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 187 毫秒
1.
The two‐dimensional mixed‐ligand network catena‐[(μ‐4,4′‐bipyridine)‐bis(μ‐L‐tryptophanato‐κ3N,O,O′)‐diaqua‐dicopper(II) dinitrate] is constructed through the bridging action of both the tridentate amino carboxylato and the bidentate 4,4′‐bipyridine ligand. The enantiomeric L‐tryptophanato ligand acts as an N,O chelate towards one copper atom and bridges through the second carboxylate oxygen atom to the adjacent copper ion. Stacking of the corrugated nets creates channels which are occupied by the hydrogen‐bonded and very weakly Cu‐coordinating nitrate ions.  相似文献   

2.
A functionalized periodic mesoporous organosilica with incorporated chiral bis(cyclohexyldiamine)‐based NiII complexes within the silica framework was developed by the co‐condensation of (1R,2R)‐cyclohexyldiamine‐derived silane and ethylene‐bridge silane, followed by the complexation of NiBr2 in the presence of (1R,2R)‐N,N′‐dibenzylcyclohexyldiamine. Structural characterization by XRD, nitrogen sorption, and TEM disclosed its orderly mesostructure, and FTIR and solid‐state NMR spectroscopy demonstrated the incorporation of well‐defined single‐site bis(cyclohexyldiamine)‐based NiII active centers within periodic mesoporous organosilica. As a chiral heterogeneous catalyst, this functionalized periodic mesoporous organosilica showed high catalytic activity and excellent enantioselectivity in the asymmetric Michael addition of 1,3‐dicarbonyl compounds to nitroalkenes, comparable to those with homogeneous catalysts. In particular, this heterogeneous catalyst could be recovered easily and reused repeatedly up to nine times without obviously affecting its enantioselectivity, thus showing good potential for industrial applications.  相似文献   

3.
The self‐assembly properties of two ZnII porphyrin isomers on Cu(111) are studied at different coverage by means of scanning tunneling microscopy (STM). Both isomers are substituted in their meso‐positions by two voluminous 3,5‐di(tert‐butyl)phenyl and two rod‐like 4′‐cyanobiphenyl groups, respectively. In the trans‐isomer, the two 4′‐cyanobiphenyl groups are opposite to each other, whereas they are located at right angle in the cis‐isomer. For coverage up to one monolayer, the cis‐substituted porphyrins self‐assemble to form oligomeric macrocycles held together by antiparallel CN???CN dipolar interactions and CN???H‐C(sp2) hydrogen bonding. Cyclic trimers and tetramers occur most frequently but everything from cyclic dimers to hexamers can be observed. Upon annealing of the samples at temperatures >150 °C, dimeric macrocyclic structures are observed, in which the two porphyrins are bridged by Cu atoms, originating from the surface, under formation of two CN???Cu???NC coordination bonds. The trans‐isomer builds up linear chains on Cu(111) at low coverage, whereas for higher coverage the molecules assemble in a periodic, densely packed structure. Both cis‐ and trans‐bis(4′‐cyanobiphenyl)‐substituted ZnII porphyrins behave very differently on Cu(111) compared to similar porphyrins in literature on less reactive surfaces such as Au(111) and Ag(111). On the latter surfaces, there is no signal visible between molecular orientation and the crystal directions of the substrate, whereas on Cu(111), very strong adsorbate–substrate interactions have a dominating influence on all observed structures. This strong porphyrin–substrate interaction enables a much broader variety of structures, including also less favorable intermolecular bonding motifs and geometries.  相似文献   

