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 640 毫秒
1.
The self‐assembled trimetallic species [L2Cu3]6+ contains a cavity that acts as a host to many different anions. By using X‐ray crystallography, ESI‐MS, and UV/Vis spectroscopy we show that these anions are encapsulated both in the solid state and aqueous systems. Upon encapsulation, the anions Br, I, CO32−, SiF62−, IO63−, VO43−, WO42−, CrO42−, SO42−, AsO43−, and PO43− are all precipitated from aqueous solution and can be removed by filtration. Furthermore, the cavity can be tuned to be selective to either phosphate or sulfate anions by variation of the pH. Phosphate anions can be removed from water, even in the presence of other common anions, reducing the concentration from 1000 to <0.1 ppm and recovering approximately 99 % of the phosphate anions.  相似文献   

2.
《化学:亚洲杂志》2017,12(15):1909-1914
A dodecavanadate, [V12O32]4−, is an inorganic bowl‐type host with a cavity entrance with a diameter of 4.4 Å in the optimized structure. Linear, bent, and trigonal planar anions are tested as guest anions and the formation of host–guest complexes, [V12O32(X)]5− (X=CN, OCN, NO2, NO3, HCO2, and CH3CO2), were confirmed by X‐ray crystallographic analyses and a 51V NMR spectroscopy study. The degree of distortion of the bowl from a regular to an oval shape depends on the type of guest anion. In 51V NMR spectroscopy, all chemical shifts of the host–guest complexes are clearly shifted after guest incorporation. The incorporation reaction rates for OCN, NO2, HCO2, and CH3CO2 are much larger than those of NO3 and halides. The incorporated nonspherical molecular anions in the dodecavanadate host are easily dissociated or exchanged for other anions, whereas spherical halides in the host are preserved without dissociation, even in the presence of the tested anions.  相似文献   

3.
《中国化学》2017,35(7):1165-1169
We synthesized a new cyanide (CN ) chemosensor CX based on a nucleophilic addition reaction prompted by cyanide ion, which could be used for highly selective and sensitive fluorescence turn‐on detection of cyanide in aqueous media. The CX showed selective fluorescence recognition for CN , the miscellaneous competitive anions (F, Cl, Br, I, AcO , H2PO4, HSO4, ClO4, S2 , PO43−, CO32− and SCN ) did not lead to any significant interference. The detection limit of the sensor towards CN is 1.15 × 10−7 mol•L−1. The sensor has been successfully applied to estimate the cyanide ion in seeds of cherries. Test strips based on CX were fabricated, which could be used as a convenient and efficient CN test kit to detect CN in aqueous solution for “in‐the‐field” measurement.  相似文献   

4.
A systematical evaluation of association constants between halide, phosphate, and carboxylate anions with N‐methylformamide ( 1 ) and the related bidentate receptors 2 – 6 (derived from, e.g., phthalic acid or ethylenediamine) in CDCl3 as solvent yielded increments of complexation free‐energy ΔΔG for each single H‐bond, which varied like, e.g., 5.1 kJ/mol (for Cl), 4.0 kJ/mol (for Br), 4.0 kJ/mol (for I) (with values taken from Tables 1 and 2), in line with expected H‐bond strength. The observed complexation induced NH‐NMR shift (CIS) values also showed a regular change, in the case of 1 , e.g., from 5.0 to 2.8 to 2.1 ppm (Table 1), with about half of these values with the bidentate ligands (Tables 2 and 3). Tridentate hosts led to a substantial binding increase, if strain‐free convergence of all NH donor functions towards the anion was possible. The tris[urea] ligand 10 yielded, even in the polar solvent DMSO, with Cl a ΔG of −21.5 kJ/mol and with Br of −10⋅5 kJ/mol, whereas with I, no association was detectable. The results demonstrated that small, inexpensive, and conformationally mobile host compounds can exhibit high affinities as well as descrimination with anions, as much as more preorganized receptors do which require multistep synthesis. The corresponding adamantyl derivative 13 allowed measurements also in CDCl3, with K=4.3⋅104 M −1 for chloride (Table 7). Complexes with nucleotide anions were again particularly strong with the tridentate urea‐based ligands, the latter being optimal ligands for chloride complexation. For the association of 10 with AMP2− and GMP2−in (D6)DMSO, the association constants were 3⋅104 M −1 (Table 8) and almost the same as with Cl. In the case of the urea derivatives 17 , 18 , and 21 , containing only one phenyl or pyrenyl substituent, however, the ΔG values decreased in the order A>C>T>G (e.g. −13.6, −11.6, −7.6, −10.5 kJ/mol in the case of 17 , resp.; Table 8). In H2O, the pyrenyl‐substituted urea derivatives allow measurements with fluorescence, and, unexpectedly, show only smaller nucleobase discrimination, with constants around 3⋅103 M −1.  相似文献   

