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 31 毫秒
1.
Solid compounds of Cd(II), Hg(II) and Pb(II) with the sodium salt of morin-5′-sulfonic acid (NaMSA) were obtained. The molecular formula of the complexes are: Cd(C15H8O10SNa)2?·?6H2O, CdOH(C15H8O10SNa)?·?4H2O, Hg(C15H8O10S)?·?4H2O and Pb(C15H8O10S)?·?3H2O. Some of their physicochemical properties such as UV-Vis, infrared, 13C NMR and mass spectra, thermogravimetric analysis, and solubility were studied. On the basis of spectroscopic data NaMSA was bound to Cd2+ via 4C=O and 3C?–?oxygen and the Hg2+ and Pb2+ ions by 5C–OH, 4C=O and 3C–OH.  相似文献   

2.
Experimental data obtained from the study of the behaviour of monoalkyl phosphates ROP(O)(OH)O? or monoalkyl phosphonates C6H5P(O)(OR)O? in aquo-alcoholic (H2O + R′OH) solutions suggest: (1) in the case of monoalkyl phosphonates the exchange of R and R′ must proceed by a direct alcoholysis; (2) in the case of monoalkyl phosphates, where the scission of the ester at about pH 4,5 proceeds via the intermediate formation of metaphosphate ion PO3?, R′OPO3H? can also be formed by the addition of R′OH to PO3? (competitive to the addition of water) besides direct alcoholysis of ROPO3H?.  相似文献   

3.
Rate constants for the reactions of OH radicals and Cl atoms with 1‐propanol (1‐C3H7OH) have been determined over the temperature range 273–343 K by the use of a relative rate technique. The value of k(Cl + 1‐C3H7OH) = (1.69 ± 0.19) × 10?12 cm3 molecule?1 s?1 at 298 K and shows a small increase of 10% between 273 and 342 K. The value of k(OH + 1‐C3H7OH) increases by 14% between 273 and 343 K with a value of (5.50 ± 0.55) × 10?12 cm3 molecule?1 s?1 at 298 K, and further when combined with a single independent experimentally determined value at 753 K gives k(OH + 1‐C3H7OH) = 4.69 × 10?17T1.8 exp(422/T) cm3 molecule?1 s?1, which fits each data point to better than 2%. Two well‐established structure–activity relationships for H abstraction by OH radicals give accurate predictions of the rate constant for OH + 1‐C3H7OH, provided the β‐CH2 group is given an increased reactivity of a factor of about 2 over that for the structurally equivalent CH2 group in alkanes at 298 K. A quantitative product analysis was carried out at 298 K for the Cl‐initiated photooxidation of 1‐C3H7OH, using both FTIR and gas chromatography. HCHO, CH3CHO, and C2H5CHO were the only major organic primary products observed, although HCOOH was found in much smaller amounts as a secondary product. A key characteristic of the analysis was that the initial values of the product ratio [CH3CHO]/[C2H5CHO] were effectively constant for NO pressures between 0.15 and 0.3 Torr, but fell by about 35% as the pressure fell to 0.0375 Torr. From a detailed consideration of the mechanism for the oxidation, it is suggested that C2H5CHO, CH3CHO (+HCHO), and 3 molecules of HCHO are formed uniquely from CH3CH2CHOH, CH3CHCH2OH, and CH2CH2CH2OH radicals, respectively. On this basis, use of the product yields gives the branching ratios of 56, 30, and 14% for Cl atom reaction at the α‐, β‐, and γ‐C? H positions in 1‐C3H7OH at 298 K. Given the very low temperature coefficients involved, little change will occur over tropospheric temperature ranges. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 34: 110–121, 2002  相似文献   

4.
The kinetics of oxidation of ethanol by cerium(IV) in presence of ruthenium(III) (in the order of 10?7 mol dm?3) in aqueous sulfuric acid media have been followed at different temperatures (25–40°C). The rate of disappearance of cerium(IV) in the title reaction increases sharply with increasing [C2H5OH] to a value independent of [C2H5OH] over a large range (0.2–1.0 mol dm?3) in which the rate law conforms to: where [Ru]T gives the total ruthenium (III) concentration. The values of 10?3kc and 10?3kd are 3.6 ± 0.1 dm3 mol?1 s?1 and 3.9 ± 0.2 s?1, respectively, at 40°C, I = 3.0 mol dm?3. The proposed mechanism involves the formation of ruthenium(III)? substrate complex which undergoes oxidation at the rate determining step by cerium(IV) to form ruthenium(IV)? substrate complex followed by the rapid red-ox decomposition giving rise to the catalyst and ethoxide radical which is oxidized by cerium(IV) rapidly. The mechanism is consistent with the existence of the complexes RuIII · (C2H5OH) and RuIII · (C2H5O?) and both are kinetically active. The overall bisulphate dependence conforms to: kobsd = A[Ru]T/{1 + C[HSO4?]} where A = 2.2 × 104 dm3 mol?1 s?1, C = 1.3 at 40°C, [H+] = 0.5 mol dm?3, and I = 3.0 mol dm?3. The observations are consistent with the Ce(SO4)2 as the kinetically active species. © 1995 John Wiley & Sons, Inc.  相似文献   