4.
Anion…π interactions are newly recognized weak supramolecular forces which are relevant to many types of electron‐deficient aromatic substrates. Being less competitive with respect to conventional hydrogen bonding, anion…π interactions are only rarely considered as a crystal‐structure‐defining factor. Their significance dramatically increases for polyoxometalate (POM) species, which offer extended oxide surfaces for maintaining dense aromatic/inorganic stacks. The structures of tetrakis(caffeinium) μ12‐silicato‐tetracosa‐μ2‐oxido‐dodecaoxidododecatungsten trihydrate, (C8H11N4O2)4[SiW12O40]·3H2O, (1), and tris(theobrominium) μ12‐phosphato‐tetracosa‐μ2‐oxido‐dodecaoxidododecatungsten ethanol sesquisolvate, (C7H9N4O2)3[PW12O40]·1.5C2H5OH, (2), support the utility of anion…π interactions as a special kind of supramolecular synthon controlling the structures of ionic lattices. Both caffeinium [(HCaf)+ in (1)] and theobrominium cations [(HTbr)+ in (2)] reveal double stacking patterns at both axial sides of the aromatic frameworks, leading to the generation of anion…π…anion bridges. The latter provide the rare face‐to‐face linkage of the anions. In (1), every square face of the metal–oxide cuboctahedra accepts the interaction and the above bridges yield flat square nets, i.e. {(HCaf+)2[SiW12O40]4?}n. Two additional cations afford single stacks only and they terminate the connectivity. Salt (2) retains a two‐dimensional (2D) motif of square nets, with anion…π…anion bridges involving two of the three (HTbr)+ cations. The remaining cations complete a fivefold anion…π environment of [PW12O40]3?, acting as terminal groups. This single anion…π interaction is influenced by the specific pairing of (HTbr)+ cations by double amide‐to‐amide hydrogen bonding. Nevertheless, invariable 2D patterns in (1) and (2) suggest the dominant role of anion…π interactions as the structure‐governing factor, which is applicable to the construction of noncovalent linkages involving Keggin‐type oxometalates.  相似文献   

5.
To model the structures of dissolved uranium contaminants adsorbed on mineral surfaces and further understand their interaction with geological surfaces in nature, we have performed periodic density funtional theory (DFT) calculations on the sorption of uranyl species onto the TiO2 rutile (110) surface. Two kinds of surfaces, an ideal dry surface and a partially hydrated surface, were considered in this study. The uranyl dication was simulated as penta‐ or hexa‐coordinated in the equatorial plane. Two bonds are contributed by surface bridging oxygen atoms and the remaining equatorial coordination is satisfied by H2O, OH?, and CO32? ligands; this is known to be the most stable sorption structure. Experimental structural parameters of the surface–[UO2(H2O)3]2+ system were well reproduced by our calculations. With respect to adsorbates, [UO2(L1)x(L2)y(L3)z]n (L1=H2O, L2=OH?, L3=CO32?, x≤3, y≤3, z≤2, x+y+2z≤4), on the ideal surface, the variation of ligands from H2O to OH? and CO32? lengthens the U? Osurf and U? Ti distances. As a result, the uranyl–surface interaction decreases, as is evident from the calculated sorption energies. Our calculations support the experimental observation that the sorptive capacity of TiO2 decreases in the presence of carbonate ions. The stronger equatorial hydroxide and carbonate ligands around uranyl also result in U?O distances that are longer than those of aquouranyl species by 0.1–0.3 Å. Compared with the ideal surface, the hydrated surface introduces greater hydrogen bonding. This results in longer U?O bond lengths, shorter uranyl–surface separations in most cases, and stronger sorption interactions.  相似文献   

6.
The asymmetric unit of the title compound, [Pb2(C8H4O4)2(C18H11N5)2]n, contains two PbII atoms, two benzene‐1,4‐dicarboxylate (1,4‐bdc) dianions and two 6‐(4‐pyridyl)‐5H‐imidazolo[4,5‐f][1,10]phenanthroline (L) ligands. Each PbII atom is eight‐coordinated by three N atoms from two different L ligands and five carboxylate O atoms from three different 1,4‐bdc dianions. The two 1,4‐bdc dianions (1,4‐bdc1 and 1,4‐bdc2) show different coordination modes. Each 1,4‐bdc1 coordinates to two PbII atoms in a chelating bis‐bidentate mode. Each carboxylate group of the 1,4‐bdc2 anion connects two PbII atoms in a chelating–bridging tridentate mode to form a dinuclear unit. Neighbouring dinuclear units are connected together by the aromatic backbone of the 1,4‐bdc dianions and the L ligands into a three‐dimensional six‐connected α‐polonium framework. The most striking feature is that two identical three‐dimensional single α‐polonium nets are interlocked with each other, thus leading directly to the formation of a twofold interpenetrated three‐dimensional α‐polonium architecture. The framework is held together in part by strong N—H...O hydrogen bonds between the imidazole NH groups of the L ligands and the carboxylate O atoms of 1,4‐bdc dianions within different α‐polonium nets.  相似文献   

