2,2,2‐crypt: 4,7,13,16,21,24‐hexaoxa‐1,10‐diazabicyclo[8.8.8]hexacosane) of the capped cluster anion [Sn9Ir(cod)]3?. Subsequently, crystals of this compound were dissolved in en, followed by the addition of triphenylphosphine or 1,2‐bis(diphenylphosphino)ethane and treatment at elevated temperatures. [Ir@Sn12]3? was obtained and characterized as the [K(2,2,2‐crypt)]+ salt. The isolation of [Sn9Ir(cod)]3? as an intermediate product establishes that the formation of the stannaspherene [Ir@Sn12]3? occurs through the oxidation of [Sn9Ir(cod)]3?. Among the structurally characterized tetrel cluster anions, [Ir@Sn12]3? is a unique example of a stannaspherene, and one of the rare spherical clusters encapsulating a metal atom that is not a member of Group 10. Single‐crystal structure determination shows that the novel Zintl ion cluster has nearly perfect icosahedral Ih point symmetry.  相似文献   

3.
We report on the structures of three unprecedented heteroleptic Sb‐centered radicals [L(Cl)Ga](R)Sb. ( 2‐R , R=B[N(Dip)CH]2 2‐B , 2,6‐Mes2C6H3 2‐C , N(SiMe3)Dip 2‐N ) stabilized by one electropositive metal fragment [L(Cl)Ga] (L=HC[C(Me)N(Dip)]2, Dip=2,6‐i‐Pr2C6H3) and one bulky B‐ ( 2‐B ), C‐ ( 2‐C ), or N‐based ( 2‐N ) substituent. Compounds 2‐R are predominantly metal‐centered radicals. Their electronic properties are largely influenced by the electronic nature of the ligands R, and significant delocalization of unpaired‐spin density onto the ligands was observed in 2‐B and 2‐N . Cyclic voltammetry (CV) studies showed that 2‐B undergoes a quasi‐reversible one‐electron reduction, which was confirmed by the synthesis of [K([2.2.2]crypt)][L(Cl)GaSbB[N(Dip)CH]2] ([K([2.2.2]crypt)][ 2‐B ]) containing the stibanyl anion [ 2‐B ]?, which was shown to possess significant Sb?B multiple‐bonding character.  相似文献   

4.
The reactions of the Zintl phase K2Cs2Sn9 with elemental tellurium and selenium in ethylenediamine have been investigated. From the reaction of K2Cs2Sn9 with elemental tellurium [K‐(2,2,2‐crypt)]4Te6Te4 ( 2 ) and [K‐(2,2,2‐crypt)]2Sn2Te3 ( 3 ) were obtained, whereas the reaction of K2Cs2Sn9 with elemental selenium led to the formation of [K‐(2,2,2‐crypt)]2Sn(Se4)3 ( 4 ) and [K‐(2,2,2‐crypt)]2Cs2Sn2Se6·2en ( 5 )1). Compounds 2 , 4 , 5 have been characterized by single crystal X‐ray structure determination.  相似文献   

5.
This work reports zinc‐catalyzed [4+2]‐annulation reactions of disubstituted N‐hydroxy allenylamines with nitrosoarenes to afford substituted 1,2‐oxazinan‐3‐ones with a skeletal rearrangement. This annulation is applicable to a reasonable scope of allenylamines and nitrosoarenes. Our control experiments indicate that nitrosobenzene can also implement this annulation through a radical annulation path, but with poor efficiency. Zn(OTf)2 or AgOTf greatly improves the efficiency of this [4+2]‐annulation; the effect of these metal species is discussed in detail.  相似文献   

6.
Black crystals of [Rb(crypt‐2,2,2)]4(I5)2(I8) were obtained from a dichloromethane/ethanol solution of RbI, I2 and Kryptofix‐2,2,2®. The crystal structure (monoclinic, P21/c (no. 14), a = 1250.1(1), b = 2555.2(2), c = 2313.4(3) pm, β = 121.45(1)°, V = 6309.9(11)·106 pm3, Z = 2) consists of [Rb(crypt‐2,2,2)]+ cations leaving three‐dimensional channels for the V‐shaped (I5)? and Z‐shaped (I8)2? anions which are isolated from each other.  相似文献   