5.
IntroductionThereisatremendousactivityforthepastyearsintheareaofinorganic organichybridmaterialsinviewoftheirdi versifiedstructuresandinterestingproperties.1,2 Theeffortshavebeenmadeinsynthesizingandcharacterizingtheclassofmaterials.3Thecontrolofinorganicstructurebyanorganiccomponentrevealsaninteractivestructuralrelationinthema terials .4 Assemblyofaninorganic organichybridmaterialcanbeachievedbyselectingorganiccomponents (multidentateligands)andinorganiccomponents (transitionmetalions ,metalo…  相似文献   

6.
A flash photolysis–resonance fluorescence technique was used to investigate the kinetics of the OH(X2Π) radical and O(3P) atom‐initiated reactions with CHI3 and the kinetics of the O(3P) atom‐initiated reaction with C2H5I. The reactions of the O(3P) atom with CHI3 and C2H5I were studied over the temperature range of 296 to 373 K in 14 Torr of helium, and the reaction of the OH (X2Π) radical with CHI3 was studied at T = 298 K in 186 Torr of helium. The experiments involved time‐resolved resonance fluorescence detection of OH (A2Σ+ → X2Π transition at λ = 308 nm) and of O(3P) (λ = 130.2, 130.5, and 130.6 nm) following flash photolysis of the H2O/He, H2O/CHI3/He, O3/He, and O3/C2H5I/He mixtures. A xenon vacuum UV (VUV) flash lamp (λ > 120 nm) served as a photolysis light source. The OH radicals were produced by the VUV flash photolysis of water, and the O(3P) atoms were produced by the VUV flash photolysis of ozone. Decays of OH radicals and O(3P) atoms in the presence of CHI3 and C2H5I were observed to be exponential, and the decay rates were found to be linearly dependent on the CHI3 and C2H5I concentrations. Measured rate coefficients for the reaction of O(3P) atoms with CHI3 and C2H5I are described by the following Arrhenius expressions (units are cm3 s?1): kO+C2H5I(T) = (17.2 ± 7.4) × 10?12 exp[?(190 ± 140)K/T] and kO+CHI3(T) = (1.80 ± 2.70) × 10?12 exp[?(440 ± 500)K/T]; the 298 K rate coefficient for the reaction of the OH radical with CHI3 is kOH+CHI3(298 K) = (1.65 ± 0.06) × 10?11 cm3 s?1. The listed uncertainty values of the Arrhenius parameters are 2σ‐standard errors of the calculated slopes by linear regression.  相似文献   

7.
Crystals of hypoxanthinium (6‐oxo‐1H,7H‐purin‐9‐ium) nitrate hydrates were investigated by means of X‐ray diffraction at different temperatures. The data for hypoxanthinium nitrate monohydrate (C5H5N4O+·NO3?·H2O, Hx1 ) were collected at 20, 105 and 285 K. The room‐temperature phase was reported previously [Schmalle et al. (1990). Acta Cryst. C 46 , 340–342] and the low‐temperature phase has not been investigated yet. The structure underwent a phase transition, which resulted in a change of space group from Pmnb to P21/n at lower temperature and subsequently in nonmerohedral twinning. The structure of hypoxanthinium dinitrate trihydrate (H3O+·C5H5N4O+·2NO3?·2H2O, Hx2 ) was determined at 20 and 100 K, and also has not been reported previously. The Hx2 structure consists of two types of layers: the `hypoxanthinium nitrate monohydrate' layers (HX) observed in Hx1 and layers of Zundel complex H3O+·H2O interacting with nitrate anions (OX). The crystal can be considered as a solid solution of two salts, i.e. hypoxanthinium nitrate monohydrate, C5H5N4O+·NO3?·H2O, and oxonium nitrate monohydrate, H3O+(H2O)·NO3?.  相似文献   

