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.
Polybissilsesquioxanes with single‐handed helical morphologies attracted much attention during the last decade, which could be applied as asymmetric catalysts and chiral stationary phases. Herein, a pair of chiral biphenylene‐bridged bissilsesquioxanes were synthesized. They self‐assembled into helical bundles in ethanol, behavior that was confirmed in field emission scanning electron microscopy images. Circular dichroism analysis indicated that the biphenylene groups twisted in a single‐handed fashion. Single‐handed helical polybissilsesquioxane bundles were prepared via polycondensation of the bissilsesquioxanes, using a self‐templating approach. Because of the shrinkage that occurred during polycondensation, the helical pitches of the bundles were shorter than those of their corresponding organic self‐assemblies. The wide‐angle X‐ray diffraction pattern indicated that there were no π–π interactions among the diphenylene groups. The circular dichroism spectra indicated that the chirality was successfully transferred from the bissilsesquioxane self‐assemblies to the polybissilsesquioxane. The polybissilsesquioxanes displayed a capacity for the adsorption of nitrobenzene and had potential application for enantioseparation. Chirality 28:44–48, 2016. © 2015 Wiley Periodicals, Inc.  相似文献   

2.
Although the preparation of single‐handed helical inorganic and hybrid organic‐inorganic nanotubes is well developed, approaches to the formation of single‐handed organopolymeric nanotubes are limited. Here, left‐handed helical m‐phenylenediamine‐formaldehyde resin and 3‐aminophenol‐formaldehyde resin nanotubes were prepared by using N‐palmitoyl glucosamine that can self‐assemble into left‐handed twisted nanoribbons in a mixture of methanol and water. In the reaction mixture, the helical pitch of the nanoribbons decreased with increasing reaction time. The resin nanotubes were obtained after removing the N‐palmitoyl glucosamine template, and circular dichroism spectroscopy indicated that the organopolymeric nanotubes had optical activity. Carbonaceous nanotubes were then prepared by carbonization of the 3‐aminophenol‐formaldehyde resin nanotubes.  相似文献   

3.
4.
Fossil hominids often processed material held between their upper and lower teeth. Pulling with one hand and cutting with the other, they occasionally left impact cut marks on the lip (labial) surface of their incisors and canines. From these actions, it possible to determine the dominant hand used. The frequency of these oblique striations in an array of fossil hominins documents the typically modern pattern of 9 right‐ to 1 left‐hander. This ratio among living Homo sapiens differs from that among chimpanzees and bonobos and more distant primate relatives. Together, all studies of living people affirm that dominant right‐handedness is a uniquely modern human trait. The same pattern extends deep into our past. Thus far, the majority of inferred right‐handed fossils come from Europe, but a single maxilla from a Homo habilis, OH‐65, shows a predominance of right oblique scratches, thus extending right‐handedness into the early Pleistocene of Africa. Other studies show right‐handedness in more recent African, Chinese, and Levantine fossils, but the sample compiled for non‐European fossil specimens remains small. Fossil specimens from Sima del los Huesos and a variety of European Neandertal sites are predominately right‐handed. We argue the 9:1 handedness ratio in Neandertals and the earlier inhabitants of Europe constitutes evidence for a modern pattern of handedness well before the appearance of modern Homo sapiens.  相似文献   

