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.
徐洁  侯万国  周维芝  台培东  王文兴 《化学学报》2007,65(13):1191-1196
分别采用电势滴定(Potentiometric titration)法, 质量滴定(Mass titration)法和惰性电解质滴定(Inert electrolyte titration)法测定了荷结构负电荷的东北草甸棕壤的零净电荷点(PZNC), 研究表明荷结构负电荷的土壤依然存在与电解质浓度无关的零净电荷点, 三种方法的实验结果分别为2.9, 5.0和3.3. 相比较而言, 电势滴定法和惰性电解质滴定法的结果相近, 而质量滴定法的结果偏高. 对文献中推算表面质子活性位密度(Ns)的方法进行了改进, 得到东北草甸棕壤样品的Ns为2.5 mmol•g-1. 根据实验测定的pHPZNC, Ns和结构电荷密度(σst)值直接计算得出东北草甸棕壤样品的界面反应特征平衡常数即1-pK模型中的pK, 2-pK模型中pKa1int和pKa2int, 分别为3.37, 2.42和4.32, 与文献报道的沉积物和蒙脱土的值相吻合.  相似文献   

2.
焦燕妮  侯万国 《中国化学》2007,25(6):756-764
The interface electrochemical properties of clay were theoretically analyzed to obtain some relationships among point of zero net charge (PZNC), point of zero net proton charge (PZNPC), intrinsic surface reaction equilibrium constants (K in 1-pK model, Ka1^int and Ka2^int in 2-pK model, *KNa^int and *KNO3^int in inert electrolyte chemical binding model) and structural negative charge density (σst) of clay, and some interface electrochemical parameters of kaolinite were measured. The following main conclusions were obtained. For clay possessing structural negative charges, the PZNC independent of electrolyte concentration (c) should exist just as amphoteric solid without structural charges such as oxides or hydroxides. A common intersection point (CIP) should appear among the potentiometric (or acid-base) titration curves at different c and the pH at the CIP should be PHPZNC. A CIP among potentiometric titration curves at different c for kaolinite was observed, and the value of PHPZNC of kaolinite was 2.16. The values of pHPZNPC were decreased with increasing c, which arises from the presence of structural negative charges of kaolinite. In addition, it was observed that a good linear relationship existed between pHPZNPC and 1g c. According to the values of PHPZNC and σst measured, the intrinsic surface reaction equilibrium constants, pK and pKa1^int and pKa2^int of 1-pK and 2-pK models could be directly calculated for clay, and the values of pK, pKa1^int and pKa2^int for kaolinite were 2.93, 1.90 and 3.97, respectively. These experimental values of pKa1^int and pKa2^int for kaolinite are obviously lower than those optimized with fitting programs in literatures, which maybe arises from the introduction of a type of permanent negatively charged sites in the models of literatures. An interesting result obtained in this study is that the inert electrolyte chemical binding does not exist for kaolinite, which also arises from the presence of structural negative charges.  相似文献   

