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 46 毫秒
1.
手性金属-有机骨架材料(MOFs)作为一种新型多孔材料,由于具有比表面积大、结构多样、孔尺寸可调和化学稳定性良好等特点,而备受色谱分离领域的关注。该文以(1R,2R)-1,2-环己烷二甲酸(H2L)和4,4'-联吡啶(bpy)为配体与铜离子反应,通过溶剂热法合成了一种具有二维手性网状结构的手性MOF[Cu_3(HL)_2(L)_2(bpy)_3]·4H_2O。将该手性MOF作为手性固定相制备了高效液相色谱柱。为了考察MOF[Cu_3(HL)_2(L)_2(bpy)_3]·4H_2O的手性识别能力,在正相色谱(流动相:正己烷-异丙醇)条件下,对一系列外消旋化合物在手性MOF柱上进行了拆分。该手性MOF柱对醇类、酮类、酸类、环氧化合物和醚类等10种手性化合物表现出较好的拆分效果。对该手性MOF柱的重现性、稳定性作了评价,考察了进样量对分离效果的影响,结果表明该手性MOF柱具有较好的重现性和稳定性。  相似文献   

2.
Zhang J  Chew E  Chen S  Pham JT  Bu X 《Inorganic chemistry》2008,47(9):3495-3497
Reported here are four homochiral and one racemic chiral compounds assembled from transition metals (Fe2+, Co2+, and Ni2+), rigid enantiopure camphoric acid, and flexible 4,4'-trimethylenedipyridine; they feature an unusual homochiral three-dimensional supramolecular assembly (resulting from catenation of 4-connected bilayers) as well as a homochiral 3-fold interpenetrating diamond net and a noninterpenetrating primitive cubic net.  相似文献   

3.
Gu X  Xue D 《Inorganic chemistry》2006,45(23):9257-9261
Two novel homochiral lanthanide-silver heterometallic coordination polymers LnAg(OAc)(IN)3 [Ln = Nd (1), Eu (2), HIN = isonicotinic acid, HOAc = acetic acid] have been prepared under hydrothermal conditions, which were characterized by elemental analysis, infrared, thermogravimetric analysis, and single-crystal X-ray diffraction. Both complexes are isostructural and crystallize in a hexagonal system, chiral space group P6(1)22. Both polymers are constructed from infinite right-handed homochiral helical chains with Ln-O-Ag connectivity, representing the first examples of homochiral lanthanide-transition metal heterometallic coordination polymers with a 3D coordination framework based on spontaneous resolution. Furthermore, the luminescent properties of 2 were studied.  相似文献   

4.
Spontaneous separation of chiral phases was observed in the monolayers of a racemate of gemini-type twin-tailed, twin-chiral amphiphiles, (2R,3R)-(+)-bis(decyloxy)succinic acid and (2S,3S)-(-)-bis(decyloxy)succinic acid. The pressure-area isotherms of the interfacial monolayers formed at the liquid-air interface, and the 2D lattice structures studied through surface probe measurements revealed that the racemate exhibits a homochiral discrimination of the enantiomers in two dimensions. An enantiomeric excess (e,e) of 20% was sufficient to break the chiral symmetry at the air-water interface for a homochiral interaction. Langmuir monolayers on ZnCl2 and CaCl2 subphases manifested chiral discrimination with Zn2+ evidencing homochiral interaction with a chelate-type complex, whereas Ca2+ resulted in a heterochiral interaction forming an ionic-type complex. For the chiral asymmetric units, oblique and rectangular unit cells of the racemic monolayer had exclusive requirements of homo- and heterochiral recognitions for Zn2+ and Ca2+ ions, respectively. Monolayers transferred from the condensed phase at 25 mN/m onto hydrophilic Si(100) and quartz substrates revealed the formation of bilayers through transfer-induced monolayer buckling. The emergence of homochiral discrimination was explained using the effective-pair-potential (EPP) approach.  相似文献   

