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.
The sheet flexibility of layered double hydroxides (LDHs) has been investigated experimentally using co-precipitation and urea hydrolysis methods in an aqueous solution of long-chain anion surfactant in this work. Using dodecylsulfate (DS) anion as morphology-controlling agent, layer-bended or contorted Mg/Al-LDH is obtained successfully. The morphology of bent layers is retained during either in situ decomposition of interlayer DS to SO42− or ion exchange of interlayer DS by CO32−. The direct synthesis of the layer-distorted LDHs intercalated with small inorganic anions is quite difficult. It has been achieved using layer-bended LDHs pillared with bulky organic anions as precursors in this paper. The morphosynthesis is expanded to Co/Al and Ni/Al-LDHs, indicative of the general flexibility of this kind of anionic clays.  相似文献   

2.
Layered double hydroxides (LDHs) have high surface area and high anion exchange capacity, so they have been proposed to be an effective scavenger for contaminants. In this study, the adsorption of 2,4-dichlorophenoxyacetate (2,4-D) on Mg/Al–NO3 LDHs with varying layer charge density was investigated with particular attention on the effect of the orientation of the interlayer nitrate. Three Mg/Al LDHs were synthesized with Al3+/(Al3+ + Mg2+) molar ratios of 3.3 (LDH3), 2.6 (LDH4) and 2.1 (LDH5). The results of adsorption experiments showed that LDH5 exhibited an S-type isotherm with a low 2,4-D adsorption capacity due to the low accessibility of 2,4-D to the interlayer space. The accessibility was restricted by the small basal spacing of LDH5 as a result of the parallel orientation of the interlayer nitrate with respect to the hydroxide sheet. Thus, the 2,4-D adsorption occurred mainly on the external surface of the material. On the contrary, LDH3, which has the highest layer charge density among the samples, contains nitrate with an orientation perpendicular to the hydroxide sheet of LDH3. The interlayer nitrate was readily exchanged by 2,4-D. Thus, in addition to the adsorption on the external surface, the replacement of the interlayer nitrate by 2,4-D contributed to a higher adsorbed amount of 2,4-D; the 2,4-D adsorption of LDH3 exhibited an L-type isotherm. For LDH4 that contained interlayer nitrate with both parallel and perpendicular orientations, the adsorption characteristics were between those of LDH3 and LDH5. This work has demonstrated the dependence of 2,4-D adsorption characteristics on the nitrate orientation in LDHs, as a consequence of changing layer charge density.  相似文献   

3.
Pillared derivatives of Mg1−xAlx layered double hydroxides (LDHs) were prepared by anion exchange reaction of a synthetic meixnerite precursor, [Mg3Al(OH)2](OH), with macromolecular polyoxometalate ions. The intercalated polyoxometalates included the lacunary Dawson ion (α-P2W17O61)10−, the Finke (Zn4(H2O)2(AsW9O34)2)10− and (WZn3(H2O)2(ZnW9O34)2)12− ions, the doubled Dawson (P4W30Zn4(H2O)2O112)16− and the polyoxocryptates (NaSb9W21O86)18− and (NaP5W30O110)14−. Anion exchange reaction of [Zn2Al(OH)2](NO3) with (NaP5W30O110)14− also resulted in a crystalline pillared product. The intercalates exhibited gallery heights up to 16.6 Å and thermal stabilities to 200°C. Nitrogen adsorption/desorption studies for the LDH intercalates showed that access to the gallery micropores was achieved upon POM intercalation. All of the intercalates contained a salt-like impurity phase, as indicated by XRD. The Zn2Al–(NaP5W30O110)14− LDH was investigated as a catalyst for the peroxide oxidation of cyclohexene. A comparison of the reactivities of three samples containing different fractions of the salt-like impurity suggested that the impurity phase contributes significantly to the observed activity.  相似文献   