5.
《化学:亚洲杂志》2017,12(8):910-919
Reduction of aluminum(III), gallium(III), and indium(III) phthalocyanine chlorides by sodium fluorenone ketyl in the presence of tetrabutylammonium cations yielded crystalline salts of the type (Bu4N+)2[MIII(HFl−O)(Pc.3−)].−(Br) ⋅ 1.5 C6H4Cl2 [M=Al ( 1 ), Ga ( 2 ); HFl−O=fluoren‐9‐olato anion; Pc=phthalocyanine] and (Bu4N+) [InIIIBr(Pc.3−)].− ⋅ 0.875 C6H4Cl2 ⋅ 0.125 C6H14 ( 3 ). The salts were found to contain Pc.3− radical anions with negatively charged phthalocyanine macrocycles, as evidenced by the presence of intense bands of Pc.3− in the near‐IR region and a noticeable blueshift in both the Q and Soret bands of phthalocyanine. The metal(III) atoms coordinate HFl−O anions in 1 and 2 with short Al−O and Ga−O bond lengths of 1.749(2) and 1.836(6) Å, respectively. The C−O bonds [1.402(3) and 1.391(11) Å in 1 and 2 , respectively] in the HFl−O anions are longer than the same bond in the fluorenone ketyl (1.27–1.31 Å). Salts 1 – 3 show effective magnetic moments of 1.72, 1.66, and 1.79 μB at 300 K, respectively, owing to the presence of unpaired S= 1/2 spins on Pc.3−. These spins are coupled antiferromagnetically with Weiss temperatures of −22, −14, and −30 K for 1 – 3 , respectively. Coupling can occur in the corrugated two‐dimensional phthalocyanine layers of 1 and 2 with an exchange interaction of J /k B=−0.9 and −1.1 K, respectively, and in the π‐stacking {[InIIIBr(Pc.3−)].−}2 dimers of 3 with an exchange interaction of J /k B=−10.8 K. The salts show intense electron paramagnetic resonance (EPR) signals attributed to Pc.3−. It was found that increasing the size of the central metal atom strongly broadened these EPR signals.  相似文献   

6.
The data on temperature, solvent, and high hydrostatic pressure influence on the rate of the ene reactions of 4‐phenyl‐1,2,4‐triazoline‐3,5‐dione ( 1 ) with 2‐carene ( 2 ), and β‐pinene ( 4 ) have been obtained. Ene reactions 1 + 2 and 1 + 4 have high heat effects: ∆Hrn ( 1 + 2 ) −158.4, ∆Hrn( 1 + 4 ) −159.2 kJ mol−1, 25°C, 1,2‐dichloroethane. The comparison of the activation volume (∆V( 1 + 2 ) −29.9 cm3 mol−1, toluene; ∆V( 1 + 4 ) −36.0 cm3 mol−1, ethyl acetate) and reaction volume values (∆Vr‐n( 1 + 2 ) −24.0 cm3 mol−1, toluene; ∆Vr‐n( 1 + 4 ) −30.4 cm3 mol−1, ethyl acetate) reveals more compact cyclic transition states in comparison with the acyclic reaction products 3 and 5 . In the series of nine solvents, the reaction rate of 1+2 increases 260‐fold and 1+4 increases 200‐fold, respectively, but not due to the solvent polarity.  相似文献   