5.
The effect of a counteranion on chiral recognition inducing conglomerate crystallization of a cobalt(III) complex is reported. An achiral tripodal ligand involving three imidazole groups, tris{[2-{(imidazol-4-yl)methylidene}amino]ethyl}amine (H3L), was prepared by condensation of tris(2-aminoethyl)amine and 4-formylimidazole in a 1:3 mole ratio. The reaction of H3L and trans-[CoIIICl2(py)4]+ afforded the chiral (Delta or Lambda) [CoIII(H3L)]3+ complex. The formally hemideprotonated complexes [CoIII(H(1.5)L)]X(1.5).nH2O (where X = Cl, Br, I, BF4, ClO4, or PF6) were synthesized by controlled deprotonation of the uncoordinated imidazole NH groups of [Co(H3L)]3+. In crystals of the hemideprotonated complex, two components, [Co(H3L)]3+ and [Co(L)], with the same absolute configuration are linked by imidazole-imidazolate hydrogen bonds to form an extended homochiral 2D sheet structure, which is composed of a hexanuclear unit with a trigonal void. There are two ways of stacking the sheets: One is via homochiral stacking, and the other is via heterochiral stacking. When the size of the counterion is small (i.e., X = Cl, Br, I, or BF4), adjacent homochiral sheets with the same chirality are stacked to form a homochiral crystal (conglomerate). With large anions (i.e., ClO4- and PF6-), a homochiral sheet consisting of Delta enantiomers and a sheet consisting of Lambda enantiomers are stacked alternately to give a heterochiral crystal (a racemic crystal). Optically active Lambda-[Co(H(1.5)L)](ClO4)(1.5).H2O was synthesized from Lambda-[Co(H3L)]3+, and the crystal structure was compared to that of the racemic complex. There are two conflicting factors governing the crystal structure: the skeletal density; the size of the channels. The 2D sheets are more closely packed in the homochiral crystal than in the heterochiral crystal. However, the channels, where the counterions are accommodated, are smaller in the homochiral crystal, and the steric congestion between the anions increases with increasing anion size. The heterochiral crystal has a flexible, zigzag channel structure, and the size of the channels can increase to accommodate larger anions. Thus, complexes with large anions (i.e., ClO4- and PF6-) preferentially form heterochiral crystals rather than homochiral crystals.  相似文献   

6.
We show that the racemic states of the B 2 phase of liquid crystals composed of banana-shaped molecules do not satisfy the Curie principle. Thus it is argued that these states cannot exist in bulk samples and the homochiral states constitute the only stable microscopic structures. A reinterpretation of the racemic states with the same macroscopic optical behaviour is proposed in terms of mixtures of the homochiral structures.  相似文献   

7.
Unique perylene diastereomeric linear and cyclic dimers were synthesized from twisted perylene monomers, revealing that pi-stacking stereoisomerism imparted specific intermolecular self-assembly and intramolecular folding. Only the homochiral twisted tetrachloroperylene monomers cyclized via a cooperative reaction, forming the homochiral diastereomers. The heterochiral tetrachloroperylene monomers proceeded through a stepwise reaction and yielded a linear heterochiral dimer, which equilibrated with the linear homochiral dimers. The linear homochiral dimers cyclized to produce the same cyclic homochiral diastereomers. These results demonstrated that homochiral and heterochiral self-assemblies were two distinct molecular codes, directing two specific chemical pathways. The homochiral cyclic dimers remain isomerically pure at -20 degrees C but can be interconverted to the heterochiral cyclic dimer meso compound at room temperature. The diastereomers were readily separated by HPLC. While driven by solvophobic forces, foldable linear dimers synthesized from the same twisted monomers using phosphoramidite chemistry folded into homodimer and heterodimer, confirming the inherent molecular codes, which were dictated by the perylene chirality, ultimately gauged the weak pi-stack forces, and directed self-assembly and folding.  相似文献   

8.
Substitution of the pillaring ligand in the homochiral open-framework [Ni(2)(L-asp)(2)(bipy)] by extended bipy-type ligands leads to a family of layer-structured, homochiral metal-organic frameworks. The 1D channel topology can be modified by the nature of the organic linker, with shape, cross-section and the chemical functionality tuneable. In addition, the volume of these channels can be increased by up to 36 % compared to the parent [Ni(2)(L-asp)(2)(bipy)]. The linker 1,4-dipyridylbenzene (3rbp) gives access to a new layered homochiral framework [Ni(2)(L-asp)(2)(3rbp)] with channels of a different shape. In specific cases, non-porous analogues with the linker also present as a guest can be activated to give porous materials after sublimation. Their CO(2) uptake shows an increase of up to 30 % with respect to the parent [Ni(2)(L-asp)(2)(bipy)] framework.  相似文献   

