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 390 毫秒
1.
Quinones can be reduced apparently by water mediated by alloxantin (AT) or alloxan radical(A·) produced photochemically from alloxan monohydrate (A-hydrate).  相似文献   

2.
In this study, using QM/QTAIM calculations in the continuum with ε = 1 under normal conditions, we have revealed for the first time the nondissociative A·T(WC)↔A·T(rWC)/A·T(rH) and A·T(H)↔A·T(rH)/A·T(rWC) conformational transitions. It was established that they proceed via the essentially nonplanar transition states (С1 symmetry) through the intermediates, which are wobbled conformers (С1 symmetry) theoretically predicted in our previous work (Brovarets’ et al., Frontiers in Chemistry, 2018, 6:8, 10.3389/fchem.2018.00008) of the classical А·Т DNA base pairs—Watson–Crick А·Т(WC), reverse Watson–Crick А·Т(rWC), Hoogsteen А·Т(Н) and reverse Hoogsteen А·Т(rН). At this, the A·T(H)↔A·T(rWC) and A·T(WC)↔A·T(rH) conformational transformations are controlled by the transition states (TSs) stabilized by the participation of the intermolecular (T)N3H···N6(A) H‐bond (∼3.70 kcal·mol−1) between the imino group N3H of T and pyramidilized amino group N6H2 of A. Gibbs free energies of activation for these processes consist 12.22 and 11.11 kcal·mol−1, accordingly, under normal conditions. TSs, which control the A·T(WC)↔A·T(rWC) and A·T(H)↔A·T(rH) conformational transitions are stabilized by the participation of the intermolecular (T)N3H···N6(A) H‐bond (5.82 kcal·mol−1) and bifurcating intermolecular (T)N3H···N6(A) (5.00) and (T)N3H···N7(A) (0.61 kcal·mol−1) H‐bonds, accordingly. Notably, in these two TSs amino group N6H2 of A is significantly pyramidilized; Gibbs free energies of activation for these reactions are 19.07 and 19.71 kcal·mol−1, accordingly.  相似文献   

3.
The extraction of thorium(IV) and uranium(VI) from nitric acid solutions has been studied using mixtures of bis(2,4,4-trimethylpentyl)phosphinic acid (Cyanex272 or HA), and synergistic extractants (S) such as tri-butylphosphate (TBP), tri-octylphosphine oxide (TOPO) or bis(2,4,4-trimethylpentyl)thiophosphinic acid (Cyanex301). The results showed that these metallic ions are extracted into kerosene as Th(OH)2(NO3)A·HA and UO2(NO3)A·HA with Cyanex272 alone. In the presence of neutral organophosphorus ligands TBP and TOPO, they are found to be extracted as Th(OH)2(NO3)A·HA·S and UO2(NO3)A·HA·S. On the other hand, Th(IV), U(VI) are extracted as Th(OH)2(NO3)A·HA·2S and UO2(NO3)A·HA·S in the presence of Cyanex301. The addition of neutral extractants such as TOPO and TBP to the extraction system enhanced the extraction efficiency of both elements while Cyanex301 as an acidic extractant has improved the selectivity between uranium and thorium. The effect of TOPO on the extraction was higher than other extractants. The equilibrium constants of above species have been estimated by non-linear regression method. The extraction amounts were determined and the results were compared with those of TBP. Also, it was found that the binding to the neutral ligands by the thorium–Cyanex272 complexes follows the neutral ligand basicity sequence.  相似文献   

4.
设计合成了一系列聚酰亚胺基的共轭骨架材料用于锂电池负极.首先,选用具有不同共轭体系的二酐分子用作共聚物构建单元,随后通过亚胺化反应与三聚氰胺共缩聚.最后,通过进一步热处理提高材料的交联程度和稳定性.将该材料用于锂离子电池负极表现出稳定的电化学性能.聚合物的倍率性能测试结果表明:在150 mA·g~(-1)的电流密度下,循环150次后,放电比容量达到471 mAh·g~(-1)以上,在2 A·g~(-1)的较大电流密度下,放电比容量达122.1 mAh·g~(-1),当电流密度返回至100 mA·g~(-1)时,其放电比容量又上升至532.3 mAh·g~(-1)左右,材料具有较好的倍率性能,聚合物材料在充放电过程中,避免了有机小分子材料在与锂离子结合后,易溶于电解液造成的容量损失.同时,共聚物骨架的共轭结构单元和极性基团,可在保证材料的导电性的同时增加材料结合锂离子的能力,因此表现出了优异的倍率性能.  相似文献   