7.
The synthesis of new molecular complexes of U2+ has been pursued to make comparisons in structure, physical properties, and reactivity with the first U2+ complex, [K(2.2.2‐cryptand)][Cp′3U], 1 (Cp′=C5H4SiMe3). Reduction of Cp′′3U [Cp′′=C5H3(SiMe3)2] with KC8 in the presence of 2.2.2‐cryptand or 18‐crown‐6 generates [K(2.2.2‐cryptand)][Cp′′3U], 2‐K(crypt) , or [K(18‐crown‐6)(THF)2][Cp′′3U], 2‐K(18c6) , respectively. The UV/Vis spectra of 2‐K and 1 are similar, and they are much more intense than those of U3+ analogues. Variable temperature magnetic susceptibility data for 1 and 2‐K(crypt) reveal lower room temperature χMT values relative to the experimental values for the 5f3 U3+ precursors. Stability studies monitored by UV/Vis spectroscopy show that 2‐K(crypt) and 2‐K(18c6) have t1/2 values of 20 and 15 h at room temperature, respectively, vs. 1.5 h for 1 . Complex 2‐K(18c6) reacts with H2 or PhSiH3 to form the uranium hydride, [K(18‐crown‐6)(THF)2][Cp′′3UH], 3 . Complexes 1 and 2‐K(18c6) both reduce cyclooctatetraene to form uranocene, (C8H8)2U, as well as the U3+ byproducts [K(2.2.2‐cryptand)][Cp′4U], 4 , and Cp′′3U, respectively.  相似文献   

8.
The synthesis and characterization of two bimetallic, cationic low‐valent gallium–cryptand[2.2.2] complexes is reported. The reaction of cryptand[2.2.2] with Ga2Cl4 gave two different cations, [Ga3Cl4(crypt‐222)]+ ( 1 ) or [Ga2Cl2(crypt‐222)]2+ ( 2 ), depending on whether or not trimethylsilyl triflate (Me3SiOTf) was added as a co‐reagent. Complexes 1 and 2 are the first examples of bimetallic cryptand[2.2.2] complexes, as well as the first low‐valent gallium–cryptand[2.2.2] complexes. Computational methods were used to evaluate the bonding in the gallium cores.  相似文献   

9.
Lanthanide triflates have been used to incorporate NdIII and SmIII ions into the 2.2.2‐cryptand ligand (crypt) to explore their reductive chemistry. The Ln(OTf)3 complexes (Ln=Nd, Sm; OTf=SO3CF3) react with crypt in THF to form the THF‐soluble complexes [LnIII(crypt)(OTf)2][OTf] with two triflates bound to the metal encapsulated in the crypt. Reduction of these LnIII‐in‐crypt complexes using KC8 in THF forms the neutral LnII‐in‐crypt triflate complexes [LnII(crypt)(OTf)2]. DFT calculations on [NdII(crypt)]2+], the first NdII cryptand complex, assign a 4f4 electron configuration to this ion.  相似文献   

10.
Details of the direct synthesis of cationic Ru(II)(η5‐Cp)(η6‐arene) complexes from ruthenocene using microwave heating are reported. Developed for the important catalyst precursor [Ru(II)(η5‐Cp)(η6‐1‐4,4a,8a‐naphthalene)][PF6] reaction time could be shortened from three days to 15 min. The method was extended to [Ru(II)(η6‐benzene)(η5‐Cp)][PF6], [Ru(II)(η5‐Cp)(η6‐toluene)][PF6], [Ru(II)(η5‐Cp)(η6‐mesitylene)][PF6], [Ru(II)(η5‐Cp)(η6‐hexamethylbenzene)][PF6], [Ru(II)(η5Cp)(η6‐indane)][PF6], [Ru(II)(η5‐Cp)(η6‐2,6‐dimethylnaphthalene)][PF6], and [Ru(II)(η5‐Cp)(η6‐pyrene)][PF6]. 1‐methylnaphthalene and 2,3‐dimethylnaphthalene afforded mixtures of regioisomeric complexes. [Ru(Cp)(CH3CN)3][PF6], derived from the naphthalene precursor provided access to the cationic RuCp complexes of naphthoquinone, tetralindione, 1,4‐dihydroxynaphthalene, and 1,4‐dimethoxynaphthalene. Reduction of the tetralindione complex afforded selectively the endo,endo diol derivative. X‐Ray structures of five complexes are reported.  相似文献   

