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 15 毫秒
1.
Cucurbit[7]uril (CB[7]), an uncharged and water‐soluble macrocyclic host, binds protonated amino saccharides (D ‐glucosamine, D ‐galactosamine, D ‐mannosamine and 6‐amino‐6‐deoxy‐D ‐glucose) with excellent affinity (Ka=103 to 104 M ?1). The host–guest complexation was confirmed by NMR spectroscopy, isothermal titration calorimetry (ITC), and MALDI‐TOF mass spectral analyses. NMR analyses revealed that the amino saccharides, except D ‐mannosamine, are bound as α‐anomers within the CB[7] cavity. ITC analyses reveal that CB[7] has excellent affinity for binding amino saccharides in water. The maximum affinity was observed for D ‐galactosamine hydrochloride (Ka=1.6×104 M ?1). Such a strong affinity for any saccharide in water using a synthetic receptor is unprecedented, as is the supramolecular stabilization of an α‐anomer by the host.  相似文献   

2.
The kinetics of σ complexation of 2,4,6‐tris(trifluoromethanesulfonyl)anisole ( 7 d ) have been investigated over a large pH range of 2–13.70 at T=20 °C in methanol. Two competitive processes associated with the initial addition of MeO? at the unsubstituted 3‐position of 7 d to give a 1,3‐dimethoxy adduct ( 9 d ‐Me) and a subsequent and slow conversion of this species into a 1,1‐dimethoxy isomer ( 8 d ‐Me) have been identified. Both adducts 8 d ‐Me and 9 d ‐Me are 105–106 times more stable than the related adducts 8 a ‐Me and 9 a ‐Me of 2,4,6‐trinitroanisole ( 7 a ), a conventional reference aromatic electrophile in Meisenheimer complex chemistry. The high stability of 8 d ‐Me and 9 d ‐Me is shown to derive from greater rates of formation and lower rates of decomposition than previously determined for 8 a ‐Me and 9 a ‐Me, thereby emphasising the especially high activation of a benzene ring by SO2CF3 group(s). Analysis of the collected rate and equilibrium data for σ complexation in the anisole series 2,4,6‐tris(SO2CF3)‐, 2,6‐bis(SO2CF3)‐4‐nitro‐, 4‐SO2CF3‐2,6‐dinitro‐ and 2,4,6‐trinitro‐ supports the idea that the especially high capacity of resonance stabilisation of the negative charge of the adducts through an Fπ‐type (as defined in ref. 49 ) polarisation effect is a major factor that accounts for the strong activation provided by SO2CF3 groups. A most significant result is the finding that the 1,1‐dimethoxy adduct 8 d ‐Me is by far the most stable benzene σ adduct so far reported. With a p value of 7.32, this adduct is formed exclusively through methanol addition up to pH≈10. This is consistent with the location of 7 d in the superelectrophilic region defined by p ≤9.5–10.5. For comparison, the solvent contribution is negligible in the formation of the 1,3‐isomer 9 d ‐Me, the p (10.59) of which is situated on the upper limit of the boundary. Taking advantage of the simple relationship linking pKa values for σ complexation in methanol and water, a ranking of the triflone 7 d on the general thermodynamic scale constructed for Meisenheimer electrophiles in water is informative. An approximate calibration on the electrophilicity scale kinetically derived by Mayr et al. has also been made.  相似文献   

3.
Aminocyclopentitol analogs of α‐L ‐fucose were synthesized stereoselectively from D ‐ribose. Alkyl substituents were attached at the NH2 group to mimic the glycosidic leaving group. The resulting (alkylamino)cyclopentitols inhibited α‐L ‐fucosidases selectively with inhibition constants in the range of Ki=10−7 M . Comparisons with stereoisomers and acyclic analogs showed that this inhibition only occurs with N‐alkyl substitution and proper configuration at the cyclopentane, as expected for transition‐state‐analog‐type inhibition. These observations were supported by molecular‐modeling comparisons between inhibitor and transition state.  相似文献   