7.
In the salt 1‐methylpiperazine‐1,4‐diium bis(dihydrogen phosphate), C5H13N22+·2H2PO4, (I), and the solvated salt 2‐(pyridin‐2‐yl)pyridinium dihydrogen phosphate–orthophosphoric acid (1/1), C10H9N2+·H2PO4·H3PO4, (II), the formation of O—H...O and N—H...O hydrogen bonds between the dihydrogen phosphate (H2PO4) anions and the cations constructs a three‐ and two‐dimensional anionic–cationic network, respectively. In (I), the self‐assembly of H2PO4 anions forms a two‐dimensional pseudo‐honeycomb‐like supramolecular architecture along the (010) plane. 1‐Methylpiperazine‐1,4‐diium cations are trapped between the (010) anionic layers through three N—H...O hydrogen bonds. In solvated salt (II), the self‐assembly of H2PO4 anions forms a two‐dimensional supramolecular architecture with open channels projecting along the [001] direction. The 2‐(pyridin‐2‐yl)pyridinium cations are trapped between the open channels by N—H...O and C—H...O hydrogen bonds. From a study of previously reported structures, dihydrogen phosphate anions show a supramolecular flexibility depending on the nature of the cations. The dihydrogen phosphate anion may be suitable for the design of the host lattice for host–guest supramolecular systems.  相似文献   

8.
Crystallization of N,N′‐dimethylpyrazinediium bis(tetrafluoroborate), C6H10N22+·2BF4, (I), and N,N′‐diethylpyrazinediium bis(tetrafluoroborate), C8H14N22+·2BF4, (II), from dried acetonitrile under argon protection has permitted their single‐crystal studies. In both crystal structures, the pyrazinediium dications are located about an inversion center (located at the ring center) and each pyrazinediium aromatic ring is π‐bonded to two centrosymmetrically related BF4 anions. Strong anion–π interactions, as well as weak C—H...F hydrogen bonds, between BF4 and pyrazinediium ions are present in both salts.  相似文献   

9.
《化学:亚洲杂志》2017,12(15):1952-1964
Because of the devastating impact of arsenic on terrestrial and aquatic organisms, the recovery, removal, disposal, and management of arsenic‐contaminated water is a considerable challenge and has become an urgent necessity in the field of water treatment. This study reports the controlled fabrication of a low‐cost adsorbent based on microscopic C‐,N‐doped NiO hollow spheres with geode shells composed of poly‐CN nanospherical nodules (100 nm) that were intrinsically stacked and wrapped around the hollow spheres to form a shell with a thickness of 500–700 nm. This C‐,N‐doped NiO hollow‐sphere adsorbent (termed CNN) with multiple diffusion routes through open pores and caves with connected open macro/meso windows over the entire surface and well‐dispersed hollow‐sphere particles that create vesicle traps for the capture, extraction, and separation of arsenate (AsO43−) species from aqueous solution. The CNN structures are considered to be a potentially attractive adsorbent for AsO43− species due to 1) superior removal and trapping capacity from water samples and 2) selective trapping of AsO43− from real water samples that mainly contained chloride and nitrate anions and Fe2+, and Mn2+, Ca2+, and Mg2+ cations. The structural stability of the hierarchal geodes was evident after 20 cycles without any significant decrease in the recovery efficiency of AsO43− species. To achieve low‐cost adsorbents and toxic‐waste management, this superior CNN AsO43− dead‐end trapping and recovery system evidently enabled the continuous control of AsO43− disposal in water‐scarce environments, presents a low‐cost and eco‐friendly adsorbent for AsO43− species, and selectively produced water‐free arsenate species. These CNN geode traps show potential as excellent adsorbent candidates in environment remediation tools and human healthcare.  相似文献   

10.
The title compound, poly[[μ4‐5‐carboxy‐4‐carboxylato‐2‐(pyridin‐4‐yl)‐1H‐imidazol‐1‐ido]disilver(I)], [Ag2(C10H5N3O4)]n, was synthesized by reacting silver nitrate with 2‐(pyridin‐4‐yl)‐1H‐imidazole‐4,5‐dicarboxylic acid (H3PyIDC) under hydrothermal conditions. The asymmetric unit contains two crystallographically independent AgI cations and one unique HPyIDC2− anion. Both AgI cations are three‐coordinated in distorted T‐shaped coordination geometries. One AgI cation is coordinated by one N and two O atoms from two HPyIDC2− anions, while the other is bonded to one O and two N atoms from two HPyIDC2− anions. It is interesting to note that the HPyIDC2− group acts as a μ4‐bridging ligand to link the AgI cations into a three‐dimensional framework, which can be simplified as a diamondoid topology. The thermal stability and photoluminescent properties of the title compound have also been studied.  相似文献   