7.
Highly transparent ultrathin films (UTFs) based on alternative layer‐by‐layer assembly of Eu‐ and Tb‐based lanthanide complexes (LCs) and Mg–Al‐layered double hydroxide (LDH) nanosheets are reported herein. UV–visible absorption and fluorescence spectroscopy showed an orderly growth of the two types of ultrathin films upon increasing the number of deposition cycles. AFM and SEM measurements indicate that the films feature periodic layered structures as well as uniform surface morphology. Luminescent investigations reveal that (LCs/LDH)n UTFs can detect Fe3+ with relative selectivity and high sensitivity (Stern–Volmer constant KSV=8.43×103 L mol?1); this suggests that (LCs/LDH)n UTFs could be a promising luminescent probe for selectively sensing Fe3+ ion.  相似文献   

8.
A series of gradient and block copolymers, based on 2‐(2‐methoxyethoxy)ethyl methacrylate (MEO2MA) and tert‐butyl acrylate (tBA), were synthesized by atom transfer radical polymerization (ATRP) in a first step. The MEO2MA monomer leads to the production of thermosensitive polymers, exhibiting lower critical solution temperature (LCST) at around room temperature, which could be adjusted by changing the proportion of tBA in the copolymer. In a second step, the tert‐butyl groups of tBA were hydrolyzed with trifluoroacetic acid to form the corresponding block and gradient copolymers of MEO2MA and acrylic acid (AA), which exhibited both temperature and pH‐responsive behavior. These copolymers showed LCST values strongly dependent on the pH. At acid pH, a slightly decrease of LCST with an increase of AA in the copolymer was observed. However, at neutral or basic conditions, ionization of acid groups increases the hydrophilic balance considerably raising the LCST values, which even become not observable over the temperature range under study. In the last step, these carboxylic functionalized copolymers were covalently bound to biocompatible and biodegradable films of poly(3‐hydroxybutyrate‐co‐3‐hydroxyhexanoate) [P(HB‐co‐HHx)] obtained by casting and, previously treated with ethylenediamine (ED) to render their surfaces with amino groups. Thereby, thermosensitive surfaces of modified P(HB‐co‐HHx) could be obtained. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2011  相似文献   

9.
New complexes containing the 1,4‐bis(pyridazin‐4‐yl)benzene ligand, namely diaquatetrakis[1,4‐bis(pyridazin‐4‐yl)benzene‐κN2]cadmium(II) hexaiodidodicadmate(II), [Cd(C14H10N4)4(H2O)2][Cd2I6], (I), and poly[[μ‐1,4‐bis(pyridazin‐4‐yl)benzene‐κ2N2:N2′]bis(μ‐thiocyanato‐κ2N:S)cadmium(II)], [Cd(NCS)2(C14H10N4)]n, (II), demonstrate the adaptability of the coordination geometries towards the demands of slipped π–π stacking interactions between the extended organic ligands. In (I), the discrete cationic [Cd—N = 2.408 (3) and 2.413 (3) Å] and anionic [Cd—I = 2.709 (2)–3.1201 (14) Å] entities are situated across centres of inversion. The cations associate via complementary O—H...N2′ hydrogen bonding [O...N = 2.748 (4) and 2.765 (4) Å] and extensive triple π–π stacking interactions between pairs of pyridazine and phenylene rings [centroid–centroid distances (CCD) = 3.782 (4)–4.286 (3) Å] to yield two‐dimensional square nets. The [Cd2I6]2− anions reside in channels generated by packing of successive nets. In (II), the CdII cation lies on a centre of inversion and the ligand is situated across a centre of inversion. A two‐dimensional coordination array is formed by crosslinking of linear [Cd(μ‐NCS)2]n chains [Cd—N = 2.3004 (14) Å and Cd—S = 2.7804 (5) Å] with N2:N2′‐bidentate organic bridges [Cd—N = 2.3893 (12) Å], which generate π–π stacks by double‐slipped interactions between phenylene and pyridazine rings [CCD = 3.721 (2) Å].  相似文献   