11.
This work reports the first success of the nitroso‐Povarov reaction, involving gold‐catalyzed [4+2] annulations of nitrsoarenes with substituted cyclopentadienes. In this catalytic sequence, nitrosoarenes presumably attack gold‐π‐dienes by a 1,4‐addition pathway, generating allylgold nitrosonium intermediates to complete an intramolecular cyclization. Acyclic dienes are also applicable substrates, and affording oxidative nitroso‐Povarov products.  相似文献   

12.
The first crystallographically characterizable complex of Sc2+, [Sc(NR2)3] (R=SiMe3), has been obtained by LnA3/M reactions (Ln=rare earth metal; A=anionic ligand; M=alkali metal) involving reduction of Sc(NR2)3 with K in the presence of 2.2.2‐cryptand (crypt) and 18‐crown‐6 (18‐c‐6) and with Cs in the presence of crypt. Dark maroon [K(crypt)]+, [K(18‐c‐6)]+, and [Cs(crypt)]+ salts of the [Sc(NR2)3] anion are formed, respectively. The formation of this oxidation state of Sc is also indicated by the eight‐line EPR spectra arising from the I =7/2 45Sc nucleus. The Sc(NR2)3 reduction differs from Ln(NR2)3 reactions (Ln=Y and lanthanides) in that it occurs under N2 without formation of isolable reduced dinitrogen species. [K(18‐c‐6)][Sc(NR2)3] reacts with CO2 to produce an oxalate complex, {K2(18‐c‐6)3}{[(R2N)3Sc]2(μ‐C2O4κ 1O:κ 1O′′)}, and a CO2 radical anion complex, [(R2N)3Sc(μ‐OCO‐κ 1O:κ 1O′)K(18‐c‐6)]n .  相似文献   

13.
N‐Nitroso compounds containing benzene, fluorene or fluorenone rings were synthesized. Photolysis of these compounds with 312‐nm UV light provided the NO . species, the presence of which was corroborated by use of an EPR method and of 2‐phenyl‐4,4,5,5‐tetramethylimidazolin‐1‐oxyl 3‐oxide (PTIO) as a trapping agent. During irradiation of N‐methyl‐N‐nitroso‐9‐fluorenone carboxamide ( 14 c ) in the absence of PTIO, it underwent decomposition followed by recombination to give the heterocyclic nitric oxide radical 15 . Incorporation of intercalating moieties endowed the N‐nitroso compounds with DNA‐cleaving ability through single‐strand scission upon UV irradiation in a phosphate buffer (pH 5.0–8.0) under aerobic conditions.  相似文献   

14.
Bis[(4,7,13,16,21,24‐hexaoxa‐1,10‐diazabicyclo[8.8.8]hexacosane)potassium(+)] pentacarbonylchromate(2−) ethylenediamine monosolvate, [K(C18H36N2O6)]2[Cr(CO)5]·C2H8N2, was obtained from the reaction between K3Cd2Sb2 and Cr(CO)6 in ethylenediamine in the presence of the macrocyclic 2,2,2‐crypt ligand. The structure provides the first crystallographic characterization of the pentacoordinated [Cr(CO)5]2− dianion. The central CrIII atom is coordinated by five carbonyl ligands in a distorted trigonal–bipyramidal geometry. The distribution of the Cr—C bond lengths indicates a greater degree of back bonding from CrIII to the equatorial carbonyl ligands compared with the axial carbonyl ligands.  相似文献   