3.
The relative equations among intrinsic surface reaction equilibrium constants (K in 1-pK model, K(a1)(int) and K(a2)(int) in 2-pK model, and *K(Na)(int) and *K(Cl)(int) in inert electrolyte chemical binding model), points of zero charge (PZC), and structural charge density (sigma(st)) for amphoteric solids with structural charge were established to investigate the effects of sigma(st) on intrinsic equilibrium constants and PZC. The intrinsic equilibrium constants of HTlc with general formulas [(Zn,Mg)(1-x)Al(x)(OH)(2)](Cl,OH)(x) and [Mg(1-x)(Fe,Al)(x)(OH)(2)](Cl,OH)(x) were evaluated. The following main conclusions were obtained. For amphoteric solids with structural charge, a point of zero net charge (PZNC) independent of electrolyte concentration (c) exists. A common intersection point (CIP) should appear among the acid-base titration curves at different c, and the pH at the CIP is pH(PZNC). The pK, pK(a1)(int), and pK(a2)(int) may be expressed as a function of pH(PZNC) and sigma(st), and these intrinsic equilibrium constants can be directly calculated from pH(PZNC) and sigma(st). The inert electrolyte chemical binding does not exist for amphoteric surfaces with structural charge. PZNC is not equal to the point of zero net proton charge (PZNPC) when sigma(st) not equal 0. pH(PZNC) > pH(PZNPC) when sigma(st)>0; pH(PZNC) < pH(PZNPC) when sigma(st)<0; and pH(PZNC) = pH(PZNPC) only when sigma(st)=0. With increasing c, the difference between pH(PZNC) and pH(PZNPC) decreases; i.e., pH(PZNPC) moves forward to pH(PZNC) with increasing c. For the HTlc samples studied, with increasing x, the pH(PZNC) and the pK(a1)(int) and pK(a2)(int) decrease, and the pK increases. These results can be explained on the basis of the affinity of metal cations for H(+) or OH(-) and the electrostatic interaction between the charging surface and H(+) or OH(-).  相似文献   

4.
A kinetic study on hydrolysis of N‐(2′‐hydroxyphenyl)phthalamic acid ( 1 ), N‐(2′‐methoxyphenyl)phthalamic acid ( 2 ), and N‐(2′‐methoxyphenyl)benzamide ( 3 ) under a highly alkaline medium gives second‐order rate constants, kOH, for the reactions of HO? with 1, 2 , and 3 as (4.73 ± 0.36) × 10?8 at 35°C, (2.42 ± 0.28) × 10?6 and (5.94 ± 0.23) × 10?5 M?1 s?1 at 65°C, respectively. Similar values of kOH for 3 , N‐methylbenzanilide, N‐methylbenzamide, and N,N‐dimethylbenzamide despite the difference between pKa values of aniline and ammonia of ~10 pK units are qualitatively explained. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 41: 1–11, 2009  相似文献   

5.
Plating battery electrodes typically deliver higher specific capacity values than insertion or conversion electrodes because the ion charge carriers represent the sole electrode active mass, and a host electrode is unnecessary. However, reversible plating electrodes are rare for electronically insulating nonmetals. Now, a highly reversible iodine plating cathode is presented that operates on the redox couples of I2/[ZnIx(OH2)4?x]2?x in a water‐in‐salt electrolyte. The iodine plating cathode with the theoretical capacity of 211 mAh g?1 plates on carbon fiber paper as the current collector, delivering a large areal capacity of 4 mAh cm?2. Tunable femtosecond stimulated Raman spectroscopy coupled with DFT calculations elucidate a series of [ZnIx(OH2)4?x]2?x superhalide ions serving as iodide vehicles in the electrolyte, which eliminates most free iodide ions, thus preventing the consequent dissolution of the cathode‐plated iodine as triiodides.  相似文献   

6.
侯万国  REN Carolyn 《中国化学》2006,24(10):1336-1341
The intrinsic surface reaction constants, pKa1^int, pKa2^int, p^*KC^int and p^*KA^int , were evaluated by a modifieddouble extrapolation (MDE) for TiO2 without structural charge and Mg-Fe hydrotalcite-like compounds (HTlc) with structural charge, respectively. The results of intrinsic surface reaction constants for TiO2 were compared with those obtained by class double extrapolation (CDE) in literature. Furthermore, the values of intrinsic surface reaction constants obtained by MDE were used to simulate the charging behaviors of the materials. The following conclusions were obtained. For TiO2 without structural charge, the pKa1^int and pKa2^int evaluated by MDE are equal to those by CDE, however the p^*KC^int and p^*KA^int evaluated by MDE are much different from those by CDE. In principle, the results of the p^*KC^int and p^*KA^int evaluated by MDE are more accurate than those by CDE. The values of intrinsic surface reaction constants obtained by MDE can excellently simulate the charging curves for TiO2 with the triple layer model (TLM). For HTlc with positive structural charge, the results of ^*KC^int=0 and ^*KA^int →∞ were obtained by MDE, which means the inert electrolyte chemical binding does not exist; the point of zero net charge (PZNC) of c-independence also exist as the same as solid without structural charge, and the PHPZNC obtained by the acid-base titration can excellently be simulated and the surface charging tendency can be simulated to a great extent using the pKa1^int and pKa2^int evaluated by MDE and the diffuse layer model (DLM).  相似文献   