10.
The title compound, C4H10N3O2+·C2H2NO3·C2H3NO3, contains at least 11 distinct hydrogen‐bond interactions showing a great variety of bond strengths. The shortest and strongest hydrogen bond [O...O = 2.5004 (12) Å] is found between the uncharged oxamic acid molecule and the oxamate monoanion. The grouping formed by such a strong hydrogen bond can thus be considered as a hydrogen bis(oxamate) monoanion. It lacks crystallographic symmetry and the two oxamate groups have different conformations, showing an asymmetric hydrogen‐bond interaction. Significantly, the asymmetry allows us to draw a direct comparison of site basicity for the two inequivalent carboxylate O atoms in the planar oxamate anion. The constituent molecular ions of (I) form ribbons, where all amide and carboxylate groups are coplanar. Graph‐set analysis of the hydrogen‐bonded networks reveals the R22(10) and R22(9) homodromic nets as important structure‐directing motifs, which appear to be a common feature of many oxamate‐containing compounds.  相似文献   

11.
Three new mixed‐ligand coordination polymers of CuII, namely, [Cu(Fbtx)(L1)(H2O)]n ( 1 ), [Cu(Fbtx)0.5(HL2)(H2O)2]n ( 2 ), and {[Cu(Fbtx)1.5(HL3)(H2O)] · H2O}n ( 3 ) [Fbtx = 2,3,5,6‐tetrafluoro‐1,4‐bis(1,2,4‐triazole‐1‐ylmethyl)benenze, H2L1 = terephthalic acid, H3L2 = trimesic acid, NaH2L3 = 5‐sulfoisophthalic acid monosodium salt], were hydrothermally synthesized and structurally characterized by elemental analysis, IR spectra, and single‐crystal and powder X‐ray diffraction techniques. All the complexes have a two‐dimensional (2D) coordination layer structure. Of these, 1 displays a planar 44‐ sql structure whereas both 2 and 3 are highly undulated 63‐ hcb nets. Moreover, their thermal stability and catalytic behaviors in the aerobic oxidation of 4‐methoxybenzyl alcohol were also investigated as well. The results indicate that the benzene dicarboxylate ligands have an effective influence on the structures and catalytic properties of the resulting coordination polymers.  相似文献   

12.
To synthesize polyesters and periodic copolymers catalyzed by nonafluorobutanesulfonimide (Nf2NH), we performed ring‐opening copolymerizations of cyclic anhydrides with tetrahydrofuran (THF) at 50–120 °C. At high temperature (100–120 °C), the cyclic anhydrides, such as succinic anhydride (SAn), glutaric anhydride (GAn), phthalic anhydride (PAn), maleic anhydride (MAn), and citraconic anhydride (CAn), copolymerized with THF via ring‐opening to produce polyesters (Mn = 0.8–6.8 × 103, Mn/Mw = 2.03–3.51). Ether units were temporarily formed during this copolymerization and subsequently, the ether units were transformed into esters by chain transfer reaction, thus giving the corresponding polyester. On the other hand, at low temperature (25–50 °C), ring‐opening copolymerizations of the cyclic anhydrides with THF produced poly(ester‐ether) (Mn = 3.4–12.1 × 103, Mw/Mn = 1.44–2.10). NMR and matrix‐assisted laser desorption/ionization time‐of‐flight mass spectra revealed that when toluene (4 M) was used as a solvent, GAn reacted with THF (unit ratio: 1:2) to produce periodic copolymers (Mn = 5.9 × 103, Mw/Mn = 2.10). We have also performed model reactions to delineate the mechanism by which periodic copolymers containing both ester and ether units were transformed into polyesters by raising the reaction temperature to 120 °C. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

13.
PrSeTe2, an Ordered Ternary Polychalcogenid with NdTe3 Structure Single crystals of PrSeTe2 have been obtained by reaction of the elements in a LiCl/RbCl flux at 970 K during 7 days. PrSeTe2 crystallizes in space group Cmcm (No. 63), with four formula units per unit cell. The lattice constants are a = 426.1(1) pm, b = 2506.0(5) pm, and c = 426.0(1) pm. The crystal structure is an ordered ternary variant of the NdTe3 type. It consists of a puckered double layer of praseodymium and selenium atoms [PrSe] sand wiched by two square planar layers of tellurium atoms [Te] yielding a stacking —[Te]—[Te]—[PrSe]— along [010]. The Te atoms build regular 44 nets with Te—Te distances of 301, 3(1) pm. DFT calculations propose that this compounds should be metallic mainly due to contributions of the Pr f‐electrons. The band structure shows no significance for a distortion in the [Te]—nets.  相似文献   