15.
The title compound [K([2,2,2]crypt)]12[Sn9]2[Sn9HgSn9] has been obtained by reaction of elemental mercury with the binary phase K4Sn9 in ethylenediamine after addition of [2,2,2]crypt and layering with toluene. The X‐ray single crystal analysis shows that the compound consists of two isolated Sn9 clusters and two Sn9 clusters connected by a mercury atom.  相似文献   

16.
It is possible that fluorous compounds could be utilized as directing forces in crystal engineering for applications in materials chemistry or catalysis. Although numerous fluorous compounds have been used for various applications, their structures in the solid state remains a lively matter for debate. The reaction of 4‐[(2,2,2‐trifluoroethoxy)methyl]pyridine with HX (X = I or Cl) yielded new fluorous ponytailed pyridinium halide salts, namely 4‐[(2,2,2‐trifluoroethoxy)methyl]pyridinium iodide, C8H9F3NO+·I, (1), and 4‐[(2,2,2‐trifluoroethoxy)methyl]pyridinium chloride, C8H9F3NO+·Cl, (2), which were characterized by IR spectroscopy, multinuclei (1H, 13C and 19F) NMR spectroscopy and single‐crystal X‐ray diffraction. Structure analysis showed that there are two types of hydrogen bonds, namely N—H…X and C—H…X. The iodide anion in salt (1) is hydrogen bonded to three 4‐[(2,2,2‐trifluoroethoxy)methyl]pyridinium cations in the crystal packing, while the chloride ion in salt (2) is involved in six hydrogen bonds to five 4‐[(2,2,2‐trifluoroethoxy)methyl]pyridinium cations, which is attributed to the smaller size and reduced polarizability of the chloride ion compared to the iodide ion. In the IR spectra, the pyridinium N—H stretching band for salt (1) exhibited a blue shift compared with that of salt (2).  相似文献   

17.
We report on the structures of three unprecedented heteroleptic Sb-centered radicals [L(Cl)Ga](R)Sb. ( 2-R , R=B[N(Dip)CH]2 2-B , 2,6-Mes2C6H3 2-C , N(SiMe3)Dip 2-N ) stabilized by one electropositive metal fragment [L(Cl)Ga] (L=HC[C(Me)N(Dip)]2, Dip=2,6-i-Pr2C6H3) and one bulky B- ( 2-B ), C- ( 2-C ), or N-based ( 2-N ) substituent. Compounds 2-R are predominantly metal-centered radicals. Their electronic properties are largely influenced by the electronic nature of the ligands R, and significant delocalization of unpaired-spin density onto the ligands was observed in 2-B and 2-N . Cyclic voltammetry (CV) studies showed that 2-B undergoes a quasi-reversible one-electron reduction, which was confirmed by the synthesis of [K([2.2.2]crypt)][L(Cl)GaSbB[N(Dip)CH]2] ([K([2.2.2]crypt)][ 2-B ]) containing the stibanyl anion [ 2-B ], which was shown to possess significant Sb−B multiple-bonding character.  相似文献   

18.
RhIII‐catalyzed N‐nitroso‐directed C?H addition to ethyl 2‐oxoacetate allows subsequent construction of indazoles, a privileged heterocycle scaffold in synthetic chemistry, through the exploitation of reactivity between the directing group and installed group. The formal [2+2] cycloaddition/fragmentation reaction pathway identified herein, a unique reactivity pattern hitherto elusive for the N‐nitroso group, emphasizes the importance of forward reactivity analysis in the development of useful C?H functionalization‐based synthetic tools. The synthetic utility of the protocol is demonstrated with the synthesis of a tricyclic‐fused ring system. The diversity of covalent linkages available for the nitroso group should enable the extension of the genre of reactivity reported herein to the synthesis of other types of heterocycles.  相似文献   