11.
In the polymeric title compound, [CuCl2(C6H6N4)]n, each CuII ion is five‐coordinated by four basal atoms (two N atoms from a 2,2′‐biimidazole mol­ecule and two Cl anions) and one axial Cl anion, in a distorted square‐pyramidal coordination geometry. Cl anions bridge the {Cu(C6H6N4)Cl} units into one‐dimensional linear chains, which are reinforced by π–π inter­actions. Adjacent linear chains are linked by N—H⋯Cl hydrogen bonds, resulting in a grid layer. The hydrogen‐bonding pattern can be described in graph‐set notation as C(9)R(9)R(14). This study extends our knowledge of the multifunctional properties of the 2,2′‐biimidazole ligand and of the coordination stereochemistry of copper(II).  相似文献   

12.
In the title compounds, C7H8NO2+·NO3, (I), C7H8NO2+·ClO4·H2O, (II), and 2C7H8NO2+·SO42−, (III), the carboxyl planes of the 4‐carboxy­phenyl­ammonium cations are twisted from the aromatic plane. A homonuclear C(8) hydrogen‐bonding motif of 4‐carboxy­phenyl­ammonium cations is observed in both (I) and (II), leading to `head‐to‐tail' layers. The cations in (III) form carboxyl group dimers, making a graph‐set motif of R22(8). In all the structures, anions connect the cationic layers and an infinite chain running along the c axis is observed, having the C22(6) graph‐set motif. Inter­estingly, in (II), the anions are connected through weak hydrogen bonds involving the water mol­ecules, leading to a graph‐set motif of R44(12). Alternate hydro­phobic and hydro­philic layers are observed in all three compounds as a result of the column‐like arrangement of the aromatic rings of the cations and the anions. Furthermore, in (I), head‐to‐tail N—H⋯O inter­actions and inter­actions linking the cations and anions form an R64(16) hydrogen‐bonding motif, resulting in a pseudo‐inversion centre at (, , 0).  相似文献   

13.
A derivative of H5ttda (=3,6,10‐tris(carboxymethyl)‐3,6,10‐triazadodecanedioic acid=N‐{2‐[bis(carboxymethyl)amino]ethyl}‐N‐{3‐[bis(carboxymethyl)amino]propyl}glycine), H5[(S)‐4‐Bz‐ttda] (=(4S)‐4‐benzyl‐3,6,10‐tris(carboxymethyl)‐3,6,10‐triazadodecanedioic acid=N‐{(2S)‐2‐[bis(carboxymethyl)amino]‐3‐phenylpropyl}‐N‐{3‐[bis(carboxymethyl)amino]propyl}glycine; 1 ) carrying a benzyl group was synthesized and characterized. The stability constants of the complexes formed with Ca2+, Zn2+, Cu2+, and Gd3+ were determined by potentiometric methods at 25.0±0.1° and 0.1M ionic strength in Me4NNO3. The observed water proton relaxivity value of [Gd{(S)‐4‐Bz‐ttda}]2− was constant with respect to pH changes over the range pH 4.5–12.0. From the 17O‐NMR chemical shift of H2O induced by [Dy{(S)‐4‐Bz‐ttda}]2− at pH 6.80, the presence of 0.9 inner‐sphere water molecules was deduced. The water proton spin‐lattice relaxation rate for [Gd{(S)‐4‐Bz‐ttda}]2− at 37.0±0.1° and 20 MHz was 4.90±0.05 mM −1 s−1. The EPR transverse electronic relaxation rate and 17O‐NMR transverse‐relaxation time for the exchange lifetime of the coordinated H2O molecule (τM), and 2H‐NMR longitudinal‐relaxation rate of the deuterated diamagnetic lanthanum complex for the rotational correlation time (τR) were thoroughly investigated, and the results were compared with those previously reported for the other lanthanide(III) complexes. The exchange lifetime (τM) for [Gd{(S)‐4‐Bz‐ttda}]2− (2.3±1.3 ns) was significantly shorter than that of the [Gd(dtpa)(H2O)]2− complex (dtpa=diethylenetriaminepentaacetic acid). The rotational correlation time τR for [Gd{(S)‐4‐Bz‐ttda}]2− (70±6 ps) was slightly longer than that of the [Gd(dtpa)(H2O)]2− complex. The marked increase of relaxivity of [Gd{(S)‐4‐Bz‐ttda}]2− mainly resulted from its longer rotational time rather than from its fast water‐exchange rate. The noncovalent interaction between human serum albumin (HSA) and the [Gd{(S)‐4‐Bz‐ttda}]2− complex containing the hydrophobic substituent was investigated by measuring the solvent proton relaxation rate of the aqueous solutions. The association constant (KA) was less than 100 M −1, indicating a weaker interaction of [Gd{(S)‐4‐Bz‐ttda}]2− with HSA.  相似文献   