14.
In the title compound, [Cd(C8H4O4)(C10H8N2O2)(H2O)]n, (I), each CdII atom is seven‐coordinated in a distorted monocapped trigonal prismatic coordination geometry, surrounded by four carboxylate O atoms from two different benzene‐1,4‐dicarboxylate (1,4‐bdc) anions, two O atoms from two distinct 4,4′‐bipyridine N,N′‐dioxide (bpdo) ligands and one water O atom. The CdII atom and the water O atom are on a twofold rotation axis. The bpdo and 1,4‐bdc ligands are on centers of inversion. Each crystallographically unique CdII center is bridged by the 1,4‐bdc dianions and bpdo ligands to give a three‐dimensional diamond framework containing large adamantanoid cages. Three identical such nets are interlocked with each other, thus directly leading to the formation of a threefold interpenetrated three‐dimensional diamond architecture. To the best of our knowledge, (I) is the first example of a threefold interpenetrating diamond net based on both bpdo and carboxylate ligands. There are strong linear O—H...O hydrogen bonds between the water molecules and carboxylate O atoms within different diamond nets. Each diamond net is hydrogen bonded to its two neighbors through these hydrogen bonds, which further consolidates the threefold interpenetrating diamond framework.  相似文献   

15.
The crystal structure of [C10N2H10][ZnCl(HPO4)]2 contains corrugated tetrahedral layers with 63 topology. Charge balance is achieved by insertion of diprotonated 4,4′‐bipyridine between the layers. Crystal data: monoclinic, P21/n (no. 14), a = 4.8832(2) Å, b = 22.673(2) Å, c = 8.1643(4) Å, β = 104.02(1)°; V = 877.0(1) Å3; Z = 4; R1 = 0.041 and wR2 = 0.088 for 1836 reflections [I > 2σ(I)]. Tetrahedral layers are also observed in other organo‐ammonium templated compounds. However, their topologies are characterized by 4.82 nets. With the title compound a layered tetrahedral net with 63‐topology is reported for the first time.  相似文献   

16.
Two metalloporphyrin octacarboxylates were used to link copper(II) nodes for the formation of two novel porous mixed‐metal metal–organic frameworks (M′MOFs) containing nanopore cages (2.1 nm in diameter) or nanotubular channels (1.5 nm in diameter). The highly active Cu2+ sites on the nanotubular surfaces of the stable porous M′MOF ZJU‐22 , stabilized by three‐connected nets, lead to the superior catalytic activity for the cross‐dehydrogenative coupling (CDC) reaction.  相似文献   

17.
A two‐step wet chemistry protocol has been developed for the surface derivatization of poly(ethylene terephthalate) (PET) track‐etched membrane used as cell culturing support, that is, (a) activation by trifluorotriazine (1 M in acetonitrile (ACN), 30 °C, 3 h); (b) coupling to amine‐terminated molecules, namely 3,5‐bis(trifluoromethyl)benzylamine ((F)Tag), (L)‐4,5‐[3H]‐lysine, and Gly‐Arg‐Gly‐Asp‐Ser (GRGDS) pentapeptide (10?3 M in PB‐ACN, 1:1 (v/v), 20 °C, 17 h). The grafting rates determined by X‐ray photoelectron spectroscopy, from the F/C and N/C atomic ratios, are in the range of 100–140 pmol/cm2 (apparent surface), whereas the liquid scintillation counting assays give higher values (180–230 pmol/cm2) corresponding to the open surface reactivity. PET‐g‐(F)Tag is reasonably stable under two usual sterilization conditions of biomaterials, that is, steam heating at 121 °C and γ‐irradiation at 25 kGy. On the other hand, PET‐g‐GRGDS is found to be stable only under ionization radiation (84% of remaining peptide molecules), but damaged in a large extent by the autoclave treatment (23% of remaining peptide molecules). The surfaces of the sterilized PET and PET‐g‐GRGDS samples have been characterized by water contact angle measurement and by atomic force microscopy analysis in air and under water. Comparatively to the corresponding nonsterilized surfaces, γ‐irradiated surfaces are slightly more hydrophilic and also slightly more rough and jagged. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 195–208, 2010  相似文献   