5.
Population hand preferences are rare in nonhuman primates, but individual hand preferences are consistent over a lifetime and considered to reflect an individual's preference to use a particular hemisphere when engaged in a specific task. Previous findings in marmosets have indicated that left‐handed individuals tend to be more fearful than their right‐handed counterparts. Based on these findings, we tested the hypotheses that left‐handed marmosets are (a) more reactive to a social stressor and (b) are slower than right‐handed marmosets in acquiring a reversal learning task. We examined the hand preference of 27 male and female marmosets (ages of 4–7 years old) previously tested in a social separation task and a reversal learning task. Hand preference was determined via a simple reaching task. In the social separation task, monkeys were separated from their partner and the colony for a single 7‐hr session. Urinary cortisol levels and behavior were assessed at baseline, during the separation and 24 hr postseparation. Hand preferences were equally distributed between left (n = 10), right‐handed (n = 10), and ambidextrous (n = 7) individuals. The separation phase was associated with an increase in cortisol levels and behavioral changes that were similar across handedness groups. However, cortisol levels at baseline were positively correlated with right‐handedness, and this relationship was stronger in females than in males. In addition, the occurrence of social behaviors (pre‐ and postseparation) was positively correlated with right‐handedness in both sexes. Baseline cortisol levels did not correlate significantly with social behavior. Acquisition of the reversals was poorer in females than males but did not differ as a function of handedness. We conclude that (a) both stress reactivity and cognitive flexibility are similar across handedness groups and (b) left‐handers exhibit less social behavior and have lower basal cortisol levels than ambidextrous and right‐handed subjects. The underlying causes for these differences remain to be established.  相似文献   

6.
《Chirality》2017,29(3-4):120-129
Novel poly(biphenylylacetylene) derivatives bearing two acetyloxy groups at the 2‐ and 2′‐positions and an alkoxycarbonyl group at the 4′‐position of the biphenyl pendants (poly‐ Ac 's) were synthesized by the polymerization of the corresponding biphenylylacetylenes using a rhodium catalyst. The obtained stereoregular (cis transoidal ) poly‐ Ac 's folded into a predominantly one‐handed helical conformation accompanied by a preferred‐handed axially twisted conformation of the biphenyl pendants through noncovalent interactions with a chiral alcohol and both the induced main‐chain helicity and the pendant axial chirality were maintained, that is, memorized, after complete removal of the chiral alcohol. The stability of the helicity memory of the poly‐ Ac 's in a solution was lower than that of the analogous poly(biphenylylacetylene)s bearing two methoxymethoxy groups at the 2‐ and 2′‐positions of the biphenyl pendants (poly‐ MOM 's). In the solid state, however, the helicity memory of the poly‐ Ac 's was much more stable and showed a better chiral recognition ability toward several racemates than that of the previously reported poly‐ MOM when used as a chiral stationary phase for high‐performance liquid chromatography. In particular, the poly‐ Ac ‐based CSP with a helicity memory efficiently separated racemic benzoin derivatives into enantiomers.  相似文献   

7.
Photolysis of five polychlorinated biphenyl (PCB) congeners [2,4,4′-trichlorobiphenyl (PCB 28), 2,2′,5,′5-tetrachlorobiphenyl (PCB 52), 2,2′,4,5,5′-pentachlorobiphenyl (PCB 101), 2,2′,4,4′,5,′5-hexachlorobiphenyl (PCB 153) and 2,2′,3,4,4′,5,′5-heptachlorobiphenyl (PCB 180)] individually and in combination were carried out in the solvents methanol, ethanol, and 2-propanol. The disappearance of parent congener generally increased with UV intensity. The solvents had significant or limited effect on the removal of PCBs depending on the congener used. Because 2-propanol was highly toxic and methoxylated products were formed when methanol was used, ethanol was selected as the optimum solvent. The results of photolysis of the PCB mixture showed that PCB 52 was formed and accumulated after 4 h of photolysis. The addition of sodium hydroxide increased the rate of photolysis of the PCB mixture. One hundred percent removal can be obtained of the PCB in mixture in 90 min under optimized conditions. Gas chromatography–mass spectrometry was used to determine the intermediates of the photolysis of PCBs under optimized conditions. For the PCB congeners and mixture studied, the major photolytic intermediates were less chlorinated congeners, and biphenyl was the major product with minor amounts of hydroxylated PCBs, ethylated, dimethylated, and methylated biphenyls. Biphenyl could be further degraded by a prolonged photolysis. Toxicity of the PCB mixture during photolysis was monitored by the Microtox® test. It was found that the toxicity increased at the early stage of photolysis, and gradually decreased as the reaction proceeded. After 90 min, the EC50 of the reaction mixture was similar to that of the untreated sample.  相似文献   