8.
The kinetics of C2H5O2 and C2H5O2 radicals with NO have been studied at 298 K using the discharge flow technique coupled to laser induced fluorescence (LIF) and mass spectrometry analysis. The temporal profiles of C2H5O were monitored by LIF. The rate constant for C2H5O + NO → Products (2), measured in the presence of helium, has been found to be pressure dependent: k2 = (1.25±0.04) × 10?11, (1.66±0.06) × 10?11, (1.81±0.06) × 10?11 at P (He) = 0.55, 1 and 2 torr, respectively (units are cm3 molecule?1 s?1). The Lindemann-Hinshelwood analysis of these rate constant data and previous high pressure measurements indicates competition between association and disproportionation channels: C2H5O + NO + M → C2H5ONO + M (2a), C2H5O + NO → CH3CHO + HNO (2b). The following calculated average values were obtained for the low and high pressure limits of k2a and for k2b : k = (2.6±1.0) × 10?28 cm6 molecule?2 s?1, k = (3.1±0.8) × 10?11 cm3 molecule?1 s?1 and k2b ca. 8 × 10?12 cm3 molecule?1 s?1. The present value of k, obtained with He as the third body, is significantly lower than the value (2.0±1.0) × 10?27 cm6 molecule?2 s?1 recommended in air. The rate constant for the reaction C2H5O2 + NO → C2H5O + NO2 (3) has been measured at 1 torr of He from the simulation of experimental C2H5O profiles. The value obtained for k3 = (8.2±1.6) × 10?12 cm3 molecule?1 s?1 is in good agreement with previous studies using complementary methods. © 1995 John Wiley & Sons, Inc.  相似文献   

9.
In two linkage isomers, bis[1,3‐di­methyl‐2,4,6(1H,3H,5H)‐pyrimidine­trionato]‐C5,O4‐(ethyl­enedi­amine‐N,N′)platinum(II), [Pt(C6H7N2O3)2(C2H8N2)], (I), and bis[1,3‐di­methyl‐2,4,6(1H,3H,5H)‐py­rim­idine­tri­on­ato‐C5](ethyl­enediamine‐N,N′)­plati­num(II) di­hyd­rate, [Pt(C6H7N2O3)2(C2H8N2)]·2H2O, (II), crystal­lized from the same aqueous solution containing [Pt(en)(OH)2] and 1,3‐di­methyl­barbituric acid (Hdmbarb) in a 1:2 molar ratio, a pair of monodentate dmbarb? anions coordinate to the Pt atom at tetrahedral C atoms for (II), while one dmbarb? anion coordinates at the carbon and the other at a deprotonated enol oxy­gen for (I). The Pt—C distances in (I) and (II) are comparable: 2.112 (4) Å for (I), and 2.114 (4) and 2.117 (4) Å for (II).  相似文献   

10.
The triazenols 4-R1? C6H4? N?N? N(OH)? R2 ( 1 ), oxidized with t-BuO radicals, produced nitroxide radicals R1? C6H4? N(O?)? N?N(R2) +O? ( 5 ). The suggested radical structure was confirmed by 15N-labeling. The reaction of triazenols 1 with PbO2 proceeded under N2 elimination, in which case nitroxides R1? C6H4? N(R2)? O?( 2 ) were observed as the final radical products. The intermediate R1? C6H radicals were identified by spin-trapping.  相似文献   

11.
Loss of an alkyl group X? from acetylenic alcohols HC?C? CX(OH)(CH3) and gas phase protonation of HC?C? CO? CH3 are both shown to yield stable HC?C? \documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm C}\limits^{\rm + } $\end{document}(OH)(CH3) ions. Ions of this structure are unique among all other [C4H5O]+ isomers by having m/z 43 [C2H3O]+ as base peak in both the metastable ion and collisional activation spectra. It is concluded that the composite metastable peak for formation of m/z 43 corresponds to two distinct reaction profiles which lead to the same product ion, CH3\documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm C}\limits^{\rm + } $\end{document}?O, and neutral, HC?CH. It is further shown that the [C4H5O]+ ions from related alcohols (like HC?C? CH(OH)(CH3)) which have an α-H atom available for isomerization into energy rich allenyl type molecular ions, consist of a second stable structure, H2C?\documentclass{article}\pagestyle{empty}\begin{document}$ \mathop {\rm C}\limits^{\rm + } $\end{document}? C(OH)?CH2.  相似文献   