14.
Carboxylate molecular crystals have been of interest due to the presence of hydrogen bonding, which plays a significant role in chemical and crystal engineering, as well as in supramolecular chemistry. Acid–base adducts possess hydrogen bonds which increase the thermal and mechanical stability of the crystal. 2,2′‐Thiodiacetic acid (Tda) is a versatile ligand that has been widely explored, employing its multidendate and chelating coordination abilities with many metals; however, charge‐transfer complexes of thiodiacetic acid have not been reported. Two salts, namely ethylenediaminium 2,2′‐thiodiacetate, C2H10N22+·C4H4O4S22−, denoted Tdaen, and 2‐aminoanilinium 2‐(carboxymethylsulfanyl)acetate, C6H9N2+·C4H5O4S, denoted Tdaophen, were synthesized and characterized by IR, 1H and 13C NMR spectroscopies, and single‐crystal X‐ray diffraction. In these salts, Tda reacts with the aliphatic (ethylenediamine) and aromatic (o‐phenylenediamine) diamines, and deprotonates them to form anions with different valencies and different supramolecular networks. In Tdaen, the divalent Tda2− anions form one‐dimensional linear supramolecular chains and these are extended into a three‐dimensional sandwich‐type supramolecular network by interaction with the ethylenediaminium cations. However, in Tdaophen, the monovalent Tda anions form one‐dimensional zigzag supramolecular chains, which are extended into a three‐dimensional supramolecular network by interaction with the 2‐aminoanilinium cations. Thus, both three‐dimensional structures display different ring motifs. The structures of these diamines, which are influenced by hydrogen‐bonded assemblies in the molecular crystals, are discussed in detail.  相似文献   

15.
In bis(2‐aminoanilinum) fumarate, 2C6H9N2+·C4H2O42−, (I), the asymmetric unit consists of two aminoanilinium cations and one fumarate dianion, whereas in 3‐methylanilinium hydrogen fumarate, C7H10N+·C4H3O4, (II), and 4‐chloroanilinium hydrogen fumarate, C6H7ClN+·C4H3O4, (III), the asymmetric unit contains two symmetry‐independent hydrogen fumate anions and anilinium cations with a slight difference in their geometric parameters; the two salts are isostructural. In (II) and (III), the carboxylic acid H atoms of the anions are disordered across both ends of the anion, with equal site occupancies of 0.50. Both the 4‐chloroanilinium cations of (III) are disordered over two orientations with major occupancies fixed at 0.60 in each case. The hydrogen fumarate anions of (II) and (III) form one‐dimensional anionic chains linked through O—H...O hydrogen bonds. Salts (II) and (III) form two‐dimensional supramolecular sheets built from R44(16), R44(18), R55(25) and C22(14) motifs extending parallel to the (010) plane, whereas in (I), an (010) sheet is formed built from two R43(13) motifs, two R22(9) motifs and an R44(18) motif.  相似文献   

16.
The salts 1‐(diaminomethylene)thiouron‐1‐ium hydrogen difluoride, C2H7N4S+·HF2, (I), and bis[1‐(diaminomethylene)thiouron‐1‐ium] hexafluoridosilicate, 2C2H7N4S+·SiF62−, (II), have both been obtained from the reaction of (1‐diaminomethylene)thiourea (HATU) with hydrofluoric acid. Both compounds contain extensive networks of N—H...F hydrogen bonds. The hydrogen difluoride salt contains four independent asymmetric [HF2] anions. In the hexafluoridosilicate salt, the centrosymmetric [SiF6]2− anion is distorted, although this distortion is not clearly correlated with the N—H...F hydrogen‐bonding network.  相似文献   

17.
The title salt, [Zn(C2N2H8)3]2[CdI4]I2, conventionally abbreviated [Zn(en)3]2[CdI4]I2, where en is ethyl­enediamine, contains discrete [Zn(en)3]2+ cations and [CdI4]2− anions with distorted octa­hedral and nearly tetra­hedral geometries, respectively, as well as uncoordinated I ions. The cation and the free I anion lie on twofold rotation axes and the [CdI4]2− anion lies on a axis in the space group I2d. The structure exhibits numerous weak inter‐ionic hydrogen bonds of two types, viz. N—H⋯I(free ion) and N—H⋯I([CdI4]2−), which support the resulting three‐dimensional framework.  相似文献   