4.
The recognition of either homomeric or heteromeric pairs of pentoses in an aromatic oligoamide double helical foldamer capsule was evidenced by circular dichroism (CD), NMR spectroscopy, and X‐ray crystallography. The cavity of the host was predicted to be large enough to accommodate simultaneously two xylose molecules and to form a 1:2 complex (one container, two saccharides). Solution and solid‐state data revealed the selective recognition of the α‐4C1‐d ‐xylopyranose tautomer, which is bound at two identical sites in the foldamer cavity. A step further was achieved by sequestering a heteromeric pair of pentoses, that is, one molecule of α‐4C1‐d ‐xylopyranose and one molecule of β‐1C4‐d ‐arabinopyranose despite the symmetrical nature of the host and despite the similarity of the guests. Subtle induced‐fit and allosteric effects are responsible for the outstanding selectivities observed.  相似文献   

5.
Seven discrete sugar‐pendant diamines were complexed to the {M(CO)3}+ (99mTc/Re) core: 1,3‐diamino‐2‐propyl β‐D ‐glucopyranoside ( L 1 ), 1,3‐diamino‐2‐propyl β‐D ‐xylopyranoside ( L 2 ), 1,3‐diamino‐2‐propyl α‐D ‐mannopyranoside ( L 3 ), 1,3‐diamino‐2‐propyl α‐D ‐galactopyranoside ( L 4 ), 1,3‐diamino‐2‐propyl β‐D ‐galactopyranoside ( L 5 ), 1,3‐diamino‐2‐propyl β‐(α‐D ‐glucopyranosyl‐(1,4)‐D ‐glucopyranoside) ( L 6 ), and bis(aminomethyl)bis[(β‐D ‐glucopyranosyloxy)methyl]methane ( L 7 ). The Re complexes [Re( L 1 – L 7 )(Br)(CO)3] were characterized by 1H and 13C 1D/2D NMR spectroscopy which confirmed the pendant nature of the carbohydrate moieties in solution. Additional characterization was provided by IR spectroscopy, elemental analysis, and mass spectrometry. Two analogues, [Re( L 2 )(CO)3Br] and [Re( L 3 )(CO)3Br], were characterized in the solid state by X‐ray crystallography and represent the first reported structures of Re organometallic carbohydrate compounds. Conductivity measurements in H2O established that the complexes exist as [Re( L 1 – L 7 )(H2O)(CO)3]Br in aqueous conditions. Radiolabelling of L 1 – L 7 with [99mTc(H2O)3(CO)3]+ afforded in high yield compounds of identical character to the Re analogues. The radiolabelled compounds were determined to exhibit high in vitro stability towards ligand exchange in the presence of an excess of either cysteine or histidine over a 24 h period.  相似文献   

6.
Three novel chiral Schiff Base ligands (H2L) were prepared from the condensation reaction of 3‐formyl acetylacetone with the amino acids L ‐alanine, L ‐phenylalanine, and L ‐threonine. X‐ray single crystal analyses revealed that the Schiff Base compounds exist as enamine tautomers in the solid state. The molecular structure of the compounds is stabilized by an intramolecular hydrogen bridge between the enamine NH function and a carbonyl oxygen atom of the pentandione residue. Treatment of the ligands H2L with copper(II) actetate in the presence of pyridine led to the formation of copper complexes [CuL(py)]. In each of the complexes the copper atoms adopt a distorted square‐pyramidal coordination. Three of the basal coordination sites are occupied by the doubly deprotonated Schiff Bases L2– which act as tridentate chelating O, N, O‐ligands. The remaining coordination sites are occupied by a pyridine ligand at the base and a carboxyl oxygen atom of a neighboring complex at the apical position. The latter coordination is responsible for a catenation of the complexes in the solid state.  相似文献   