12.
《Electroanalysis》2004,16(19):1622-1627
The pH‐dependence of the stationary open‐circuit potential Ei=0st of rhodium electrode with a surface layer of anodically formed insoluble compounds has been studied in sulfate and phosphate solutions by means of cyclic voltammetry and chronopotentiometry. The range of potentials of the investigations performed has been confined to the region of rhodium electrochemical oxidation/reduction, i.e., 0.2<E<1.2 V (RHE) in order to prevent any possible interference of other reactions such as H2 and O2 evolution. It has been shown that rhodium electrode with a layer of surface compounds formed anodically at E<<1.23 V (RHE) behaves like a reversible metal‐oxide electrode within the range of pH values from ca. 1.0 to ca. 8.0. It has been presumed that the stationary potential of such electrode is determined by the equilibrium of the following electrochemical reaction: Rh+3H2O??Rh(OH)3+3H++3e?. The pH‐dependence of the reversible potential of Eequation/tex2gif-inf-6.gif electrode has been found to be: Eequation/tex2gif-inf-8.gif=Ei=0st=0.69?0.059 pH, V. In acid solutions (pH<2.0) rhodium hydroxide dissolves into the electrolyte, therefore, to reach equilibrium, the solution must be saturated with Rh(OH)3. This has been achieved by adding Rh3+ ions in the form of Rh2(SO4)3. The solubility product of Rh(OH)3, estimated from the experimental Eequation/tex2gif-inf-16.gif?pH dependence obtained, is ca. 1.0×10?48, which is close to the value given in literature.  相似文献   

13.
The aerobic decarboxylation of saturated carboxylic acids (from C2 to C5) in water by TiO2 photocatalysis was systematically investigated in this work. It was found that the split of C1? C2 bond of the acids to release CO2 proceeds sequentially (that is, a C5 acid sequentially forms C4 products, then C3 and so forth). As a model reaction, the decarboxylation of propionic acid to produce acetic acid was tracked by using isotopic‐labeled H218O. As much as ≈42 % of oxygen atoms of the produced acetic acids were from dioxygen (16O2). Through diffuse reflectance FTIR measurements (DRIFTS), we confirmed that an intermediate pyruvic acid was generated prior to the cut‐off of the initial carboxyl group; this intermediate was evidenced by the appearance of an absorption peak at 1772 cm?1 (attributed to C?O stretch of α‐keto group of pyruvic acid) and the shift of this peak to 1726 cm?1 when H216O was replaced by H218O. Consequently, pyruvic acid was chosen as another model molecule to observe how its decarboxylation occurs in H216O under an atmosphere of 18O2. With the α‐keto oxygen of pyruvic acid preserved in the carboxyl group of acetic acid, ≈24 % new oxygen atoms of the produced acetic acid were from molecular oxygen at near 100 % conversion of pyruvic acid. The other ≈76 % oxygen atoms were provided by H2O through hole/OH radical oxidation. In the presence of conduction band electrons, O2 can independently accomplish such C1? C2 bond cleavage of pyruvic acid to generate acetic acid with ≈100 % selectivity, as confirmed by an electrochemical experiment carried out in the dark. More importantly, the ratio of O2 participation in decarboxylation increased along with the increase of pyruvic acid conversion, indicating the differences between non‐substituted acids and α‐keto acids. This also suggests that the O2‐dependent decarboxylation competes with hole/OH‐radical‐promoted decarboxylation and depends on TiO2 surface defects at which Ti4c sites are available for the simultaneous coordination of substrates and O2.  相似文献   

14.
The hydrothermal reaction of Cu(CH3COO)2·H2O, H3BO3, ethylenediamine and H2O in a molar ratio of 3:20:9:222 at 140°C for 5 d yields the deep blue crystals of a new copper polyborate [Cu(en)2B(OH)3]· [B5O5(OH)7] (en?H2NCH2CH2NH2) in 70% yield. It crystallizes in monoclinic system, space group P21/c, with unit cell dimensions, a=1.2779(2) nm, b=1.0167(15) nm, c=1.5019(2) nm, β=90.30(2)°, Z=4. The crystal structure of this compound consists of [Cu(en)2B(OH)3]2+ cation and [B5O5(OH)7]2? anion, which are linked together through hydrogen bonding interactions and electrostatic forces, forming an interesting three‐dimensional framework. The [B5O5(OH)7]2? anion is constituted of [B4O5(OH)4]2? anion and discrete B(OH)3 group which attaches to the side of [B4O5(OH)4]2? through intramolecular hydrogen bonds. Fundamental vibrational modes of this compound were identified and band assignments were made. The middle bands observed at 882 and 575 cm?1 in Raman spectrum are the characteristic peak of B(OH)3 group and [B4O5(OH)4]2? anion, respectively. Additionally the thermal behavior of title compound was recorded and its decomposition mechanism was discussed.  相似文献   