5.
以L-天冬氨酸为原料,磷酸为催化剂,在不同溶剂中进行缩聚反应,合成中间体聚丁二酰亚胺(PSI),当混合溶剂为m三甲苯/m环丁砜=7/3时,可得到较高分子量的PSI。当催化剂与单体的质量比为0.14时,分子量达到最大值。将PSI碱解得到聚天冬氨酸。  相似文献   

6.
Herein, we first address the question posed in the title by establishing the tautomerization trajectory via the double proton transfer of the adenine·guanine (A·G) DNA base mispair formed by the canonical tautomers of the A and G bases into the A*·G* DNA base mispair, involving mutagenic tautomers, with the use of the quantum‐mechanical calculations and quantum theory of atoms in molecules (QTAIM). It was detected that the A·G ? A*·G* tautomerization proceeds through the asynchronous concerted mechanism. It was revealed that the A·G base mispair is stabilized by the N6H···O6 (5.68) and N1H···N1 (6.51) hydrogen bonds (H‐bonds) and the N2H···HC2 dihydrogen bond (DH‐bond) (0.68 kcal·mol?1), whereas the A*·G* base mispair—by the O6H···N6 (10.88), N1H···N1 (7.01) and C2H···N2 H‐bonds (0.42 kcal·mol?1). The N2H···HC2 DH‐bond smoothly and without bifurcation transforms into the C2H···N2 H‐bond at the IRC = ?10.07 Bohr in the course of the A·G ? A*·G* tautomerization. Using the sweeps of the energies of the intermolecular H‐bonds, it was observed that the N6H···O6 H‐bond is anticooperative to the two others—N1H···N1 and N2H···HC2 in the A·G base mispair, while the latters are significantly cooperative, mutually strengthening each other. In opposite, all three O6H···N6, N1H···N1, and C2H···N2 H‐bonds are cooperative in the A*·G* base mispair. All in all, we established the dynamical instability of the А*·G* base mispair with a short lifetime (4.83·10?14 s), enabling it not to be deemed feasible source of the A* and G* mutagenic tautomers of the DNA bases. The small lifetime of the А*·G* base mispair is predetermined by the negative value of the Gibbs free energy for the A*·G* → A·G transition. Moreover, all of the six low‐frequency intermolecular vibrations cannot develop during this lifetime that additionally confirms the aforementioned results. Thus, the A*·G* base mispair cannot be considered as a source of the mutagenic tautomers of the DNA bases, as the A·G base mispair dissociates during DNA replication exceptionally into the A and G monomers in the canonical tautomeric form. © 2013 Wiley Periodicals, Inc.  相似文献   

7.
The redox cycle between alloxan, a mild oxidizing agent, and its reduction partner, dialuric acid, is investigated using density functional theory. It is found that the initial step is the one‐electron reduction of alloxan followed by protonation, yielding a stable neutral radical, AH·. The radical can then accept another electron to form the dialuric acid anion. The formation of this anion is thermodynamically favored in both the gas phase and in solution. The radical may also undergo dimerization to alloxantin, followed by the transfer of a proton from one moiety to another, yielding alloxan and dialuric acid. This reduction is thermodynamically feasible in the gas phase, but not in aqueous solution. In the case of reduction of alloxan by glutathione at the physiological pH, computed redox potentials indicate that a two‐electron reduction is the favored course of reaction, yielding directly the dialuric acid anion, which then undergoes aerial oxidation to yield the superoxide radical. The redox cycling between alloxan and dialuric acid is responsible for the diabetogenic activity of alloxan, producing cytotoxic radicals on reoxidation of dialuric acid. © 2013 Wiley Periodicals, Inc.  相似文献   

8.
Experimentally observed changes in the electronic absorption spectrum of sodium salt of synthetic polyadenyl–polyuridyl acid (Poly[A]·Poly[U]) in the aqueous solution (pH = 7.0, I = 0.15 mol/L) with the change in temperature were explained in terms of co-existence of two different forms of the Poly[A]·Poly[U] strands. Chemometrics analysis of full set of melting curves measured at 120 wavelengths in the UV spectral range has been performed using evolving factor analysis (EFA) and KALS designed for the calculation of equilibrium constants from spectrometry data without a priori set model of the system. Using the Van’t Hoff equation, we have determined thermodynamic constants of the biopolymer transition from the double stranded helix into the random coil conformation (helix–coil transition), they can be further used to quantify the possibility of these structures to be involved in biochemical processes. Molecular model of partially denatured Poly[A]′·Poly[U]′ complex has been biult by means of molecular mechanics.  相似文献   