7.
The phase‐transition behaviors, crystal structures, and dielectric properties of four kinds of simple 1:1 organic salts of (C12H25NH3+)(benzenesulfonate) and (C12H25NH3+)(pyridine sulfonates) were examined from the viewpoint of intermolecular hydrogen‐bonding interactions and dynamic conformational transformation in molecular assemblies. Crystals of (C12H25NH3+)(benzenesulfonate) and (C12H25NH3+)(3‐pyridinesulfonate) were isostructural and solid–solid and solid–liquid‐crystal smectic A (SmA) phase transitions were observed. These two crystals formed rodlike cation–anion assemblies. However, the two salts, (C12H25NH3+)(2‐pyridinesulfonate) and (C12H25NH3+)(4‐pyridinesulfonate), formed largely bent L ‐shaped cation–anion conformations. Interesting conformational transformations from rodlike to L ‐shaped assemblies were observed in (C12H25NH3+)(2‐pyridinesulfonate) and (C12H25NH3+)(3‐pyridinesulfonate).  相似文献   

8.
The ability to quantify the local electrostatic environment of proteins and protein/peptide assemblies is key to gaining a microscopic understanding of many biological interactions and processes. Herein, we show that the ester carbonyl stretching vibration of two non‐natural amino acids, L ‐aspartic acid 4‐methyl ester and L ‐glutamic acid 5‐methyl ester, is a convenient and sensitive probe in this regard, since its frequency correlates linearly with the local electrostatic field for both hydrogen‐bonding and non‐hydrogen‐bonding environments. We expect that the resultant frequency–electric‐field map will find use in various applications. Furthermore, we show that, when situated in a non‐hydrogen‐bonding environment, this probe can also be used to measure the local dielectric constant (ε). For example, its application to amyloid fibrils formed by Aβ16–22 revealed that the interior of such β‐sheet assemblies has an ε value of approximately 5.6.  相似文献   

9.
《化学:亚洲杂志》2017,12(17):2207-2210
Tetraphenylethylene (TPE)–carborane hybrids are constructed, and the impact of carborane substituents on the aggregation‐induced emission (AIE) characteristics of TPE‐cores has been investigated. When altering the 2‐R‐group on the carborane unit with ‐H, ‐CH3 or phenyl group, the luminescent quantum yield of the corresponding TPE derivatives can be manipulated from 0.18 to 0.63 in the solid state. The emission color exhibits an obvious 100 nm shift (from blue to yellow).  相似文献   

10.
A convenient method to isolate inverted cucurbit[7]uril (iQ[7]) from a mixture of water‐soluble Q[n]s was established by eluting the soluble mixture of Q[n]s on a Dowex (H+ form) column so that iQ[7] could be selected as a ligand for coordination and supramolecular assembly with alkaline earth cations (AE2+) in aqueous HCl solutions in the presence of [ZnCl4]2? and [CdCl4]2? anions as structure‐directing agents. Single‐crystal X‐ray diffraction analysis revealed that both iQ[7]–AE2+–[ZnCl4]2?–HCl and iQ[7]–AE2+–[CdCl4]2?–HCl interaction systems yielded supramolecular assemblies, in which the [ZnCl4]2? and [CdCl4]2? anions presented a honeycomb effect, and this resulted in the formation of linear iQ[7]/AE2+ coordination polymers through outer‐surface interactions of Q[n]s.  相似文献   