19.
The first crystallographically characterizable complex of Sc2+, [Sc(NR2)3] (R=SiMe3), has been obtained by LnA3/M reactions (Ln=rare earth metal; A=anionic ligand; M=alkali metal) involving reduction of Sc(NR2)3 with K in the presence of 2.2.2‐cryptand (crypt) and 18‐crown‐6 (18‐c‐6) and with Cs in the presence of crypt. Dark maroon [K(crypt)]+, [K(18‐c‐6)]+, and [Cs(crypt)]+ salts of the [Sc(NR2)3] anion are formed, respectively. The formation of this oxidation state of Sc is also indicated by the eight‐line EPR spectra arising from the I =7/2 45Sc nucleus. The Sc(NR2)3 reduction differs from Ln(NR2)3 reactions (Ln=Y and lanthanides) in that it occurs under N2 without formation of isolable reduced dinitrogen species. [K(18‐c‐6)][Sc(NR2)3] reacts with CO2 to produce an oxalate complex, {K2(18‐c‐6)3}{[(R2N)3Sc]2(μ‐C2O4κ 1O:κ 1O′′)}, and a CO2 radical anion complex, [(R2N)3Sc(μ‐OCO‐κ 1O:κ 1O′)K(18‐c‐6)]n .  相似文献   

20.
The new anionic complexes [K(18-crown-6)][WH5(PMe2Ph)3], [K(1,10-diaza-18-crown-6)][WH5(PMe2Ph)3], [K(2,2,2-crypt)][ReH4(PMePh2)3], and [K(1,10-diaza-18-crown-6)][ReH4(PMePh2)3] were prepared by reaction of KH/crown or KH/crypt with the appropriate neutral polyhydride WH6(PMe2Ph)3 or ReH5(PMePh2)3. The rate of deprotonation of the rhenium hydride in THF is much greater for the reaction involving crypt compared with that of crown. The structure of [ReH4(PMePh2)3]- is distorted pentagonal bipyramidal as determined by an X-ray diffraction study of the crypt salt. No hydridic-protonic M-H...HN bonding is detected between the hydrides of the anionic hydrides and the amino hydrogens of the cations [K(1,10-diaza-18-crown-6)]+ suggesting that stronger M-H...K interactions are present. Acid dissociation constants Ka of polyhydride complexes in THF, approximately corrected for ion pairing, are determined by NMR in order to better understand the periodic trends of metal hydrides. The pKalphaTHF of (WH6(PMe2Ph)3/[WH5(PMe2Ph)3]-) is 42+/-4 according to the equilibrium set up by reacting WH6(PMe2Ph)3 with [K(2,2,2-crypt)][ReH6(PCy3)2]. The pKalphaTHF for ReH5(PMePh2)3 can be estimated as greater than the pKalphaTHF of 38 for HNPh2 and less than the pKalphaTHF of 41 for ReH7(PCy3)2. Reaction of the phosphazene base P4-tBu with ReH7(PCy3)2 gave an equilibrium with [HP4-tBu]+[ReH6(PCy3)2]- whereas reaction with WH6(PMe2Ph)3 gave an equilibrium with [HP4-tBu]+[WH5(PMe2Ph)3]-. From these and a related equilibrium, the pKalphaTHF of [HP4-tBu]+ is found to be 40+/-4. In general, neutral complexes MHx(PR3)n (M=W, Re, Ru, Os, Ir; n=3, 2) studied to date have pKalphaTHF values from 30 to 44 on going from phenyl-substituted to alkyl-substituted phosphine ligands whereas MHx(PR3)n+ (M=Re, Fe, Ru, Os, Co, Rh, Ni, Pd, Pt; n=4, 3), including diphosphine ligands ((PR3)2=PR2-PR2), have values from 12 to 23. From the equilibrium established from the reaction of [HP2-tBu][BPh4] and [K(2,2,2-crypt)][OP(OEt)2NPh], [HP2-tBu]+ was calculated to have a pKalphaTHF of 30+/-4. The equilibrium constant for the similar deprotonation reaction with [K(18-crown-6)][{ReH2(PMePh2)2}2(mu-H)3] confirmed this value.  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