9.
Reaction of Cu(ClO(4))(2) x 6H(2)O with a racemic mixture of the novel chiral ligand N-(1,2-bis(2-pyridyl)ethyl)pyridine-2-carboxamide (PEAH) affords only the homochiral dimeric copper(II) complexes [Cu(2)((R)()PEA)(2)](ClO(4))(2) and [Cu(2)((S)()PEA)(2)](ClO(4))(2) in a 1:1 ratio. The phenomenon of molecular self-recognition is also observed when a racemic mixture of the monomeric copper(II) complex [Cu((R(S))()PEA)(Cl)(H(2)O)] is converted into the homochiral dimeric species [Cu(2)((R(S))()PEA)(2)](ClO(4))(2) via reaction with Ag(+) ion. This is the first report of direct conversion of a racemic mixture of a chiral monomeric copper(II) complex to a mixture of the homochiral dimers.  相似文献   

10.
We show that the racemic states of the B 2 phase of liquid crystals composed of banana-shaped molecules do not satisfy the Curie principle. Thus it is argued that these states cannot exist in bulk samples and the homochiral states constitute the only stable microscopic structures. A reinterpretation of the racemic states with the same macroscopic optical behaviour is proposed in terms of mixtures of the homochiral structures.  相似文献   

11.
A new strategy for creating homochiral metal–organic frameworks through a fusion of pillaring and templating concepts is demonstrated. This strategy makes use of the synergy among various chemical interactions during self‐assembly processes, and leads to the synthesis of a series of homochiral frameworks. In the presence of only pillar‐to‐pillar π–π interactions, inter‐pillar forces compete against metal–pillar interactions, resulting in mismatch between pillar‐to‐pillar and metal‐to‐metal separations and consequently 2D materials without pillaring. To create 3D materials, a method was developed to use various aromatic molecules, polycyclic aromatic hydrocarbons in particular, as templates to modulate the inter‐pillar interaction and separation, leading to the formation of 3D homochiral frameworks. The use of aromatic molecules, especially hydrocarbons, as structure‐directing agents, represents a new approach in the development of crystalline porous materials. Aromatic templates can be post‐synthetically extracted to yield flexible porous homochiral materials with gate‐opening gas sorption behaviors for both N2 and CO2 at partial pressures tunable by temperature.  相似文献   

12.
By incorporating an achiral diazine ligand, 2-pyridylmethylketazine, which can be locked in a chiral conformation upon coordination, into the manganese(II)-azido system, we induced a homochiral 2D network, in which neighboring Mn(II) ions are bridged via a diazine and two end-on azido ligands into chiral dimeric units, and neighboring units are interlinked via single end-to-end azido bridges. The interdimer chirality preservation is achieved via the homochiral 1D helical linkage formed by Mn(II) and end-to-end azido ions. The 2D layers are stacked in hetero- and homochiral fashion to yield simultaneously racemic and chiral crystals, indicating a partial spontaneous resolution. Both compounds behave as spin-canted weak ferromagnets, but the critical temperatures are different.  相似文献   

13.
《Chemistry & biology》1997,4(4):309-320
Background: Why did Nature choose furanosyl-RNA and not pyranosyl-RNA as her molecular genetic system? An experimental approach to this problem is the systematic comparison of the two isomeric oligonucleotide systems with respect to the chemical properties that are fundamental to the biological role of RNA, such as base pairing and nonenzymic replication. Pyranosyl-RNA has been found to be not only a stronger, but also a more selective pairing system than natural RNA; both form hairpin structures with comparable ease. Base sequences of pyranosyl-RNA can be copied by template-controlled replicatioe ligation of short activated oligomers (e.g. tetramer-2′,3′-cyclophosphates) under mild and potentially natural conditions. The copying proceeds with high regio-selectivity as well as chiroselectivity: homochiral template sequences mediate the formation of the correct (4′→2′)-phosphodiester junction between homochiral tetramer units provided they have the same sense of chirality as the template. How could homochiral template sequences assemble themselves in the first place?Results: Higher oligomers of pyranosyl-RNA can self-assemble in dilute solutions under mild conditions by ligative oligomerization of tetramer-2′,3′-cyclophosphates containing hemi self-complementary base sequences. The only side reaction that effectively competes with ligation is hydrolytic deactivation of 2′,3′-cyclophosphate end groups. The ligation reaction is highly chiroselective; it is slower by at least two orders of magnitude when one of the (d)-ribopyranosyl units of a homochiral (d)-tetramer-2′,3′-cyclophosphate is replaced by a corresponding (l)-unit, except when the (l)-unit is at the 4′ end of the tetramer and carries a purine, when the oligomerization rate can be ∼ 10% of that shown for a homochiral isomer. The oligomerization of homochiral tetramers is not, or only weakly, inhibited by the presence of the non-oligomerizing diastereomers.Conclusions: Available data on the chiroselective self-directed oligomerization of tetramer-2′,3′-cyclophosphates allow us to extrapolate that sets of tetramers with different but mutually fitting base sequences can be expected to co-oligomerize stochastically and generate sequence libraries consisting of predominantly homochiral (d)- and (l)-oligomers, starting from the racemic mixture of tetramers containing all possible diastereomers. Such a capability of an oligonucleotide system deserves special attention in the context of the problem of the origin of biomolecular homochirality: breaking molecular mirror symmetry by de-racemization is an intrinsic property of such a system whenever the constitutional complexity of the products of co-oligomerization exceeds a critical level.  相似文献   