7.
Zinc oxide photocatalyst was synthesized through a low‐temperature co‐precipitation process using zinc sulfate as precursor for the degradation of Acid Red 57 (AR57) under UV irradiation. The activities of the prepared photocatalyst at different calcination temperatures (400, 500, and 600 °C) were investigated. The synthesized zinc oxides were characterized by different techniques such as X‐ray diffraction, scanning electron microscopy, energy‐dispersive X‐ray spectroscopy, N2 adsorption–desorption, and pH titration for the determination of the zero‐point charge (pHZPC). The efficiency of photocatalytic degradation of ZnO prepared at the calcination temperatures of 400, 500, and 600 °C was 90.03, 77.67, and 72.71%, respectively, after 190 min. The kinetics and scavengers of the reactive species during the degradation were also investigated. It was found that the degradation of AR57 fitted first‐order kinetics and the OH? radicals were the main species. During irradiation, the formation of OH? free radicals was ascertained by photoluminescence studies using terephthalic acid as the probe molecule.  相似文献   

8.
A new chromophore has been identified in copper‐doped apatite pigments having the general composition [Sr10(PO4)6(CuxOH1?x?y)2], in which x=0.1, 0.3 and y=0.01–0.42. By using X‐ray absorption spectroscopy, low‐temperature magnetization measurements, and synchrotron X‐ray powder structure refinement, it has been shown that the oxygenated compounds contain simultaneously diamagnetic Cu1+ and paramagnetic Cu3+ with S=1. Cu3+ is located at the same crystallographic position as Cu1+, being linearly coordinated by two oxygen atoms and forming the OCuO? anion. The Raman spectroscopy study of [A10(PO4)6(CuxOH1?x?y)2,], in which A=Ca, Sr, Ba, reveals resonance bands at 651–656 cm?1 assigned to the symmetric stretching vibration (ν1) of OCuO?. The strontium apatite pigment exhibits a strong paramagnetic anisotropy with an unprecedentedly large negative zero‐field splitting parameter (D) of ≈?400 cm?1. The extreme magnetic anisotropy causes slow magnetization relaxation with relaxation times (τ) up to 0.3 s at T=2 K, which relates the compounds to single‐ion magnets. At low temperature, τ is limited by a spin quantum‐tunneling, whereas at high temperature a thermally activated relaxation prevails with Ueff≈48 cm?1. Strong dependence of τ on the paramagnetic center concentration at low temperature suggests that the spin‐spin relaxation dominates in the spin quantum‐tunneling process. The compound is the first example of a d‐metal‐based single‐ion magnet with S=1, the smallest spin at which an energy barrier arises for the spin flipping.  相似文献   