9.
A basic host polymer exhibiting pH-regulatable saccharide recognition has been investigated. Poly(m-ethynylpyridine) bearing dialkylamino groups forms helical complexes with saccharides to show induced circular dichroism (ICD). When trifluoroacetic acid was titrated on these complexes, the ICD was gradually enhanced until the amount of the acid reached ca. 0.5 molar equivalence versus the pyridine rings in the polymer, and further addition of the acid suppressed the ICD. The proper addition of the acid also increased the binding constants between the polymer and saccharides. These findings would be due to stabilization of the helical structure consisting of cisoid conformations for each of the adjacent pyridine pairs, which were caused by half-protonation of the pyridine rings. Computational analyses indicated that the pyridinium-pyridine dimeric structure prefers its cisoid conformation to its transoid one.  相似文献   

10.
采用同轴静电纺丝法制备了碳包覆纳米SnO2中空纤维超级电容器电极材料.利用X射线衍射(XRD)、拉曼光谱、扫描电子显微镜(SEM)、透射电子显微镜(TEM)和比表面积分析仪(BET)对材料进行表征.结果表明,纤维呈现中空形貌,平均直径为1 μm; SnO2颗粒均匀分布于碳壳结构中,平均粒径为3-15 nm.材料的比表面积为565 m2·g-1.在三电极体系中,当电流密度为0.25 A·g-1时,电极材料的比容量达397.5 F·g-1;在1.0A·g-1电流密度下,充放电循环3000次后比容量仍保持为初始值的88%.在对称型双电极体系中,电流密度为0.25 A·g-1时,电极材料的比容量达162.0 F·g-1,在1.0 A·g-1电流密度下,充放电循环3000次后比容量仍保持为初始值的84%.  相似文献   

11.
选取溴代噻唑和三乙炔基苯为单体,利用聚合反应自下而上构建含噻唑共轭微孔聚合物(NSCMP),通过热解和KOH活化热解NSCMP制备了氮、硫杂原子硬炭(NSHC)和活化NSHC(KNSHC)。利用扫描电子显微镜、能量色散谱、氮气吸附-脱附和恒流充放电等表征2个样品的结构与电化学性能。研究表明KNSHC中N和S的质量分数分别为10.42%和2.23%,KNSHC比表面积高达2 140 m2·g-1。在0.2 A·g-1电流密度下循环500次后KNSHC和NSHC的可逆比容量分别为946.2和493.7 mAh·g-1。KNSHC的优异电化学性能归因于其独特的孔结构和氮、硫杂原子的协同作用。  相似文献   

12.
《Chemistry & biology》1997,4(8):569-578
Background: Cell-permeable small molecules that target predetermined DNA sequences with high affinity and specificity have the potential to control gene expression. A binary code has been developed to correlate DNA sequence with side-by-side pairings between N-methylpyrrole (Py) and N-methylimidazole (lm) carboxamides in the DNA minor groove. We set out to determine the relative energetics of pairings of Im/Py, Py/Im, Im/Im, and Py/Py for targeting G·C and A·T base pairs. A key specificity issue, which has not been previously addressed, is whether an Im/Im pair is energetically equivalent to an Im/Py pair for targeting G·C base pairs.Results: Equilibrium association constants were determined at two five-base-pair sites for a series of four six-ring hairpin polyamides, in order to test the relative energetics of the four aromatic amino-acid pairings opposite G·C and A·T base pairs in the central position. We observed that a G·C base pair was effectively targeted with Im/Py but not Py/Im, Py/Py, or Im/Im. The A·T base pair was effectively targeted with Py/Py but not Im/Py, Py/Im, or Im/Im.Conclusions: An Im/Im pairing is energetically disfavored for the recognition of both A·T and G·C. This specificity will create important limitations on undesirable slipped motifs that are available for unlinked dimers in the minor groove. Baseline energetic parameters will thus be created which, using the predictability of the current pairing rules for specific molecular recognition of double-helical DNA, will guide further second-generation polyamide design for DNA recognition.  相似文献   