11.
Two types of dendritically functionalized iron(II) porphyrins were prepared (Scheme) and investigated in the presence of 1,2‐dimethylimidazole (1,2‐DiMeIm) as the axial ligand as model systems for T(tense)‐state hemoglobin (Hb) and myoglobin (Mb). Equilibrium O2‐ and CO‐binding studies were performed in toluene and aqueous phosphate buffer (pH 7). UV/VIS Titrations (Fig. 4) revealed that the two dendritic receptors 1 ⋅ Fe II ‐1,2‐DiMeIm and 2 ⋅ Fe II ‐1,2‐DiMeIm (Fig. 2) with secondary amide moieties in the dendritic branching undergo reversible complexation (Fig. 5) with O2 and CO in dry toluene. Whereas the CO affinity is similar to that measured for the natural receptors, the O2 affinity is greatly enhanced and exceeds that of T‐state Hb by a factor of ca. 1500 (Table). The oxygenated complexes possess half‐lives of several h (Fig. 6). This remarkable stability originates from both dendritic encapsulation of the iron(II) porphyrin and formation of a H‐bond between bound O2 and a dendritic amide NH moiety (Fig. 11). Whereas reversible CO binding was also observed in aqueous solution (Fig. 10), the oxygenated iron(II) complexes are destabilized by the presence of H2O with respect to oxidative decay (Fig. 9), possibly as a result of the weakening of the O2⋅⋅⋅H−N H‐bond by the competitive solvent. The comparison between the two dendrimers with amide branchings and ester derivative 3 ⋅ Fe II ‐1,2‐DiMeIm (Fig. 2), which lacks H‐bond donor centers in the periphery of the porphyrin, further supports the role of H‐bonding in stabilizing the O2 complex against irreversible oxidation. All three derivatives bind CO reversibly and with similar affinity (Fig. 8) in dry toluene, but the oxygenated complex of 3 ⋅ Fe II ‐1,2‐DiMeIm undergoes much more rapid oxidative decomposition (Fig. 7).  相似文献   

12.
The first excited‐state proton transfer (ESPT ) in 7AI ‐H2O complex and its derivatives, in which the hydrogen atom at the C2 position in pyrrole ring was replaced by halogen atom X (X = F, Cl, Br), were studied at the TD ‐M06‐2X/6‐31 + G(d, p) level. The double proton transfer took place in a concerted but asynchronous protolysis pathway. The vibrational‐mode selectivity of excited‐state double proton transfer in the model system was confirmed. The specific vibrational‐mode could shorten the reaction path and accelerate the reaction rate. The substituent effects on the excited‐state proton transfer process were discussed. When the H atom at C2 position in 7AI ‐H2O complex was replaced by halogen atom, some geometrical parameters changed obviously, the barrier height of ESDPT reduced, and the asynchronicity of proton transfer enlarged. The above changes correlated with the Pauling electronegativity of halogen atom.  相似文献   

13.
Two modified DNA 14‐mers have been prepared, containing either a 2‐deoxy‐D ‐erythrose‐derived adenosine analogue carrying a C(8)−CH2O group (deA*), or a 2‐deoxy‐D ‐erythrose‐derived uridine analogue, possessing a C(6)−CH2O group (deU*). These nucleosides are linked via a phosphinato group between O−C(3′) (deA* and deU*) and O−C(5′) of one neighbouring nucleotide, and between C(8)−CH2O (deA*), or C(6)−CH2O (deU*) and O−C(3′) of the second neighbour. N6‐Benzoyl‐9‐(β‐D ‐erythrofuranosyl)adenine ( 3 ) and 1‐(β‐D ‐erythrofuranosyl)uracil ( 4 ) were prepared from D ‐glucose, deoxygenated at C(2′), and converted into the required phosphoramidites 1 and 2 . The modified tetradecamers 31 and 32 were prepared by solid‐phase synthesis. Pairing studies show a decrease in the melting temperature of 7 to 8 degrees for the duplexes 31 ⋅ 30 and 32 ⋅ 29 , as compared to the unmodified DNA duplex 29 ⋅ 30 . A comparison with the pairing properties of tetradecamers similarly incorporating deoxyribose‐ instead of the deoxyerythrose‐derived nucleotides evidences that the CH2OH substituent at C(4′) has no significant effect on the pairing.  相似文献   

14.
The impact of the molecular architecture on the transfection efficiency of PEGylated poly(amino acid) block copolymers was investigated for PEG‐b‐p(l ‐Lys)x‐b‐p(l ‐Leu)y, PEG‐b‐p(l ‐Leu)x‐b‐p(l ‐Lys)y, and PEG‐b‐p((l ‐Leu)x‐co‐(l ‐Lys)y). The block lengths of p(l ‐Lys) and p(l ‐Leu) were varied between 10, 20, and 40; and 10 and 20, respectively, to study the influence of the ionic/hydrophobic balance. The results show that ABC triblock copolymers form smaller and more stable polyplexes with plasmid DNA than AB diblock copolymers—as verified by long‐term aggregation and ethidium bromide exclusion studies—protect the DNA more effectively against nucleases, and provide better transfection efficiencies, as indicated by total protein as well as luciferase expression. More detailed studies revealed that triblock copolymers with p(l ‐Leu) forming the C‐block were most efficient in DNA complexation with a 2.3 times higher transfection rate. Furthermore, increasing the cationic character by increasing the p(l ‐Lys) chain length led to up to 25% higher transfection but at the same time induced some cytotoxicity. Diblock copolymers, where the amino acid–building blocks exist as a random copolymer, bind more loosely with DNA leading to less compact and less stable aggregates with lower transfection efficiencies.  相似文献   