9.
MP2/aug′‐cc‐pVTZ calculations were performed to investigate boron as an electron‐pair donor in halogen‐bonded complexes (CO)2(HB):ClX and (N2)2(HB):ClX, for X=F, Cl, OH, NC, CN, CCH, CH3, and H. Equilibrium halogen‐bonded complexes with boron as the electron‐pair donor are found on all of the potential surfaces, except for (CO)2(HB):ClCH3 and (N2)2(HB):ClF. The majority of these complexes are stabilized by traditional halogen bonds, except for (CO)2(HB):ClF, (CO)2(HB):ClCl, (N2)2(HB):ClCl, and (N2)2(HB):ClOH, which are stabilized by chlorine‐shared halogen bonds. These complexes have increased binding energies and shorter B?Cl distances. Charge transfer stabilizes all complexes and occurs from the B lone pair to the σ* Cl?A orbital of ClX, in which A is the atom of X directly bonded to Cl. A second reduced charge‐transfer interaction occurs in (CO)2(HB):ClX complexes from the Cl lone pair to the π* C≡O orbitals. Equation‐of‐motion coupled cluster singles and doubles (EOM‐CCSD) spin–spin coupling constants, 1xJ(B‐Cl), across the halogen bonds are also indicative of the changing nature of this bond. 1xJ(B‐Cl) values for both series of complexes are positive at long distances, increase as the distance decreases, and then decrease as the halogen bonds change from traditional to chlorine‐shared bonds, and begin to approach the values for the covalent bonds in the corresponding ions [(CO)2(HB)?Cl]+ and [(N2)2(HB)?Cl]+. Changes in 11B chemical shieldings upon complexation correlate with changes in the charges on B.  相似文献   

10.
The mechanism of the reaction of trans‐ArPdBrL2 (Ar=p‐Z‐C6H4, Z=CN, H; L=PPh3) with Ar′B(OH)2 (Ar′=p‐Z′‐C6H4, Z′=H, CN, MeO), which is a key step in the Suzuki–Miyaura process, has been established in N,N‐dimethylformamide (DMF) with two bases, acetate (nBu4NOAc) or carbonate (Cs2CO3) and compared with that of hydroxide (nBu4NOH), reported in our previous work. As anionic bases are inevitably introduced with a countercation M+ (e.g., M+OH?), the role of cations in the transmetalation/reductive elimination has been first investigated. Cations M+ (Na+, Cs+, K+) are not innocent since they induce an unexpected decelerating effect in the transmetalation via their complexation to the OH ligand in the reactive ArPd(OH)L2, partly inhibiting its transmetalation with Ar′B(OH)2. A decreasing reactivity order is observed when M+ is associated with OH?: nBu4N+> K+> Cs+> Na+. Acetates lead to the formation of trans‐ArPd(OAc)L2, which does not undergo transmetalation with Ar′B(OH)2. This explains why acetates are not used as bases in Suzuki–Miyaura reactions that involve Ar′B(OH)2. Carbonates (Cs2CO3) give rise to slower reactions than those performed from nBu4NOH at the same concentration, even if the reactions are accelerated in the presence of water due to the generation of OH?. The mechanism of the reaction with carbonates is then similar to that established for nBu4NOH, involving ArPd(OH)L2 in the transmetalation with Ar′B(OH)2. Due to the low concentration of OH? generated from CO32? in water, both transmetalation and reductive elimination result slower than those performed from nBu4NOH at equal concentrations as Cs2CO3. Therefore, the overall reactivity is finely tuned by the concentration of the common base OH? and the ratio [OH?]/[Ar′B(OH)2]. Hence, the anionic base (pure OH? or OH? generated from CO32?) associated with its countercation (Na+, Cs+, K+) plays four antagonist kinetic roles: acceleration of the transmetalation by formation of the reactive ArPd(OH)L2, acceleration of the reductive elimination, deceleration of the transmetalation by formation of unreactive Ar′B(OH)3? and by complexation of ArPd(OH)L2 by M+.  相似文献   