13.
Polybenzidine (PBz) particles were formed by the chemical polymerization in micelles of block copolymer by using Poly (ethylene oxide)-b-Polystyrene [PEO113-b-PSx (x = 50, 58 and 100)] as templates. The samples were characterized by IR, UV-vis absorption spectroscopy, transmission electron microscopy, cyclic voltammetry, constant current charge-discharge and electrochemical impedance to determine their morphologies and electrochemical properties. The results show that the prepared PBz is submicron to nanometer rod-like conducting particles with uniform sizes. Removing the templates did not affect the morphology but slightly reduced the size of the PBz particles. The size and morphology of PBz particles can be tuned by adjusting the amount of monomer. The PBz submicrorods showed 412 F·g?1 specific capacitance in 0.3 mol·L?1HClO4 at the current density of 1 A·g?1, indicating its better electrochemical activity. The specific capacitance of the PBz particles reduced less than 10% after 500 charge-discharge cycles at the current density of 3 A·g?1, indicating its good cycling stability.  相似文献   

14.
Here for the first time we present four novel routes of the tautomerisation via the sequential DPT that links biologically important A·C*(WC) DNA base mispair with Watson–Crick (WC) geometry and wobble (w) A*·C*(w), A·C*O2(w), A*·C*(w 1) and A·C(w ) mismatches, pursuing the goal of estimation of their contribution into the transition mutations during DNA biosynthesis. These processes occur without opening of the pairs and are accompanied by the substantial changes in their geometry. A detailed analysis of these pathways leads to an identification of the A·C*(WC)?A*·C*(w) tautomerisation route as the most suitable among these processes from the point of view of the spontaneous point mutagenesis, since it proceeds via the time that is significantly less than the time used by the replicative DNA-polymerase for the incorporation of one incoming nucleotide into the synthesised DNA double helix. This non-dissociative transition occurs through the planar, highly stable, zwitterionic \( {\text{TS}}_{{{\text{A}} \cdot {\text{C*(WC)}} \leftrightarrow {\text{A}}^{ + } \cdot {\text{C}}^{\text{ - }} (w )}}^{{{\text{A}}^{ + } \cdot {\text{C}}^{\text{ - }} }} \) transition state and dynamically unstable intermediate A+·C?(w) ion pair and is accompanied by the consistent rearrangement of the 10 unique patterns of the specific intermolecular interactions, among which there are from 2 to 4 AH···B H-bonds and 2 loosened A–H–B covalent bridges. Basic physico-chemical properties of this mutual tautomeric transformation, which is internally inherent to the A·C*(WC) and A*·C*(w) base mispairs, are documented, and its possible biological assignment is discussed here.  相似文献   

15.
We have prepared supramolecular polymer gels by mixing solutions of a polystyrene bearing cyclic amidine pendant groups (Poly‐A) and copolymers of acrylic acid and n‐butyl acrylate (Poly‐C), followed by evaporation. FT‐IR analysis indicated that the gels were formed through three‐dimensional network of the amidinium‐carboxylate salt bridge. DSC study showed that the Poly‐A and the Poly‐C were miscible when the salt bridge content was high. On the other hand, the mixtures with small salt bridge content showed phase separation. Dynamic shear measurements showed that the gel prepared from Poly‐A and Poly‐C with acrylic acid unit content of 15% had G' higher than G″ over a temperature range of ?22 °C to 32 °C, in which the G' value reached almost 1 MPa. The gel had a crossover point of G' and G″ at 32 °C, very close to room temperature, which suggested facile processability. © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 765–770  相似文献   

16.
Poly(L ‐lactic acid) (PLLA) was produced by the melt polycondensation of L ‐lactic acid. For the optimization of the reaction conditions, various catalyst systems were examined at different temperature and reaction times. It was discovered that Sn(II) catalysts activated by various proton acids can produce high molecular weight PLLA [weight‐average molecular weight (Mw ) ≥ 100,000] in a relatively short reaction time (≤15 h) compared with simple Sn(II)‐based catalysts (SnO, SnCl2 · 2H2O), which produce PLLA with an Mw of less than 30,000 after 20 h. The new catalyst system is also superior to the conventional systems in regard to racemization and discoloration of the resultant polymer. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 1673–1679, 2000  相似文献   

