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 twisted lateral tetraalkyloxy ortho‐terphenyl units in dibenzo[18]crown‐6 ethers 1 a – f were readily converted into the flat tetraalkyloxytriphenylene systems 2 a – f by oxidative cyclization with FeCl3 in nitromethane. Reactions of the latter with potassium salts gave complexes KX ?2 , which displayed mesomorphic properties. The aromatization increased both the clearing and melting points; the mesophase stabilities, however, were mainly influenced by the respective anions upon complexation with various potassium salts. In contrast, the alkyl chain lengths played only a secondary role. Among the potassium complexes of triphenylene‐substituted crown ethers KX ?2 , only those with the soft anions I? and SCN? displayed mesophases with expanded phase temperature ranges of 93 °C and 132 °C (for KX ?2 e ), respectively, as compared to the corresponding o‐terphenyl‐substituted crown ether complexes KI ?1 e (ΔT=51 °C) and KSCN ?1 e (plastic crystal phase). Anions such as Br?, Cl?, and F? decreased the mesophase stability, and PF6? led to complete loss of the mesomorphic properties of KPF6 ?2 although not for KPF6 ?1 . For crown ether complexes KX ?2 (X=F, Cl, Br, I, BF4, and SCN), columnar rectangular mesophases of different symmetries (c2 mm, p2 mg, and p2 gg) were detected. In contrast to findings for the twisted o‐terphenyl crown ether complexes KX ?1 , the complexation of the flat triphenylene crown ethers 2 with KX resulted in the formation of organogels. Characterization of the organogel of KI ?2 e in CH2Cl2 revealed a network of fibers.  相似文献   

2.
An M2L4 quadruple helicate, formed by wrapping four molecules of 1,4‐bis(3‐pyridyloxy)benzene ( L1 ) about two palladium(II) centers, is shown to bind anions within its internal cavity. 1H NMR exchange experiments provide a quantitative measure of anion selectivity and reveal a preference for ClO4? over the other tetrahedral anions BF4? and ReO4? and the octahedral anion PF6?. X‐ray crystal structures of [Pd2( L1 )4]4+ helicates containing ClO4, BF4? and I? reveal that the cavity size can dynamically change in response to the size of the guest.  相似文献   

3.
A series of half‐sandwich rhodium‐based metallamacrocycles with tetra‐ and hexanuclearities have been synthesized. They are assembled by linking the deprotonated 2,4‐diacetyl‐5‐hydroxy‐5‐methyl‐3‐(3‐pyridinyl)cyclohexanone (HL) ligand in the presence of counteranions. When the counteranion was the tetrahedral BF4? ion, tetranuclear metallamacrocycle [(Cp*RhL)4][BF4]4 ( 1 d ) was formed. However, the larger OTf?, PF6?, and SbF6? counterions favored the formation of hexanuclear metallamacrocycles [(Cp*RhL)6?2OTf][OTf]4 ( 1 a ), [(Cp*RhL)6?2PF6][PF6]4 ( 1 b ), and [(Cp*RhL)6?2SbF6][SbF6]4 ? 6CH3CN ( 1 c ) when the reactions were performed under the same conditions. Single‐crystal X‐ray analysis indicated that, in the solid state, two counteranions were encapsulated in each belt‐like host molecule of hexanuclear metallamacrocycles 1 a , b , and c . Based on the results of 1H NMR analysis in methanol, the nuclearities of 1 a – c and the two encapsulated anions in each molecular cavity were maintained in solution. In addition, tetranuclear metallamacrocycle 1 d was converted into the hexanuclear metallamacrocycles 1 a′ , b , and c after addition of the appropriate anion as its [NBu4]+ salt. The compound 1 a′ was characterized by single‐crystal X‐ray diffraction to have the formula [(Cp*RhL)6?2OTf][BF4]4 ? 2M eOH ? 2H2O. From the interconversion of the hexanuclear metallamacrocycles, we have concluded that the hexanuclear belt‐like host in 1 a – c has an clear selectivity for larger anions, in the sequence SbF6?≈PF6?>OTf?>BF4?>Cl?.  相似文献   