8.
Four Ala‐Ala dipeptides with a perfluoroalkyl chain at the N‐terminal were synthesized. They were able to self‐assemble into helical nanofibers and/or twisted nanobelts in a mixture of DMSO/H2O. The handedness of nanofibers and nanobelts was controlled by the chirality of the alanine at the N‐terminal. The stacking handedness of the phenylene groups and the helicity of the perfluoroalkyl chain were studied using circular dichroism spectroscopy and vibrational circular dichroism, respectively. The chirality of the alanine at N‐terminal controlled the stacking handedness of the neighboring phenylene groups. Moreover, due to the low potential barrier between M‐ and P‐helices of the perfluorocarbon chain, the handedness of the organic self‐assemblies eventually controlled the helicity of the perfluorocarbon chain. X‐ray diffraction indicated that a lamellar structure was formed by the dimers of the dipeptides.  相似文献   

9.
One‐handed helical polyphenylacetylenes having achiral amino alcohol moieties, but no chiral side groups, were synthesized by the helix‐sense‐selective copolymerization of an achiral phenylacetylene having an amino alcohol side group with a phenylacetylene having two hydroxyl groups. Since the resulting helical copolymers were successfully utilized as chiral ligands for the enantioselective alkylation of benzaldehyde with diethylzinc, we can conclude that the main‐chain chirality based on the one‐handed helical conformation is useful for the chiral catalysis of an asymmetric reaction for the first time. The enantioselectivities of the reaction were controlled by the optical purities of the helical polymer ligands. In addition, the polymer ligands could be easily recovered by precipitation after the reaction. Chirality 27:454–458, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

10.
Most humans are right‐handed and, like many behavioral traits, there is good evidence that genetic factors play a role in handedness. Many researchers have argued that non‐human animal limb or hand preferences are not under genetic control but instead are determined by random, non‐genetic factors. We used quantitative genetic analyses to estimate the genetic and environmental contributions to three measures of chimpanzee handedness. Results revealed significant population‐level handedness for two of the three measures—the tube task and manual gestures. Furthermore, significant additive genetic effects for the direction and strength of handedness were found for all three measures, with some modulation due to early social rearing experiences. These findings challenge historical and contemporary views of the mechanisms underlying handedness in non‐human animals.  相似文献   

11.
A new one‐dimensional (1D) copper(II) coordination polymer {[Cu2(dmaepox)(dabt)](NO3)·0.5 H2O}n, where H3dmaepox and dabt denote N‐benzoato‐N′‐(3‐methylaminopropyl)oxamide and 2,2′‐diamino‐4,4′‐bithiazole, respectively, was synthesized and characterized by single‐crystal X‐ray diffraction and other methods. The crystal structure analysis revealed that the two copper(II) ions are bridged alternately by cis‐oxamido and carboxylato groups to form a 1‐D coordination polymer with the corresponding Cu···Cu separations of 5.1946(19) and 5.038(2) Å. There is a three‐dimensional supramolecular structure constructed by hydrogen bonding and π–π stacking interactions in the crystal. The reactivity towards herring sperm DNA (HS‐DNA) and bovine serum albumin (BSA) indicated that the copper(II) polymer can interact with the DNA in the mode of intercalation, and bind to BSA responsible for quenching of tryptophan fluorescence by the static quenching mechanism. The in vitro cytotoxicity suggested that the copper(II) polymer exhibits cytotoxic effects against the selected tumor cell lines.  相似文献   

12.
Several diaminodiphenyl analogs were assessed in vivo for their capacity to inhibit seizure induction and propagation in rodents. Both 3,4′- and 4,4′-diaminodiphenyl compounds prevented seizures for as long as 4 h after maximal electric shock induction. 4,4′-Diphenyl compounds bridged by a methylene, sulfide, or carbonyl linker also attenuated focal seizure acquisition in a kindling model. Of these analogs, based upon data generated in two rodent species, 4,4′-thiodianiline (1) was identified as the most active compound, significantly reducing seizure staging scores and after-discharge duration for several hours after systemic administration. All compounds were devoid of acute in vivo neurotoxicity at doses well above those required for anticonvulsant activity.  相似文献   