15.
Peroxynitrates (RO2NO2), in particular acyl peroxynitrates (R = R′C(O) with R′ = alkyl), are prominent constituents of polluted air. In this work, a systematic study on the thermal decomposition rate constants of the first five members of the series of homologous R′C(O)O2NO2 with R′ = CH3 ( =PAN), C2H5, n‐C3H7, n‐C4H9, and n‐C5H11 is undertaken to verify the conclusions from previous laboratory data (Grosjean et al., Environ. Sci. Technol. 1994, 28, 1099–1105; Grosjean et al., Environ. Sci. Technol. 1996, 30, 1038–1047; Bossmeyer et al., Geophys. Res. Lett. 2006, 33, L18810) that the longer chain peroxynitrates may be considerably more stable than PAN. Experiments are performed in a temperature‐controlled, evacuable 200 L‐photoreactor made from quartz. n‐Acyl peroxynitrates are generated by stationary photolysis of mixtures of molecular bromine, O2, NO2, and the corresponding parent aldehydes, highly diluted in N2. Thermal decomposition of R′C(O)O2NO2 is initiated by the addition of an excess of NO. First‐order decomposition rate constants k1 of the reactions R′C(O)O2NO2 (+M) → R′C(O)O2 + NO2 (+M) are derived at 298 K and a total pressure of 1 bar from the measured loss rates of R′C(O)O2NO2, correcting for wall loss of R′C(O)O2NO2 and several percentages of reformation of R′C(O)O2NO2 by the reaction of R′C(O)O2 radicals with NO2. With increasing chain length of R′, k1(298 K) slightly decreases from 4.4 × 10?4 s?1 (R′ = CH3) to 3.7 × 10?4 s?1 (R′ = C2H5), leveling off at (3.4 ± 0.1) × 10?4 s?1 for R′ = n‐C3H7, n‐C4H9, and n‐C5H11. Temperature dependencies of k1 were measured for CH3C(O)O2NO2 and n‐C5H11C(O)O2NO2 in the temperature range 289–308 K, resulting in the same activation energy within the statistical error limits (2σ) of 0.9 and 1.5 kJ mol?1, respectively. A few experiments on n‐C6H13C(O)O2NO2, n‐C7H15C(O)O2NO2, and n‐C8H17C(O)O2NO2 were also performed, but the results were considered to be unreliable due to strong wall loss of the peroxynitrate and possible complications caused by radical‐sinitiated side reactions.  相似文献   

16.
The [C4H6O] ion of structure [CH2?CHCH?CHOH] (a) is generated by loss of C4H8 from ionized 6,6-dimethyl-2-cyclohexen-1-ol. The heat of formation ΔHf of [CH2?CHCH?CHOH] was estimated to be 736 kJ mol?1. The isomeric ion [CH2?C(OH)CH?CH2] (b) was shown to have ΔHf, ? 761 kJ mol?1, 54 kJ mol?1 less than that of its keto analogue [CH3COCH?CH2]. Ion [CH2?C(OH)CH?CH2] may be generated by loss of C2H4 from ionized hex-1-en-3-one or by loss of C4H8 from ionized 4,4-dimethyl-2-cyclohexen-1-ol. The [C4H6O] ion generated by loss of C2H4 from ionized 2-cyclohexen-1-ol was shown to consist of a mixture of the above enol ions by comparing the metastable ion and collisional activation mass spectra of [CH2?CHCH?CHOH] and [CH2?C(OH)CH?CH2] ions with that of the above daughter ion. It is further concluded that prior to their major fragmentations by loss of CH3˙ and CO, [CH2?CHCH?CHOH]+˙ and [CH2?C(OH)CH?CH2] do not rearrange to their keto counterparts. The metastable ion and collisional activation characteristics of the isomeric allenic [C4H6O] ion [CH2?C?CHCH2OH] are also reported.  相似文献   