4.
The 1H NMR chemical shifts of the C(α)? H protons of arylmethyl triphenylphosphonium ions in CD2Cl2 solution strongly depend on the counteranions X?. The values for the benzhydryl derivatives Ph2CH? PPh3+ X?, for example, range from δH=8.25 (X?=Cl?) over 6.23 (X?=BF4?) to 5.72 ppm (X?=BPh4?). Similar, albeit weaker, counterion‐induced shifts are observed for the ortho‐protons of all aryl groups. Concentration‐dependent NMR studies show that the large shifts result from the deshielding of the protons by the anions, which decreases in the order Cl? > Br? ? BF4? > SbF6?. For the less bulky derivatives PhCH2? PPh3+ X?, we also find C? H???Ph interactions between C(α)? H and a phenyl group of the BPh4? anion, which result in upfield NMR chemical shifts of the C(α)? H protons. These interactions could also be observed in crystals of (p‐CF3‐C6H4)CH2? PPh3+ BPh4?. However, the dominant effects causing the counterion‐induced shifts in the NMR spectra are the C? H???X? hydrogen bonds between the phosphonium ion and anions, in particular Cl? or Br?. This observation contradicts earlier interpretations which assigned these shifts predominantly to the ring current of the BPh4? anions. The concentration dependence of the 1H NMR chemical shifts allowed us to determine the dissociation constants of the phosphonium salts in CD2Cl2 solution. The cation–anion interactions increase with the acidity of the C(α)? H protons and the basicity of the anion. The existence of C? H???X? hydrogen bonds between the cations and anions is confirmed by quantum chemical calculations of the ion pair structures, as well as by X‐ray analyses of the crystals. The IR spectra of the Cl? and Br? salts in CD2Cl2 solution show strong red‐shifts of the C? H stretch bands. The C? H stretch bands of the tetrafluoroborate salt PhCH2? PPh3+ BF4? in CD2Cl2, however, show a blue‐shift compared to the corresponding BPh4? salt.  相似文献   

5.
Preparation and Characterization of Cationic η2-1-Butene and Acetonitrile Complexes The reaction of the species η5-C5H5M(CO)n-σ-C4H7 (M = Fe, Mo, W; n = 2, 3) with (C6H5)3CBF4 yielded – instead of the expected cationic butadiene complexes of the type [η5-CpM(CO)n?14-C4H6][BF4], which would have been formed in case of hydride cleavage – compounds of the type [η5-CpM(CO)n η2-C4H8][BF4], which were formed by protonation of the σ-C4H7 ligands. The reaction proceeded quantitatively. The BF4? anion can be substituted by other anions, such as ClO4?, B(C6H5)4?, PF4?, and [Cr(SCN)4(NH3)2]? in the complexes obtained. The mechanism of the reaction leading to the η2-bonded 1-butene complexes was determined by isotope experiments. In trying to recrystallize the butene complexes from acetonitrile the cationic complexes [η5-C5H5 Fe(CO)2CH3CN]BF4 and [η5-C5H5 M(CO)3CH3CN]BF4 were observed; the X-ray structure analysis of the former is reported.  相似文献   

6.
As is well‐known, the C2?H proton of 1‐ethyl‐3‐methylimidazolium tetrafluoroborate ([Emim]BF4) and 1‐butyl‐3‐methylimidazolium tetrafluoroborate ([Bmim]BF4) has a strong ability to form hydrogen bonds. The purpose of this work is to evaluate the effect of the interactions of the C4?H and C5?H protons on the microstructure of [Emim]BF4 and [Bmim]BF4 with water by using 1H NMR spectroscopy. The differences between the relative 1H NMR chemical shifts of C2?H, C4?H, and C5?H and between the interaction‐energy parameters obtained from these chemical shifts are minor, thus suggesting that the interactions of C4?H and C5?H may have a considerable effect on the microstructure. To confirm this, the viscosities of the systems are estimated by using the interaction‐energy parameters obtained from the 1H NMR chemical shifts of the three studied aromatic protons and water, showing that the interactions of C4?H and C5?H also play an important role in the microstructure.  相似文献   