13.
The adsorption of diastereoisomers of tartaric acid, namely, meso (R,S)‐tartaric acid, (R,R)‐tartaric acid, and the racemic mixture of (R,R) and (S,S) tartaric acid on the (111) surface of a copper single crystal has been studied by means of reflection‐absorption IR Spectroscopy, X‐ray photoelectron spectroscopy, low‐energy electron diffraction, and thermal desorption spectroscopy. Two distinct adsorption modes are identified for all three adsorbate systems. All molecules undergo an identical thermally induced autocatalytic decomposition reaction above 510 K. The pure enantiomers show 2D chiral long‐range ordered structures of opposite handedness.  相似文献   

14.
The aim of this study was to explore differences between left‐and right‐handed subjects in sleep duration. Sleep and activity patterns were continuously registered for 12 days using actometers on 20 left‐handed and 20 right‐handed medical students in Berlin. Handedness was determined by a modified version of the Edinburgh handedness inventory. Each participant wore one actometer on each wrist. Actiwatch® Sleep Analysis Software (CNT, UK) was used to evaluate the data, and statistical calculations were performed with a non‐parametric variance analysis. A significant difference in mean sleep duration between left‐handers (7.9 h) and right‐handers (7.3 h) was determined (p=0.025 for measurement made on the dominant hand and p=0.013 for ones made on the non‐dominant hand). In contrast, the maximal phase of daily activity (acrophase) did not show any difference between the two groups. The difference in sleep duration might be caused by either the greater effort required for left‐handers to cope in a right‐handed world or by structural brain differences.  相似文献   

15.
The use of Boc as a nucleobase‐protecting group in the synthesis of sugar‐modified thymidine analogs is reported. Boc was easily inserted at N(3) by a simple and high‐yielding reaction and found to be stable to standard treatments for the removal of Ac and tBuMe2Si (TBDMS) groups, as well as to ZnBr2‐mediated 4,4′‐dimethoxytrityl (DMTr) deprotection. Boc Protection proved to be completely resistant to the strong basic conditions required to regioselectively achieve O‐alkylation, therefore, providing synthetic access to a variety of sugar‐alkylated nucleoside analogs. To demonstrate the feasibility of this approach, two 3′‐O‐alkylated thymidine analogs have been synthesized in high overall yields and fully characterized.  相似文献   

16.
Despite the large number of publications on three‐helix protein folding, there is no study devoted to the influence of handedness on the rate of three‐helix protein folding. From the experimental studies, we make a conclusion that the left‐handed three‐helix proteins fold faster than the right‐handed ones. What may explain this difference? An important question arising in this paper is whether the modeling of protein folding can catch the difference between the protein folding rates of proteins with similar structures but with different folding mechanisms. To answer this question, the folding of eight three‐helix proteins (four right‐handed and four left‐handed), which are similar in size, was modeled using the Monte Carlo and dynamic programming methods. The studies allowed us to determine the orders of folding of the secondary‐structure elements in these domains and amino acid residues which are important for the folding. The obtained data are in good correlation with each other and with the experimental data. Structural analysis of these proteins demonstrated that the left‐handed domains have a lesser number of contacts per residue and a smaller radius of cross section than the right‐handed domains. This may be one of the explanations of the observed fact. The same tendency is observed for the large dataset consisting of 332 three‐helix proteins (238 right‐ and 94 left‐handed). From our analysis, we found that the left‐handed three‐helix proteins have some less‐dense packing that should result in faster folding for some proteins as compared to the case of right‐handed proteins.Proteins 2013; © 2013 Wiley Periodicals, Inc.  相似文献   