17.
Poly(dimethylaminoethyl methacrylate N-oxide) (poly(DMAEMNO)) was prepared by oxidation of poly(dimethylaminoethyl methacrylate) with hydrogen peroxide in methanol. From thermogravimetric and IR spectroscopic investigations Cope elimination of amine oxide group in poly(DMAENO) was found to occur at 120–150°C. The postpolymerization of partially pyrolyzed polymer carrying vinyl ester group as pendant was performed with azobisisobutyronitrile at 60°C in methanol to give cross-linked polymer that was found to form hydrogel. Poly(DMAEMNO) gave metal–polymer complexes with CuCl2, ZnCl2, and CoCl2. Cobalt–polymer complex had a constitution of 1:2 of metal ion to amine oxide group, while copper– and zinc–polymer complexes seemed to have structures of 1:1 and 1:2 of metal ion to amine oxide group. Furthermore, polymer complexes of poly(DMAEMNO) with poly(methacrylic acid) and poly(acrylic acid) were found to be formed by mixing aqueous solutions of both polymers and also by radical polymerization of the acid monomers in the presence of poly(DMAEMNO). From elemental analysis, thermogravimetric investigation, and measurement of turbidity it was concluded that the resulting polymer–polymer complexes contained more than one acid monomer unit per one N-oxide unit.  相似文献   

18.
The reduced graphene oxide (RGO)/bisphenol A (BPA) composites were prepared by an adsorption‐reduction method. The composites are characterized by X‐ray diffraction (XRD), UV‐vis, thermogravimetric (TG) analysis, field emission scanning electron microscopy (FESEM), transmission electron microscopy (TEM). The results confirm that BPA is adsorbed on the basal plane of RGO by π‐π stacking interaction. Furthermore, the electrochemical behaviors were evaluated by cyclic voltammetry, galvanostatic charge/discharge techniques and electrochemical impedance spectroscopy (EIS). The results show that the RGO/BPA nanocomposites exhibit ultrahigh specific capacitance of 466 F·g?1 at a current density of 1 A·g?1, excellent rate capability (more than 81% retention at 10 A·g?1 relative to 1 A·g?1) and superior cycling stability (90% capacitance decay after 4000 cycles). Consequently, the RGO/BPA nanocomposites can be regarded as promising electrode materials for supercapacitor applications.  相似文献   

19.

Poly(2‐octadecyl‐butanedioic acid), prepared from polyanhydride PA‐18, possesses novel heavy metal adsorption characteristics. The adsorption capacity of this water insoluble polymer for lead (II) was substantially higher than other heterogeneous adsorbants and is equivalent to those obtained with homogeneous sorbants. The polymer exhibited pseudo‐second‐order kinetics and nearly complete adsorption of lead occurred in 15 min with initial lead (II) concentrations greater than 100 mg · L?1. Adsorptive behavior was accurately predicted by the Dubinin‐Radushkevich isotherm model. The mean free energy of adsorption of lead (II) onto poly(2‐octadecyl‐butanedioic acid) was determined to be 31.6 kJ · mol?1, suggesting an ion exchange component to the adsorption mechanism. Gibb's free energy values for this process indicate that it is spontaneous. Adsorption was relatively independent of pH in the range of 3–5, due to the utilization of the sodium carboxylate form of the chelating groups, and was not influenced by high Na+ concentration and moderate concentrations (up to 200 mg · L?1) of Ca+2. Lead (II) solutions containing 2000 mg · L?1 Ca+2 did reduce the adsorption of 2000 mg · L?1 lead (II) by 28%.  相似文献   

20.
Poly(α-isobutyl-L -aspartate) was prepared by the polycondensation reaction of p-nitrophenyl ester of α-isobutyl-L -aspartate and the conformation of the poly(β-amino acid) was investigated by X-ray diffraction, polarized infrared, circular dichroism (CD), optical rotatory dispersion (ORD), and NMR spectroscopy. α-Isobutyl β-p-nitrophenyl-L -aspartate hydrochloride and hydrobromide were used as monomers and dimethylformamide, chloroform, and chlorobenzene, as solvents. A high-molecular-weight polymer with [η] 1.0 dl/g (dichloroacetic acid, 25°C) was formed in the polymerization of the hydrochloride in chloroform at 25°C. The X-ray diagram and polarized infrared spectrum of the stretched polymer film obtained from a chloroform solution suggested a cross-β-form as the most probable structure in the solid state. The CD spectra of the polymer in a 2,2,2-trifluoroethanol (TFE) solution and its film cast from the solution showed a peak at 205 nm and a trough at 190 nm which were assigned to a β-structure. The polymer was associated in chloroform. The NMR and ORD spectra in chloroform were similar to those in TFE, which suggests that the polymer also exists in the β-structure in chloroform. The addition of small amounts of dichloroacetic acid and sulfuric acid to chloroform and TFE solutions, respectively, destroyed the β-structure. A random copolymer of α-isobutyl-L -aspartate with β-alanine was also prepared by polycondensation reaction. The copolymer apparently did not form an ordered structure in the solid state or in solution.  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