18.
The spreading behaviour of defined trisiloxane surfactants of general formula [(CH3)3SiO]2 CH3Si(CH2)3(OCH2CH2) nOCH3 (n = 3–9) on five different solid surfaces has been investigated. Maximum spreading areas and rates are found on non‐polar or slightly polar surfaces of 30 to 40 mN m−1 surface energy. Extremely low or high surface energies substantially reduce the spreading rates. On non‐polar surfaces rapid spreading is observed for 1 wt % solutions of the relatively short‐chained penta‐ and hexa‐ethylene glycol derivatives. On slightly polar surfaces dilute 0.1 wt % solutions of longer‐chained derivatives spread faster. This spreading pattern shift coincides with a change of the phase behaviour. Solutions of Silwet L77 do not prefer one specific surface, since 1 wt % solutions abruptly stop spreading after a few seconds and the maximum spreading rates are found for 0.1 wt % solutions. Therefore, Silwet L77 essentially belongs among the long‐chained derivatives. Copyright © 2000 John Wiley & Sons, Ltd.  相似文献   

19.
In recent years, the design and construction of crystalline coordination complexes by the assembly of metal ions with multitopic ligands have attracted considerable attention because of the unique architectures and potential applications of these compounds. Two new coordination polymers, namely poly[[μ‐trans‐1‐(2‐aminopyridin‐3‐yl)‐2‐(pyridin‐4‐yl)ethene‐κ2N:N′](μ3‐5‐methylisophthalato‐κ4O1,O1′:O3:O3′)cadmium(II)], [Cd(C9H6O4)(C12H11N3)]n or [Cd(5‐Me‐ip)(2‐NH2‐3,4‐bpe)]n, ( I ), and poly[[μ‐trans‐1‐(2‐aminopyridin‐3‐yl)‐2‐(pyridin‐4‐yl)ethene‐κ2N:N′](μ2‐5‐hydroxyisophthalato‐κ4O1,O1′:O3:O5)cadmium(II)], [Cd(C8H4O5)(C12H11N3)]n or [Cd(5‐HO‐ip)(2‐NH2‐3,4‐bpe)]n, ( II ), have been prepared hydrothermally by the self‐assembly of Cd(NO3)2·4H2O and trans‐1‐(2‐aminopyridin‐3‐yl)‐2‐(pyridin‐4‐yl)ethene (2‐NH2‐3,4‐bpe) with two similar dicarboxylic acids, i.e. 5‐methylisophthalic acid (5‐Me‐H2ip) and 5‐hydroxyisophthalic acid (5‐HO‐H2ip). The coordination network of ( I ) is a two‐dimensional sql net parallel to (101). Adjacent sql nets are further linked to form a three‐dimensional supramolecular framework via hydrogen‐bonding interactions. Compound ( II ) is a two‐dimensional (3,5)‐connected coordination network parallel to (010) with the point symbol (63)(55647). As the other reactants and reaction conditions are the same, the structural differences between ( I ) and ( II ) are undoubtedly determined by the different substituent groups in the 5‐position of isophthalic acid. Both ( I ) and ( II ) exhibit good thermal stabilities and photoluminescence properties.  相似文献   

20.
The asymmetric unit of the title compound, [Cd(C8H4O4)(C17H8ClN5)(H2O)]n, contains one CdII atom, two half benzene‐1,4‐dicarboxylate (1,4‐bdc) anions, one 11‐chloropyrido[2′,3′:2,3]pyrimidino[5,6‐f][1,10]phenanthroline (L) ligand and one coordination water molecule. The 1,4‐bdc ligands are on inversion centers at the centroids of the arene rings. The CdII atom is six‐coordinated by two N atoms from one L ligand, three carboxylate O atoms from two different 1,4‐bdc ligands and one water O atom in a distorted octahedral coordination sphere. Each CdII center is bridged by the 1,4‐bdc dianions to give a one‐dimensional chain. π–π stacking interactions between L ligands of neighboring chains extend adjacent chains into a two‐dimensional supramolecular (6,3) network. Neighboring (6,3) networks are interpenetrated in an unusual inclined mode, resulting in a three‐dimensional framework. Additionally, the water–carboxylate O—H...O hydrogen bonds observed in the network consolidate the interpenetrating nets.  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