7.
The mononuclear η5-cyclopentadienyl complexes [(η5-C5H5)Ru(PPh3)2Cl], [(η5-C5H5)Os(PPh3)2Br] and pentamethylcyclopentadienyl complex [(η5-C5Me5)Ru(PPh3)2Cl] react in the presence of 1 eq. of the tetradentate N,N′-chelating ligand 3,5-bis(2-pyridyl)pyrazole (bpp-H) and 1 eq. of NH4PF6 in methanol to afford the mononuclear complexes [(η5-C5H5)Ru(PPh3)(bpp-H)]PF6 ([1]PF6), [(η5-C5H5)Os(PPh3)(bpp-H)]PF6 ([2]PF6) and [(η5-C5Me5)Ru(PPh3)(bpp-H)]PF6 ([3]PF6), respectively. The dinuclear η5-pentamethylcyclopentadienyl complexes [(η5-C5Me5)Rh(μ-Cl)Cl]2 and [(η5-C5Me5)Ir(μ-Cl)Cl]2 as well as the dinuclear η6-arene ruthenium complexes [(η6-C6H6)Ru(μ-Cl)Cl]2 and [(η6-p-iPrC6H4Me)Ru(μ-Cl)Cl]2 react with 2 eq. of bpp-H in the presence of NH4PF6 or NH4BF4 to afford the corresponding mononuclear complexes [(η5-C5Me5)Rh(bpp-H)Cl]PF6 ([4]PF6), [(η5-C5Me5)Ir(bpp-H)Cl]PF6 ([5]PF6), [(η6-C6H6)Ru(bpp-H)Cl]BF4 ([6]BF4) and [(η6-p-iPrC6H4Me)Ru(bpp-H)Cl]BF4 ([7]BF4). However, in the presence of 1 eq. of bpp-H and NH4BF4 the reaction with the same η6-arene ruthenium complexes affords the dinuclear salts [(η6-C6H6)2Ru2(bpp)Cl2]BF4 ([8]BF4) and [(η6-p-iPrC6H4Me)2Ru2(bpp)Cl2]BF4 ([9]BF4), respectively. These compounds have been characterized by IR, NMR and mass spectrometry, as well as by elemental analysis. The molecular structures of [1]PF6, [5]PF6 and [8]BF4 have been established by single crystal X-ray diffraction studies and some representative complexes have been studied by UV–vis spectroscopy.  相似文献   

8.
The formation of complexes between hexafluorophosphate (PF6) and tetraisobutyloctahydroxypyridine[4]arene has been thoroughly studied in the gas phase (ESI‐QTOF‐MS, IM‐MS, DFT calculations), in the solid state (X‐ray crystallography), and in chloroform solution (1H, 19F, and DOSY NMR spectroscopy). In all states of matter, simultaneous endo complexation of solvent molecules and exo complexation of a PF6 anion within a pyridine[4]arene dimer was observed. While similar ternary complexes are often observed in the solid state, this is a unique example of such behavior in the gas phase.  相似文献   