15.
Reaction of 2‐isopropyl‐(N,N‐diisopropyl)‐benzamide 5 with tBuLi in ether results in ortho deprotonation and the formation of a hemisolvate based on a tetranuclear dimer of ( 5 ‐Lio)2?Et2O. The solid‐state structure exhibits a dimer core in which the amide oxygen atoms fail to stabilize the metal ions but are instead available for interaction with two metalated monomers that reside peripheral to the core. Reaction of 5 with tBuLi in the presence of the tridentate Lewis base PMDTA (N,N,N′,N′′,N′′‐pentamethyldiethylenetriamine) takes a different course. In spite of the tertiary aliphatic group at the 2‐position in 5 , X‐ray crystallography revealed that a remarkable benzylic (lateral) deprotonation had occurred, giving the tertiary benzyllithium 5 ‐Lil?PMDTA. The solid‐state structure reveals that amide coordination and solvation by PMDTA combine to distance the Li+ ion from the deprotonated α‐C of the 2‐iPr group (3.859(4) Å), thus giving an essentially flat tertiary carbanion and a highly distorted aromatic system. DFT analysis suggests that the metal ion resides closer to the carbanion center in solution. In line with this, the same (benzylic) deprotonation is noted if the reaction is attempted in the presence of tridentate diglyme, with X‐ray crystallography revealing that the metal is now closer to the tertiary carbanion (2.497(4) Å). Electrophilic quenches of lithiated 5 have allowed, for the first time, the formation of quaternary benzylic substituents by lateral lithiation.  相似文献   

16.
The interaction of PEGylated poly(amino acid)s with their biological targets depends on their chemical nature and spatial arrangement of their building blocks. The synthesis, self‐assembly, and DNA complexation of ABC terblock copolymers consisting of poly(ethylene glycol), (PEG), poly(l ‐lysine), and poly(l ‐leucine), are reported. Block copolymers are produced by a metal‐free, living ring‐opening polymerization of respective amino acid N‐carboxyanhydrides using amino‐terminated PEG as macroinitiator: (PEG‐b‐p(l ‐Lys)xb‐p(l ‐Leu)y, PEG‐b‐p(l ‐Leu)xb‐p(l ‐Lys)y, and PEG‐b‐p((l ‐Lys)xco‐p(l ‐Leu)y). Sizes of self‐assembled nanoparticles depend on the formation method. The nanoprecipitation method proves useful for copolymers with the poly(l ‐lysine) block protected as trifluoroacetate, effective diameters range between 92 and 132 nm, while direct dissolution in distilled water is suitable for the deprotected copolymers, yielding effective diameters between 52 and 173 nm. Critical micelle concentration (CMC) analyses corroborate particle size analyses and show a distinct impact of the molecular architecture; the lowest CMC (8 µg mL−1) is observed when the poly(l ‐leucine) segment forms the C‐block and the hydrophilic, disassembly driving poly(l ‐lysine) segment is short. DNA complexation, evaluated by gel motility and RiboGreen analyses, depends strongly on the molecular architecture. A more efficient DNA complexation is observed when poly(l ‐lysine) and poly(l ‐leucine) form individual blocks as opposed to them forming a copolymer.  相似文献   