17.
In this article, we review the relevant results obtained during almost 60 years of research on a specific aspect of stereochemistry, namely handedness preference and switches between right‐handed and left‐handed helical peptide structures generated by protein amino acids or appropriately designed, side‐chain modified analogs. In particular, we present and discuss here experimental and theoretical data on three categories of those screw‐sense issues: (i) right‐handed/left‐handed α‐helix transitions underwent by peptides rich in Asp, specific Asp β‐esters, and Asn; (ii) comparison of the preferred conformations adopted by helical host–guest peptide series, each characterized by an amino acid residue (e.g. Ile or its diastereomer aIle) endowed with two chiral centers in its chemical structure; and (iii) right‐handed (type I)/left‐handed (type II) poly‐(Pro)n helix transitions monitored for peptides rich in Pro itself or its analogs with a pyrrolidine ring substitution, particularly at the biologically important position 4. The unique modular and chiral properties of peptides, combined with their relatively easy synthesis, the chance to shape them into the desired conformation, and the enormous chemical diversity of their coded and non‐coded α‐amino acid building blocks, offer a huge opportunity to structural chemists for applications to bioscience and nanoscience problems. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
The synthetic peptide Z‐Gly‐Aib‐Gly‐Aib‐OtBu was dissolved in methanol and crystallized in a mixture of ethyl acetate and petroleum ether. The crystals belong to the centrosymmetric space group P4/n that is observed less than 0.3% in the Cambridge Structural Database. The first Gly residue assumes a semi‐extended conformation (φ ±62°, ψ ?131°). The right‐handed peptide folds in two consecutive β‐turns of type II' and type I or an incipient 310‐helix, and the left‐handed counterpart folds accordingly in the opposite configuration. In the crystal lattice, one molecule is linked to four neighbors in the ab‐plane via hydrogen bonds. These bonds form a continuous network of left‐ and right‐handed molecules. The successive ab‐planes stack via apolar contacts in the c‐direction. An ethyl acetate molecule is situated on and close to the fourfold axis. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
A simple and effective method of the methylation on the 2′-O position of adenosine is described. Adenosine is treated with CH3I in an anhydrous alkaline medium at 0°C for 4 h. The major products of this reaction are monomethylated adenosine at either the 2′-O or 3′-O position (total of 64%) and the side products are dimethylated adenosine (2′,3′-O-dimethyladenosi, 21%, and N6-2′-O-dimethyladenosine, 11%). The ratio of 2′-O- and 3′-O-methyladenosine has been found to be 8 to 1. Therefore, this reaction preferentially favors the synthesis of 2′-O-methyladenosine. The monomethylated adenosine is isolated from reaction mixture by a silica gel column chromatography. Then the pure 2′-O-methyladenosine can be separated by crystallization in ethanol from the mixture of 2′-O and 3′-O-methylated isomers. The overall yield of 2′-O-methyladenosine is 42%.  相似文献   

20.
The remarkable predominance of right‐handedness in beta‐alpha‐beta helical crossovers has been previously explained in terms of thermodynamic stability and kinetic accessibility, but a different kinetic trapping mechanism may also play a role. If the beta‐sheet contacts are made before the crossover helix is fully formed, and if the backbone angles of the folding helix follows the energetic pathway of least resistance, then the helix would impart a torque on the ends of the two strands. Such a torque would tear apart a left‐handed conformation but hold together a right‐handed one. Right‐handed helical crossovers predominate even in all‐alpha proteins, where previous explanations based on the preferred twist of the beta sheet do not apply. Using simple molecular simulations, we can reproduce the right‐handed preference in beta‐alpha‐beta units, without imposing specific beta‐strand geometry. The new kinetic trapping mechanism is dubbed the “phone cord effect” because it is reminiscent of the way a helical phone cord forms superhelices to relieve torsional stress. Kinetic trapping explains the presence of a right‐handed superhelical preference in alpha helical crossovers and provides a possible folding mechanism for knotted proteins.  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