17.
Six solvated salts of a mononuclear manganese(III) complex with a chelating hexadentate Schiff base ligand are reported. One member of the series, [MnL]PF6.0.5 CH3OH ( 1 ), shows a rare low‐spin (LS) electronic configuration between 10–300 K. The remaining five salts, [MnL]NO3? C2H5OH ( 2 ), [MnL]BF4?C2H5OH ( 3 ), [MnL]CF3SO3?C2H5OH ( 4 ), [MnL]ClO4?C2H5OH ( 5 ) and [MnL]ClO4?0.5 CH3CN ( 6 ), all show gradual incomplete spin‐crossover (SCO) behaviour. The structures of all were determined at 100 K, and also at 293 K in the case of 3 – 6 . The LS salt [MnL]PF6.0.5 CH3OH is the only member of the series that does not exhibit strong hydrogen bonding. At 100 K two of the four SCO complexes ( 2 and 4 ) assemble into 1D hydrogen‐bonded chains, which weaken or rupture on warming. The remaining SCO complexes 3 , 5 and 6 do not form 1D hydrogen‐bonded chains, but instead exhibit discrete hydrogen bonding between cation/counterion, cation/solvent or counterion/solvent and show no significant change on warming.  相似文献   

18.
The kinetics of ozonation of C2H4 and C2H2 have been studied in the gas phase from ?40 to ?95°C (C2H4) and +10 to ?30°C (C2H2). The O3 concentrations were near 10?4 M, and the hydrocarbons were present in 2- to 25-fold excess. A few experiments with propylene were also carried out. The reactions were followed by observing the rate of decay of O3 absorption at 2537 Å. Reaction stoichiometries and effects of added O2 were investigated. The second-order rate constant for C2H4 was log k(M?1 sec?1) = (6.3 ± 0.2) – (4.7 ± 0.2)/θ (θ = 2.3RT). The rate was independent of the presence of excess O2. Rate measurements for C3H6 were less accurate because of aerosol interference. Combined with room temperature measurements of other workers, the C3H6 rate constant was log k(M?1 sec?1) = (6.0 ± 0.4) – (3.2 ± 0.6)/θ. The C2H2 rate constant was log k(M?1 sec?1) = (9.5 ± 0.4) – (10.8 ± 0.4)/θ. In the case of C3H6 the major product was propylene ozonide. Ethylene did not yield the ozonide, and the products of the O3–C2H4 and O3–C2H2 reactions were not identified. Pre-exponential factors for the olefin reactions are consistent with a five-membered ring transition state formed by 1,3 dipolar cycloaddition of O3. For C2H2, however, the much higher observed A factor suggests a different mechanism. Possible transition states for the O3–C2H2 reaction are discussed.  相似文献   

19.
MXenes, 2D compounds generated from layered bulk materials, have attracted significant attention in energy‐related fields. However, most syntheses involve HF, which is highly corrosive and harmful to lithium‐ion battery and supercapacitor performance. Here an alkali‐assisted hydrothermal method is used to prepare a MXene Ti3C2Tx (T=OH, O). This route is inspired from a Bayer process used in bauxite refining. The process is free of fluorine and yields multilayer Ti3C2Tx with ca. 92 wt % in purity (using 27.5 m NaOH, 270 °C). Without the F terminations, the resulting Ti3C2Tx film electrode (ca. 52 μm in thickness, ca. 1.63 g cm?3 in density) is 314 F g?1 via gravimetric capacitance at 2 mV s?1 in 1 m H2SO4. This surpasses (by ca. 214 %) that of the multilayer Ti3C2Tx prepared via HF treatments. This fluorine‐free method also provides an alkali‐etching strategy for exploring new MXenes for which the interlayer amphoteric/acidic atoms from the pristine MAX phase must be removed.  相似文献   

20.
Pb2(OH)2[p‐O2C‐C6H4‐CO2]: Synthesis and Crystal Structure Single crystals of Pb2(OH)2[p‐O2C‐C6H4‐CO2] ( 1 ) were obtained by hydrothermal reaction of terephthalic acid and PbCO3 at 180 °C (10 days). 1 crystallizes in the monoclinic space group P21/c with Z = 2 (a = 1115.6(2) pm, b = 380.10(4) pm, c = 1141.3(2) pm, β = 93.39(1)°, V = 0.4831(1) nm3). The crystal structure is characterized by ladder‐type Pb(OH)3/3 double chains, which are connected to a three‐dimensional framework by terephthalate dianions.  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