17.
The inclusion‐complexation behavior of coenzyme Q10 (CoQ10) with the three polyamine‐modified β‐cyclodextrins (CDs) 1 – 3 was investigated in both solution and the solid state by means of NMR, XRD, and FT‐IR spectroscopy. The results showed that the apparent solubility of CoQ10 increased linearly upon addition of hosts 1 – 3 , giving AL‐type phase‐solubility curves. These hosts 1 – 3 were able to solubilize CoQ10 to high levels, up to 1.35, 1.52, and 1.44 mg/ml (calculated as CoQ10), respectively. The host 2 with a moderate‐length chain is the most suitable for inclusion complexation of CoQ10. Accroding to the ROESY experiments, the MeO groups of CoQ10 and the tether of 2 can be co‐included into the cavity of β‐CD through the induced‐fit interaction between host and guest. The binding ability of modified β‐CDs 1 – 3 upon complexation with CoQ10 are discussed from the viewpoints of the size/shape‐matching relationship and the induced‐fit concept between host CDs and guest CoQ10 molecule.  相似文献   

18.
A family of four self‐assembling lipopeptides containing Ala‐Lys peptides attached to a C16 aliphatic chain were synthesised. These compounds form two enantiomeric pairs that bear a diastereomeric relationship to one another (C16‐l ‐Ala‐l ‐Lys/C16‐d ‐Ala‐d ‐Lys) and (C16‐d ‐Ala‐l ‐Lys/C16‐l ‐Ala‐d ‐Lys). These diastereomeric pairs have very different critical micelle concentrations (CMCs). The self‐assembled multivalent (SAMul) systems bind biological polyanions as a result of the cationic lysine groups on their surfaces. For heparin binding, there was no significant enantioselectivity, but there was a binding preference for the diastereomeric assemblies with lower CMCs. Conversely, for DNA binding, there was significant enantioselectivity for systems displaying d ‐lysine ligands, with a further slight preference for attachment to l ‐alanine, with the CMC being irrelevant.  相似文献   

19.
We previously reported that chiral Zn2+ complexes that were designed to mimic the actions of class‐I and class‐II aldolases catalyzed the enantioselective aldol reactions of acetone and its analogues thereof with benzaldehyde derivatives. Herein, we report the synthesis of new chiral Zn2+ complexes that contain Zn2+? tetraazacyclododecane (Zn2+? [12]aneN4) moieties and amino acids that contain aliphatic, aromatic, anionic, cationic, and dipeptide side chains. The chemical and optical yields of the aldol reaction were improved (up to 96 % ee) by using ZnL complexes of L ‐decanylglycyl‐pendant [12]aneN4 (L ‐ZnL7), L ‐naphthylalanyl‐pendant [12]aneN4 (L ‐ZnL10), L ‐biphenylalanyl‐pendant [12]aneN4 (L ‐ZnL11), and L ‐phenylethylglycyl‐pendant [12]aneN4 ligands (L ‐ZnL12). UV/Vis and circular dichroism (CD) titrations of acetylacetone (acac) with ZnL complexes confirmed that a ZnL? (acac)? complex was exclusively formed and not the enaminone of ZnL and acac, as we had previously proposed. Moreover, the results of stopped‐flow experiments indicated that the complexation of (acac)? with ZnL was complete within milliseconds, whereas the formation of an enaminone required several hours. X‐ray crystal‐structure analysis of L ‐ZnL10 and the ZnL complex of L ‐diphenylalanyl‐pendant [12]aneN4 (L ‐ZnL13) shows that the NH2 groups of the amino‐acid side chains of these ligands are coordinated to the Zn2+ center as the fourth coordination site, in addition to three nitrogen atoms of the [12]aneN4 rings. The reaction mechanism of these aldol reactions is discussed and some corrections are made to our previous mechanistic hypothesis.  相似文献   

20.
A controlling influence on the self‐assembly in the complexation reaction of a mixture of methylene‐ and ethylene‐bridged bis(catechol) ligands ( 1 ‐H4 and 2 ‐H4, respectively) with titanium(IV ) ions is exerted by alkali metal cations (see scheme). Thus, not a complicated mixture of complexes, but as a result of a self‐recognition of the ligands only well‐defined products are formed.  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