11.
The ionization (or basicity) constants (pKb) were determined for many 2‐substituted 4,6‐diamino‐s‐tri‐azines ( I ) by means of the electrometric titration. I includes 2‐alkoxy or aryloxy‐( Ia ), 2‐alkyl‐ or 2‐aryl‐( Ib ), and 2‐alkylamino‐ or 2‐arylamino‐4,6‐diamino‐s‐triazines ( Ic ). For the series with the same alkyl or aryl group, the order of the basicity was found to be Ic < Ib < Ia . A study was made of relationships between the pKb, values of I , and the substituent constants, σp, σm, σp+, σm+, σpO, σmo, σI, σn, and σ*. The Hammett relationships were observed between the pKa values of I, and the substituent constants σm, (or the combination ones, [0.97σm + 0.03σp] as well as another [0.77σI + 0.23σR]). The Taft relationships were also found between the pKa values of Ia , Ib , and Ic and the constants σ*, respectively. Furthermore, in the case of Ic a linear relationship was observed between the pKa values and Σσ8.  相似文献   

12.
Borocarbonitrides (BxCyNz) with a graphene‐like structure exhibit a remarkable high lithium cyclability and current rate capability. The electrochemical performance of the BxCyNz materials, synthesized by using a simple solid‐state synthesis route based on urea, was strongly dependent on the composition and surface area. Among the three compositions studied, the carbon‐rich compound B0.15C0.73N0.12 with the highest surface area showed an exceptional stability (over 100 cycles) and rate capability over widely varying current density values (0.05–1 A g?1). B0.15C0.73N0.12 has a very high specific capacity of 710 mA h g?1 at 0.05 A g?1. With the inclusion of a suitable additive in the electrolyte, the specific capacity improved drastically, recording an impressive value of nearly 900 mA h g?1 at 0.05 A g?1. It is believed that the solid–electrolyte interphase (SEI) layer at the interface of BxCyNz and electrolyte also plays a crucial role in the performance of the BxCyNz .  相似文献   

13.
Quantitative data on the stability of mono‐, di‐ and trimethyltin(IV)‐carboxylate complexes (acetate, malonate, succinate, malate, oxydiacetate, diethylenetrioxydiacetate, tricarballylate, citrate, butanetetracarboxylate and mellitate) are reported at t = 25 °C and I→ 0 mol l?1. Several mononuclear, mixed proton, mixed hydroxo and polynuclear species are formed in these systems. As expected, the stability trend is mono‐ > di‐ > trimethyltin(IV) and mono < di < tri < tetra < hexa for the organotin moieties and carboxylate ligands investigated, respectively. Moreover, ligands containing, in addition to carboxylic,? O? and? OH groups show a significantly higher stability with respect to analogous ligands with the same number of carboxylic binding sites. The results obtained from all the systems investigated allowed us to formulate the following empirical predictive equation for correlation between complex stability and some simple structural parameters, (1) where ncarb and nOH are the number of carboxylic and alcoholic groups in the ligand, respectively, r is the stoichiometric coefficient of H+ (positive) or OH? (negative) and zcat is the methyltin cation charge (CH3)xSnz+ (z+ = 4 ? x). Distribution diagrams for some representative systems are also reported and are discussed in the light of speciation studies in natural waters. A literature data comparison is made with carboxylate complexes of other metal ions with the same charge as the organotin cations investigated here. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

14.
We demonstrate a unique synthetic route for oxygen‐deficient mesoporous TiOx by a redox–transmetalation process by using Zn metal as the reducing agent. The as‐obtained materials have significantly enhanced electronic conductivity; 20 times higher than that of as‐synthesized TiO2 material. Moreover, electrochemical impedance spectroscopy (EIS) and galvanostatic intermittent titration technique (GITT) measurements are performed to validate the low charge carrier resistance of the oxygen‐deficient TiOx. The resulting oxygen‐deficient TiOx battery anode exhibits a high reversible capacity (~180 mA h g?1 at a discharge/charge rate of 1 C/1 C after 400 cycles) and an excellent rate capability (~90 mA h g?1 even at a rate of 10 C). Also, the full cell, which is coupled with a LiCoO2 cathode material, exhibits an outstanding rate capability (>75 mA h g?1 at a rate of 3.0 C) and maintains a reversible capacity of over 100 mA h g?1 at a discharge/charge of 1 C/1 C for 300 cycles.  相似文献   