9.
Syntheses and Structures of Bis(4,4′‐t‐butyl‐2,2′‐bipyridine) Ruthenium(II) Complexes with functional Derivatives of Tetramethyl‐bibenzimidazole [(tbbpy)2RuCl2] reacts with dinitro‐tetramethylbibenzimidazole ( A ) in DMF to form the complex [(tbbpy)2Ru( A )](PF6)2 ( 1a ) (tbbpy: bis(4,4′‐t‐butyl)‐2,2′bipyridine). Exchange of the two PF6? anions by a mixture of tetrafluor‐terephthalat/tetrafluor‐terephthalic acid results in the formation of 1b in which an extended hydrogen‐bonded network is formed. According to the 1H NMR spectra and X‐ray analyses of both 1a and 1b , the two nitro groups of the bibenzimidazole ligand are situated at the periphery of the complex in cis position to each other. Reduction of the nitro groups in 1a with SnCl2/HCl results in the corresponding diamino complex 2 which is a useful starting product for further functionalization reactions. Substitution of the two amino groups in 2 by bromide or iodide via Sandmeyer reaction results in the crystalline complexes [(tbbpy)2Ru( C )](PF6)2 and [(tbbpy)2Ru( D )](PF6)2 ( C : dibromo‐tetrabibenzimidazole, D : diiodo‐tetrabibenzimidazole). Furthermore, 2 readily reacts with 4‐t‐butyl‐salicylaldehyde or pyridine‐2‐carbaldehyde under formation of the corresponding Schiff base RuII complexes 5 and 6 . 1H NMR spectra show that the substituents (NH2, Br, I, azomethines) in 2 ‐ 6 are also situated in peripheral positions, cis to each other. The solid state structure of both 2 , and 3 , determined by X‐ray analyses confirm this structure. In addition, the X‐ray diffraction analyses of single crystals of the complexes [(tri‐t‐butyl‐terpy)(Cl)Ru( A )] ( 7 ) and [( A )PtCl2] ( 8 ) display also that the nitro groups in these complexes are in a cis‐arrangement.  相似文献   

10.
Several bis(triazolium)‐based receptors have been synthesized as chemosensors for anion recognition. The central naphthalene core features two aryltriazolium side‐arms. NMR experiments revealed differences between the binding modes of the two triazolium rings: one triazolium ring acts as a hydrogen‐bond donor, the other as an anion–π receptor. Receptors 92+?2BF4 ? (C6H5), 112+?2BF4 ? (4‐NO2?C6H4), and 132+?2BF4? (ferrocenyl) bind HP2O73? anions in a mixed‐binding mode that features a combination of hydrogen‐bonding and anion–π interactions and results in strong binding. On the other hand, receptor 102+?2 BF4 ? (4‐CH3O?C6H4) only displays combined Csp2?H/anion–π interactions between the two arms of the receptors and the bound anion rather than triazolium (CH)+???anion hydrogen bonding. All receptors undergo a downfield shift of the triazolium protons, as well as the inner naphthalene protons, in the presence of H2PO4? anions. That suggests that only hydrogen‐bonding interactions exist between the binding site and the bound anion, and involve a combination of cationic (triazolium) and neutral (naphthalene) C?H donor interactions. Theoretical calculations relate the electronic structure of the substituent on the aromatic group with the interaction energies and provide a minimum‐energy conformation for all the complexes that explains their measured properties.  相似文献   

11.
The new bis-4-(hexyloxy)benzamide and bis-4-(dodecyloxy)benzamide derivatives of macrocyclic ligands [18]aneN2O4 (L1 and L2) and [18]aneN2S4 (L3 and L4) have been prepared and characterized by1H and13C NMR spectroscopy, differential scanning calorimetry and optical microscopy. The nonmesogenic ligands L3 and L4 react with [Pd(CH3CN)4][BF4]2 to give the cationic complexes [Pd(L3)][BF4]2 and [Pd(L4)][BF4]2. [Pd(L4)][BF4]2 shows mesomorphic behaviour affording a smectic mesophase.  相似文献   

12.
New biligand complexes of iron(III) are synthesized based on 3,4,5-tri(tetradecyloxy)benzoyloxy-4-salicylidene-N′-ethyl-N-ethylenediamine azomethine with the outer sphere NO3?, PF6?, Cl?, BF4?, ClO4?, and CNS anions. All the target compounds are characterized by gel exclusion chromatography, elemental analysis, and electron, IR, and NMR spectroscopy. The presence of complex-forming ions is confirmed by FT-IR spectra in the far region. The formation of biligand polychelate complexes with an octahedral packing of the iron ion is observed. Phase transitions in the resulting coordination compounds are studied by differential scanning calorimetry and optical polarizing thermomicroscopy. The presence of several polymorphic crystalline modifications, as well as mesophases, is established. Mesomorphic properties are found for complexes with chloride and tetrafluoroborate anions.  相似文献   