18.
The reaction of (diaqua)(N,N′‐ethylene‐bis(salicylidiniminato)manganese(III) with aqueous sulphite buffer results in the formation of the corresponding mono sulphito complex, [Mn(Salen)(SO3)] (S‐bonded isomer) via three distinct paths: (i) Mn(Salen)(OH2)2+ + HSO3 → (k1); (ii) Mn(Salen)(OH2)2+ + SO32− → (k2); (III) Mn(Salen)(OH2)(OH) + SO32− → (k3) in the stopped flow time scale. The fact that the mono sulphito complex does not undergo further anation with SO32−/HSO3 may be attributed to the strong trans‐activating influence of the S‐bonded sulphite. The values of the rate constants (10−2ki/dm2 mol−1 s−1 at 25°C, I = 0.3 mol dm−3), ΔHi#/kJ mol−1 and ΔSi#/J K−1 mol−1 respectively are: 2.97 ± 0.27, 42.4 ± 0.2, −55.3 ± 0.6 (i = 1); 11.0 ± 0.8, 33 ± 3, −75 ± 10 (i = 2); 20.6 ± 1.9, 32.4 ± 0.2, −72.9 ± 0.6 (i = 3). The trend in reactivity (k2 > k1), a small labilizing effect of the coordinated hydroxo group (k3/k2 < 2), and substantially low values of ΔS# suggest that the mechanism of aqua ligand substitution of the diaqua, and aqua‐hydroxo complexes is most likely associative interchange (Ia). No evidence for the formation of the O‐bonded sulphito complex and the ligand isomerization in the sulphito complex, (MnIII‐OSO2 → MnIII‐SO3), ensures the selectivity of the MnIII centre toward the S‐end of the SIV species. The monosulphito complex further undergoes slow redox reaction in the presence of excess sulphite to produce MnII, S2O62− and SO42−. The formation of dithionate is a consequence of the fast dimerization of the SO3−. generated in the rate determining step and also SO42− formation is attributed to the fast scavenging of the SO3−. by the MnIII species via a redox path. The internal reduction of the MnIII centre in the monosulphito complex is insignificant. The redox reaction of the monosulphitomanganese(III) complex operates via two major paths, one involving HSO3− and the other SO32−. The electron transfer is believed to be outersphere type. The substantially negative values of activation entropies (ΔS# = −(1.3 ± 0.2) × 102 and −(1.6 ± 0.2) × 102 J K−1 mol−1 for the paths involving HSO3− and SO32− respectively) reflect a considerable degree of ordering of the reactants in the act of electron transfer. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 627–635, 1999  相似文献   

19.
In the title compound, (C6H8N4)[AuCl4]Cl, the 4,4′‐bi(1H‐pyrazol‐2‐ium) dication, denoted [H2bpz]2+, is situated across a centre of inversion, the [AuCl4] anion lies across a twofold axis passing through Cl—Au—Cl, and the Cl anion resides on a twofold axis. Conventional N—H...Cl hydrogen bonding [N...Cl = 3.109 (3) and 3.127 (3) Å, and N—H...Cl = 151 and 155°] between [H2bpz]2+ cations (square‐planar node) and chloride anions (tetrahedral node), as complementary donors and acceptors of four hydrogen bonds, leads to a three‐dimensional binodal four‐connected framework with cooperite topology (three‐letter notation pts). The framework contains channels along the c axis housing one‐dimensional stacks of square‐planar [AuCl4] anions [Au—Cl = 2.2895 (10)–2.2903 (16) Å; interanion Au...Cl contact = 3.489 (2) Å], which are excluded from primary hydrogen bonding with the [H2bpz]2+ tectons.  相似文献   

20.
Shock‐induced crystallization of the supercooled ionic liquid 1‐butyl‐3‐methyl­imidazolium hexa­fluoro­phosphate, C8H15N2+·PF6, allows for the first time precise X‐ray diffraction analysis directly pertinent to the fluid state. This inter­mediate‐chain‐length structure shows features of both short‐ and long‐chain analogs. Two types of inter­planar distances between imidazolium rings are observed. The anions are located in channels formed by the imidazolium rings and alkyl chains.  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