15.
Fundamental photocatalytic limitations of solar CO2 reduction remain due to low efficiency, serious charge recombination, and short lifetime of catalysts. Herein, two-dimensional graphitic carbon nitride nanosheets with nitrogen vacancies (g-C3Nx) located at both three-coordinate N atoms and uncondensed terminal NHx species were prepared by one-step tartaric acid-assistant thermal polymerization of dicyandiamide. Transient absorption spectra revealed that the defects in g-C3N4 act as trapped states of charges to result in prolonged lifetimes of photoexcited charge carriers. Time-resolved photoluminescence spectroscopy revealed that the faster decay of charges is due to the decreased interlayer stacking distance in g-C3Nx in favor of hopping transition and mobility of charge carriers to the surface of the material. Owing to the synergic virtues of strong visible-light absorption, large surface area, and efficient charge separation, the g-C3Nx nanosheets with negligible loss after 15 h of photocatalysis exhibited a CO evolution rate of 56.9 μmol g−1 h−1 under visible-light irradiation, which is roughly eight times higher than that of pristine g-C3N4. This work presents the role of defects in modulating light absorption and charge separation, which opens an avenue to robust solar-energy conversion performance.  相似文献   

16.
A simple one‐pot synthesis of metal selenide/reduced graphene oxide (rGO) composite powders for application as anode materials in sodium‐ion batteries was developed. The detailed mechanism of formation of the CoSex–rGO composite powders that were selected as the first target material in the spray pyrolysis process was studied. The crumple‐structured CoSex–rGO composite powders prepared by spray pyrolysis at 800 °C had a crystal structure consisting mainly of Co0.85Se with a minor phase of CoSe2. The bare CoSex powders prepared for comparison had a spherical shape and hollow structure. The discharge capacities of the CoSex–rGO composite and bare CoSex powders in the 50th cycle at a constant current density of 0.3 A g?1 were 420 and 215 mA h g?1, respectively, and their capacity retentions measured from the second cycle were 80 and 46 %, respectively. The high structural stability of the CoSex–rGO composite powders for repeated sodium‐ion charge and discharge processes resulted in superior sodium‐ion storage properties compared to those of the bare CoSex powders.  相似文献   

17.
High‐valent iron‐oxo species have been invoked as reactive intermediates in catalytic cycles of heme and nonheme enzymes. The studies presented herein are devoted to the formation of compound II model complexes, with the application of a water soluble (TMPS)FeIII(OH) porphyrin ([meso‐tetrakis(2,4,6‐trimethyl‐3‐sulfonatophenyl)porphinato]iron(III) hydroxide) and hydrogen peroxide as oxidant, and their reactivity toward selected organic substrates. The kinetics of the reaction of H2O2 with (TMPS)FeIII(OH) was studied as a function of temperature and pressure. The negative values of the activation entropy and activation volume for the formation of (TMPS)FeIV?O(OH) point to the overall associative nature of the process. A pH‐dependence study on the formation of (TMPS)FeIV?O(OH) revealed a very high reactivity of OOH? toward (TMPS)FeIII(OH) in comparison to H2O2. The influence of N‐methylimidazole (N‐MeIm) ligation on both the formation of iron(IV)‐oxo species and their oxidising properties in the reactions with 4‐methoxybenzyl alcohol or 4‐methoxybenzaldehyde, was investigated in detail. Combined experimental and theoretical studies revealed that among the studied complexes, (TMPS)FeIII(H2O)(N‐MeIm) is highly reactive toward H2O2 to form the iron(IV)‐oxo species, (TMPS)FeIV?O(N‐MeIm). The latter species can also be formed in the reaction of (TMPS)FeIII(N‐MeIm)2 with H2O2 or in the direct reaction of (TMPS)FeIV?O(OH) with N‐MeIm. Interestingly, the kinetic studies involving substrate oxidation by (TMPS)FeIV?O(OH) and (TMPS)FeIV?O(N‐MeIm) do not display a pronounced effect of the N‐MeIm axial ligand on the reactivity of the compound II mimic in comparison to the OH? substituted analogue. Similarly, DFT computations revealed that the presence of an axial ligand (OH? or N‐MeIm) in the trans position to the oxo group in the iron(IV)‐oxo species does not significantly affect the activation barriers calculated for C?H dehydrogenation of the selected organic substrates.  相似文献   