14.
1,4-Diazabicyclo[2.2.2]octane (dabco) and its derivatives have been extensively utilized as building units of excellent molecular ferroelectrics for decades. However, the homochiral dabco-based ferroelectric remains a blank. Herein, by adding a methyl (Me) group accompanied by the introduction of homochirality to the [H2dabco]2+ in the non-ferroelectric [H2dabco][TFSA]2 (TFSA=bis(trifluoromethylsulfonyl)ammonium), we successfully designed enantiomeric ferroelectrics [R and S-2-Me-H2dabco][TFSA]2. The two enantiomers show two sequential phase transitions with transition temperature (Tc) as high as 405.8 K and 415.8 K, which is outstanding in both dabco-based ferroelectrics and homochiral ferroelectrics. To our knowledge, [R and S-2-Me-H2dabco][TFSA]2 are the first examples of dabco-based homochiral ferroelectrics. This finding opens an avenue to construct dabco-based homochiral ferroelectrics and will inspire the exploration of more eminent enantiomeric molecular ferroelectrics.  相似文献   

15.
This communication asks whether homochiral or heterochiral interaction is preferred between enantiomeric beta-sheets and finds that homochiral pairing is strongly preferred. Interactions between beta-sheets occur widely among proteins through pairing of the hydrogen-bonding edges. Although the hydrogen-bonding edges of both l- and d-beta-sheets put forth the same pattern of hydrogen-bond donor and acceptor groups, the side chains point in opposite directions. Homochiral pairing of beta-sheets generates structures in which the pleats and side chains of adjacent beta-strands are parallel to each other, while heterochiral pairing of beta-sheets generates structures in which the pleats and side chains are antiparallel. To test which pairing is preferred, we have prepared and studied the interactions of beta-sheets 1a-d, which comprise all l-amino acids, and beta-sheets 2a-c, which comprise all d-amino acids. Previous studies in our laboratory have established that these compounds form well-defined dimers in organic solvents. In the current study, 1H NMR experiments establish that when the l-beta-sheets (1) are mixed with the enantiomeric d-beta-sheets (2), homochiral beta-sheet dimers predominate, and only small quantities of heterochiral beta-sheet dimers form. Ratios of homochiral and heterochiral dimers ranging from 95.8:4.2 to 98.5:1.5 are measured in CDCl3 at 253 K, which correspond to statistically corrected free-energy differences of 3.1-4.2 kcal/mol (0.6-0.8 kcal/mol per interacting residue). Possible explanations for the high enantioselectivity of molecular recognition between beta-sheets include favorable nonbonded contacts between the adjacent beta-strands of the homochiral beta-sheets and poor fit of the heterochiral beta-strands, which should twist in opposite directions.  相似文献   