13.
Considering the ionic nature of ionic liquids (ILs), ionic association is expected to be essential in solutions of ILs and to have an important influence on their applications. Although numerous studies have been reported for the ionic association behavior of ILs in solution, quantitative results are quite scarce. Herein, the conductivities of the ILs [Cnmim]Br (n=4, 6, 8, 10, 12), [C4mim][BF4], and [C4mim][PF6] in various molecular solvents (water, methanol, 1‐propanol, 1‐pentanol, acetonitrile, and acetone) are determined at 298.15 K as a function of IL concentration. The conductance data are analyzed by the Lee–Wheaton conductivity equation in terms of the ionic association constant (KA) and the limiting molar conductance (Λm0). Combined with the values for the Br? anion reported in the literature, the limiting molar conductivities and the transference numbers of the cations and [BF4]? and [PF6]? anions are calculated in the molecular solvents. It is shown that the alkyl chain length of the cations and type of anion affect the ionic association constants and limiting molar conductivities of the ILs. For a given anion (Br?), the Λm0 values decrease with increasing alkyl chain length of the cations in all the molecular solvents, whereas the KA values of the ILs decrease in organic solvents but increase in water as the alkyl chain length of the cations increases. For the [C4mim]+ cation, the limiting molar conductivities of the ILs decrease in the order Br?>[BF4]?>[PF6]?, and their ionic association constants follow the order [BF4]?>[PF6]?>Br? in water, acetone, and acetonitrile. Furthermore, and similar to the classical electrolytes, a linear relationship is observed between ln KA of the ILs and the reciprocal of the dielectric constants of the molecular solvents. The ILs are solvated to a different extent by the molecular solvents, and ionic association is affected significantly by ionic solvation. This information is expected to be useful for the modulation of the IL conductance by the alkyl chain length of the cations, type of anion, and physical properties of the molecular solvents.  相似文献   

14.
The synthesis, structure, and properties of bischloro, μ‐oxo, and a family of μ‐hydroxo complexes (with BF4?, SbF6?, and PF6? counteranions) of diethylpyrrole‐bridged diiron(III) bisporphyrins are reported. Spectroscopic characterization has revealed that the iron centers of the bischloro and μ‐oxo complexes are in the high‐spin state (S=5/2). However, the two iron centers in the diiron(III) μ‐hydroxo complexes are equivalent with high spin (S=5/2) in the solid state and an intermediate‐spin state (S=3/2) in solution. The molecules have been compared with previously known diiron(III) μ‐hydroxo complexes of ethane‐bridged bisporphyrin, in which two different spin states of iron were stabilized under the influence of counteranions. The dimanganese(III) analogues were also synthesized and spectroscopically characterized. A comparison of the X‐ray structural parameters between diethylpyrrole and ethane‐bridged μ‐hydroxo bisporphyrins suggest an increased separation, and hence, less interactions between the two heme units of the former. As a result, unlike the ethane‐bridged μ‐hydroxo complex, both iron centers become equivalent in the diethylpyrrole‐bridged complex and their spin state remains unresponsive to the change in counteranion. The iron(III) centers of the diethylpyrrole‐bridged diiron(III) μ‐oxo bisporphyrin undergo very strong antiferromagnetic interactions (J=?137.7 cm?1), although the coupling constant is reduced to only a weak value in the μ‐hydroxo complexes (J=?42.2, ?44.1, and ?42.4 cm?1 for the BF4, SbF6, and PF6 complexes, respectively).  相似文献   