18.
This paper deals with the problem of defining, and measuring, the pH inside the water pool (which we define as pHwp) of reverse micelles, i.e. micelles formed by surfactants dissolved in apolar solvents in the presence of minimal amounts of water. The conceptual and experimental difficulties are discussed, and it is argued that no absolute determination of pHwp is possible, mostly because water in the water pools of reverse micelles is a new solvent, for which no standardization of acidity is available. The problem can be approached only on the basis of an empirical acidity scale. An empirical acidity scale for water pools in reverse micelles of bis (2-ethyl-hexyl) sodium sulfosuccinate (AOT) in isooctane has been defined by measuring the 31P-chemical shifts of phosphate buffers. The chemical shifts in bulk water were compared to those found in reverse micelles under the assumption that the pK of phosphate ion is the same in the two systems. It was found that in most cases there was little difference (less than 0.4 pH units) between pHwp and the pH of the starting buffer in bulk water (which we define as pHst). However, this difference between pHwp and pHst may become much larger in certain cases. The difference (pHwp–pHst) is measured under a variety of conditions, and this permits the determination of an operational acidity in the micelle water pools as a function of the pHst with which the aqueous micelles are prepared. The significance of such data for interpreting the behaviour of enzymes confined in the micelles water pool is discussed. Based on the pHwp scale, the apparent pKa of phenol-red and 4-nitrophenol were determined in reverse micelles containing different buffers and different water content. The pKa values obtained were rather sensitive to changes of both these factors, which was taken to signify that organic dies have only a very limited applicability to measure the acidity of the water pools of reverse micelles.  相似文献   

19.
This study introduces an in situ fabrication of nanoporous hematite with a Ti‐doped SiOx passivation layer for a high‐performance water‐splitting system. The nanoporous hematite with a Ti‐doped SiOx layer (Ti‐(SiOx/np‐Fe2O3)) has a photocurrent density of 2.44 mA cm?2 at 1.23 VRHE and 3.70 mA cm?2 at 1.50 VRHE. When a cobalt phosphate co‐catalyst was applied to Ti‐(SiOx/np‐Fe2O3), the photocurrent density reached 3.19 mA cm?2 at 1.23 VRHE with stability, which shows great potential of the use of the Ti‐doped SiOx layer with a synergistic effect of decreased charge recombination, the increased number of active sites, and the reduced hole‐diffusion pathway from the hematite to the electrolyte.  相似文献   

20.
Anion sensor properties of N‐alkyl‐substituted 1,4′‐diazaflavonium bromides in methanol–water were evaluated by UV–vis spectrometry. Pronounced changes were observed in the absorption spectra of all compounds for only OH?, CO32?, and CN? among F?, Cl?, Br?, I?, OH?, CO32?, NO3?, PO43?, CN?, SO42?, HSO4?, HCO3?, SCN?, NO2?, and P2O72? ions. Two new absorption bands at 385 and 685 nm accompanying the distinct color change for OH?, CO32?, and CN? ions were observed in case of all compounds. The color changes were from pink to blue for CO32? and OH? ions and from pink to purple for CN? ion. Thanks to the distinct color change, the compounds can be used as selective colorimetric anion sensors. Linear changes of absorbance of N‐heptyl‐substituted compound at 385 nm as a function of the ion concentration were used to determine CN? ion in water samples. Detection and quantification limits of the proposed method were 0.94 and 2.82 mg/L, respectively.  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