16.
A 2D homochiral inorganic-organic framework {[Mn(NPTA)(4,4'-bpy)(H(2)O)]·(H(2)O)(2)}(n) was prepared by assembling achiral polar 4-nitrophthalic acid, manganese ions, and ancillary 4,4'-bipyridine ligands (NPTA = 4-nitrophthalate) (4,4'-bpy = 4,4'-bipyridine). The isomorphous Zn(ii) compound was also prepared as a diamagnetic analogue. Adjacent manganese spin centres are linked by the syn-anti carboxylate to form a helical chain, and chains of the same chirality are connected by 4,4'-bpy ligands to generate a homochiral layered framework. Edge-to-face aromatic interactions between neighboring layers lead to a 3D homochiral supramolecular structure. Magnetization and heat capacity measurements indicate that the framework is a weak antiferromagnet at low applied field. The magnetic interactions between adjacent manganese ions in the helical chain can be fitted using the 1D Fisher model, with 2J/k = -0.68 K and g = 2.00. Moreover, the compound displays a unique field-dependent spin-flop transition in high magnetic fields, with a critical field of 23.6 kOe at 1.9 K.  相似文献   

17.
1,4‐Diazabicyclo[2.2.2]octane (dabco) and its derivatives have been extensively utilized as building units of excellent molecular ferroelectrics for decades. However, the homochiral dabco‐based ferroelectric remains a blank. Herein, by adding a methyl (Me) group accompanied by the introduction of homochirality to the [H2dabco]2+ in the non‐ferroelectric [H2dabco][TFSA]2 (TFSA=bis(trifluoromethylsulfonyl)ammonium), we successfully designed enantiomeric ferroelectrics [R and S‐2‐Me‐H2dabco][TFSA]2. The two enantiomers show two sequential phase transitions with transition temperature (Tc) as high as 405.8 K and 415.8 K, which is outstanding in both dabco‐based ferroelectrics and homochiral ferroelectrics. To our knowledge, [R and S‐2‐Me‐H2dabco][TFSA]2 are the first examples of dabco‐based homochiral ferroelectrics. This finding opens an avenue to construct dabco‐based homochiral ferroelectrics and will inspire the exploration of more eminent enantiomeric molecular ferroelectrics.  相似文献   

18.
Homochiral metal–organic framework (MOF) membranes have been recently reported for chiral separations. However, only a few high‐quality homochiral polycrystalline MOF membranes have been fabricated due to the difficulty in crystallization of a chiral MOF layer without defects on porous substrates. Alternatively, mixed matrix membranes (MMMs), which combine potential advantages of MOFs and polymers, have been widely demonstrated for gas separation and water purification. Here we report novel homochiral MOF–polymer MMMs for efficient chiral separation. Homochirality was successfully incorporated into achiral MIL‐53‐NH2 nanocrystals by post‐synthetic modification with amino acids, such as l ‐histidine (l ‐His) and l ‐glutamic acid (l ‐Glu). The MIL‐53‐NH‐l ‐His and MIL‐53‐NH‐l ‐Glu nanocrystals were then embedded into polyethersulfone (PES) matrix to form homochiral MMMs, which exhibited excellent enantioselectivity for racemic 1‐phenylethanol with the highest enantiomeric excess value up to 100 %. This work, as an example, demonstrates the feasibility of fabricating diverse large‐scale homochiral MOF‐based MMMs for chiral separation.  相似文献   

19.
在水热条件下, 分别用过渡金属离子Zn(Ⅱ)和Co(Ⅱ)与混合手性羧酸和含氮配体反应, 合成了2个手性金属-有机配位聚合物[M(D-cam)(H2mbdpz)]n[M=Zn(1), Co(2); D-H2cam=D-(+)-樟脑酸; H2mbdpz=4,4'-亚甲基二(3,5-二甲基吡唑)]. 通过单晶X射线衍射、粉末X射线衍射、元素分析、热重分析及荧光分析对其结构、组成和性质进行了表征. 单晶结构分析表明, 配合物1和2是异质同晶的手性三维开放骨架结构, 具有单节点三连接的ThSi2型网络拓扑结构.  相似文献   

20.
The polymer network: The reaction of quinine (QN) with CuI under solvothermal, as well as liquid-assisted grinding, conditions afforded a unique 1D homochiral coordination polymer {[Cu(4)(μ(3)-I)(4)(QN)(2)][Cu(3)(μ(3)-I)(2)(μ(2)-I)(QN)(2)](2)}(n), containing both triangular Cu(3)I(3) and cubane Cu(4)I(4) clusters as connecting nodes (see scheme). Van der Waals interactions between the adjacent 1D polymer chains lead to an extended quasi-honeycomb homochiral pillared 3D network with solvent-free 1D channels.  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