15.
The reaction of [Pt(CH2COMe)(Ph)(cod)] (cod=1,5‐cyclooctadiene) with (ArCH2NH2CH2‐C6H4COOH)+(PF6)? (Ar=4‐tBuC6H4 or 9‐anthryl) in the presence of cyclic oligoethers such as dibenzo[24]crown‐8 (DB24C8) and dicyclohexano[24]crown‐8 (DC24C8) produces {(ce)[ArCH2NH2CH2C6H4COOPt(Ph)(cod)]}+(PF6)? (ce=DB24C8 or DC24C8, Ar=4‐tBuC6H4 or 9‐anthryl) with interlocked structures. FABMS and NMR spectra of a solution of these compounds indicate that the Pt complexes with a secondary ammonium group and DB24C8 (or DC24C8) make up the axis and cyclic components, respectively. Temperature‐dependent 1H NMR spectra of a solution of {(DB24C8)[4‐tBuC6H4CH2NH2CH2‐C6H4COOPt(Ph)(cod)]}+(PF6)? ({(DB24C8)[ 4 ‐H]}+(PF6)?) show equilibration with free DB24C8 and the axis component. The addition of DB24C8 to a solution of {(DC24C8)[ 4 ‐H]}+(PF6)? causes partial exchange of the macrocyclic component of the interlocked molecules, giving a mixture of {(DC24C8)[ 4 ‐H]}+(PF6)?, {(DB24C8)[ 4 ‐H]}+(PF6)?, and free macrocyclic compounds. The reaction of 3,5‐Me2C6H3COCl with {(DB24C8)[ 4 ‐H]}+(PF6)? affords the organic rotaxane {(DB24C8)(4‐tBuC6H4CH2NH2CH2‐C6H4COOCOC6H3Me2‐3,5)}+(PF6)? through C? O bond formation between the aroyl group and the carboxylate ligand of the axis component. The addition of 2,2′‐bipyridine (bpy) to a solution of {(DB24C8)[ 4 ‐H]}+(PF6)? induces the degradation of the interlocked structure to form a complex with trigonal bipyramidal coordination, [Pt(Ph)(bpy)(cod)]+(PF6)?, whereas the reaction of bpy with [Pt(OCOC6H4Me‐4)(Ph)(cod)] produces the square‐planar complex [Pt(OCOC6H4Me‐4)(Ph)(bpy)].  相似文献   

16.
2‐Thienylpyridyl iridium(III) complexes containing an o‐, m‐, or p‐carboranylvinyl‐2,2′‐bipyridine ligand and various counteranions (denoted o ‐ PF6 , m ‐ BF4 , m ‐ PF6 , m ‐ SbF6 , m ‐ ClO4 , m ‐ OTf , m ‐ NO3 , m ‐ BPh4 , m ‐ F , m ‐ Cl , and p ‐ PF6 ) were synthesized by using C‐formyl carboranes as starting materials. The solid‐state structures of o ‐ PF6 , m ‐ PF6 , m ‐ ClO4 , and m ‐ BF4 showed that the cations form twisted cavities in which the anions are fixed by multiple hydrogen bonds. Anion–hydrogen interactions were investigated for nine m‐carborane‐based complexes with different counteranions. All carborane‐based iridium(III) complexes show similar phosphorescence yields in solution but significantly different emission in the solid state. Anion‐exchange titration and theoretical calculations revealed the relationships between structures and optical properties. The size of the anion and C?H ??? X anion–hydrogen bonds strongly influence the phosphorescence quantum yield in the solid state. In particular, the Ccar?H ??? X hydrogen bonds between the carboranyl unit and the anion play an important role in solid‐state phosphorescence. Complex p ‐ PF6 was successfully applied in phosphorescence‐lifetime bioimaging owing to its low toxicity and near‐infrared emission.  相似文献   

17.
The pseudo‐square‐planar complexes [Rh(cod)(Hbbtm)]BF4 ( 3 ), [Rh(bbte)(cod)]BF4 ( 4 ), [Rh(CO)2(Hbbtm)]BF4 ( 5 ), [Rh(bbte)(CO)2]BF4 ( 6 ), [Rh(bbtm)(cod)] ( 7 ) and [Rh(bbtm)(CO)2] ( 8 ) (Hbbtm=bis(benzothiazol‐2‐yl)methane=2,2′‐methylenebis[benzothiazole], bbte=bis(benzothiazol‐2‐yl)ethane=2,2′‐(ethane‐1,2‐diyl)bis[benzothiazole], and cod=cycloocta‐1,5‐diene) were synthesized and characterized. Diastereotopic protons were observed for the protons at the bridge in the 1H‐NMR of 3 and 5 . Twisting of the ethane‐1,2‐diyl bridge in 4 and 6 effects chemical equivalence of the CH2 groups in solution. Unusually large downfield shifts occur on coordination of the deprotonated ligand Hbbtm as the negative charge is delocalized in 7 and 8 . The NMR signals of the cod ligand in 4 could be differentiated. The X‐ray crystal structures of 3, 4 , and 6 are reported.  相似文献   

18.
A series of dicarbene‐bridged metallacycles [Ag2( 1 )2](PF6)2, [Ag2( 2 )2](BF4)2, [Ag2( 3 )2](PF6)2, [Ag2( 7 )2](BF4)2, [Ag2( 8 )2](BF4)2 and [Ag2( 11 )2](PF6)2 were obtained in high yields via the reactions of 1,2,4‐triazole‐, 1,2,3‐triazole‐ and imidazo[1,5‐a]pyridine‐based ligands with Ag2O in CH3CN. The C=C double bonds in all of the newly synthesized metallacycles went through [2 + 2] photodimerization under UV irradiation condition (λ = 365 nm, T = 298 K) yielding the dinuclear rctt‐cyclobutane‐silver(I) complexes [Ag2( 4 )](PF6)2, [Ag2( 5 )](BF4)2, [Ag2( 6 )](PF6)2, [Ag2( 9 )](BF4)2, [Ag2( 10 )](BF4)2 and [Ag2( 12 )](PF6)2, respectively with quantitative yields. Treatment of the these cyclobutane‐bridged silver(I) complexes with NH4Cl resulted in the exclusive formation of cyclobutane derivatives after removal of the silver(I) metal ions.  相似文献   

19.
A series of fluorescent imidazolium‐based salts containing the cation [AnCH2MeIm]+ (in which An=anthracene and Im=the imidazolium cation) with Cl?, BF4?, PF6?, SO3CF3?, [N(CN)2]?, [N(SO2CF3)2]?, or PhBF3? anions have been prepared and characterized. X‐ray diffraction analysis of four of the salts reveals a number of C? H???X‐type (X=O, N, F) hydrogen bonds between the hydrogen atoms from the imidazolium ring and in some cases from the anthracene ring with the electronegative atoms of the anions. Additionally, C? H???π interactions can be found in all the salts analyzed by X‐ray diffraction, whereas π–π stacking is observed only in the salt containing the phenyltrifluoroborate anion. Fluorescence emission analysis in acetonitrile shows that the fluorescence of these salts varies significantly according to the nature of the anion, and correlates to the extent of ion pairing present in solution. Photodimerization of these salts was observed, and in one case a dimer has been isolated and characterized by X‐ray crystallography.  相似文献   

20.
Four new picolyl hydrazones were prepared via Schiff-base condensation of picolonic acid hydrazide with α-formyl-(L1), α-acetyl-(L2), α-benzoyl-(L3) pyridine and α-formyl-(L4) thiophene. Copper(II) complexes of these hydrazones and a series of copper(II) complexes containing (L2) and various anions (Cl, Br, NO3, SCN, SO4, ClO4, AcO, PF6 and BF4) have been synthesized. Elemental, thermal analysis, molar conductivity, magnetic moment measurements and spectral (i.r., electronic and e.s.r.) studies have been used to characterize the prepared compounds. The overall structure and reactivity of the reported copper(II) chelates critically depend on the ligand structure and the nature of counter anion incorporated in the complex molecule. Octahedral [complex (7)], square-pyramidal [complex (8)] and square-planar monomeric species [complexes (1–6), (9) and (10)] and a dimeric structure with oxygen bridge in square-planar geometry [complexes (11) and (12)] were suggested. The reported copper(II) complexes exhibit promising oxidase catalytic activity towards the aerobic oxidation of vitamin C. A linear correlation exists between the oxidase catalytic activity and the Lewis-acidity of the central copper(II) ion created by the donating properties of the parent ligand, as well the irregularity of the coordination environment. The probable mechanistic implications of the catalytic oxidation reactions are discussed.  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