4.
The hydration of layered double hydroxides (LDHs) was investigated by changing the interlayer anion species, the Mg/Al ratio of the LDH hosts and the relative humidity (RH). The anions were CO32−, Cl, Br, NO3, I, SO42−, and ClO4 (listed in the order of ion size, small to large) and LDHs with Mg/Al = 1.90 (LDH2) and 2.91 (LDH3) were used. Their XRD profiles were measured by an XRD diffractometer while controlling the RH in the range 0–95% at 25 °C. Only I, SO42−, and ClO4 LDH2s and SO42− LDH3 showed a large step-wise basal-spacing expansion, 0.24–0.28 nm, under high RH conditions (> ca. 60%) probably due to the insertion of one water layer into the interlayer space. Such hydration occurred more favorably for the LDHs with larger anions and those with a higher layer charge (LDH2). Among them, I and ClO4 LDH2s exhibited the second staging – alternate stacking of hydrated (H) and non-hydrated (NH) interlayers – in the intermediate RH region.  相似文献   

5.
Chitosan, a natural biopolymeric cation, is a candidate to modify montmorillonite for the adsorption of anions. As an anionic organic pollutant the adsorption of tannic acid was studied. Because of protonation/deprotonation reactions of both chitosan-montmorillonite and tannic acid, the adsorption process is strongly pH-dependent. The objective of this work is to characterize the pH dependency of adsorption in combination with surface charge determinations.Montmorillonite was modified with different amounts of chitosan, corresponding to 20–1000% of the cation exchange capacity (CEC). The deacetylation degree of chitosan was determined by polyelectrolyte titration and was found to be 74%. The uptake of chitosan was determined by the C-content. The interlayer expansion was investigated by X-ray powder diffraction. The adsorption capacity for tannic acid was investigated with the batch technique at pH 3, 4, 5 and 8. As a measure for the adsorption properties, the electrokinetic surface charge was determined with a particle charge detector.The uptake of chitosan by montmorillonite is up to 152% (1.69 molc kg− 1) of the CEC. The resulting anion exchange capacity of chitosan-montmorillonite calculated from C-content is 0.43 molc kg− 1. At low loadings with chitosan (24.7 and 49.5% uptake), a monolayer is formed in montmorillonite. At an uptake of 96.8%, a bilayer structure is observed, which becomes more dominant at higher loadings. On the external surface, a monolayer of chitosan was formed. From pH 4 to 8, the surface charge of all modified montmorillonites is with − 9 to 8 mmolc kg− 1 close to the point of zero charge. The maximal adsorption capacity for tannic acid is found with 240 g kg− 1 (0.14 molc kg− 1) at pH 4. The adsorption process fits in well with the Freundlich isotherm. At lower as well as higher pH values the adsorption capacity decreases up to about 25%. Most probably the exchange sites in the interlayer do not contribute to the adsorption of tannic acid. The observed surface charge is lower than the adsorbed amount of tannin. It is thought that tannin is adsorbed also by van der Waals forces besides ionic forces.  相似文献   

6.
Chromate anion was effectively concentrated on an anion exchanger, layered double hydroxide (LDH), from aqueous solution. The LDHs used in this study were synthesized by hydrothermal reactions using urea hydrolysis as reported previously (Ogawa and Kaiho, 2002). The ion exchange was conducted at pH = 3 to concentrate larger amount of dichromate (Cr2O72 ). The adsorption isotherms indicated the strong interactions of chromate anions with LDHs. Among the tested LDHs (MgAl, NiAl, CoAl, and LiAl-LDHs), LiAl-LDH gave the best result in terms of the adsorbed Cr amount. Since the reaction was started from carbonate forms of layered double hydroxides, only the carbonate anions are released to the environment after the anion exchange. All these facts indicated that the present layered double hydroxides are promising materials for the environmental remediation from Cr contaminated water. If compared with commercially available LDHs and those synthesized by the conventional co-precipitation methods, the LDHs synthesized by hydrothermal reaction using urea hydrolysis were shown to be superior because of the stability and the ease of solid–liquid separation after the ion exchange.  相似文献   

7.
Layered double hydroxides (LDHs) are extensively studied as precursors for catalysts, following a calcination at high temperatures to yield mixed oxides. However, these materials are less used as layered materials, i.e., without undergoing thermal activation. We have focused in this work on the use of a series of as-synthesised LDHs for the catalytic preparation of glycol ethers, which is a reaction of primary commercial importance. Two main systems are considered, namely the [Cu–Cr] and [Mg–Al] LDHs. The [Cu–Cr–Cl] LDH is obtained by the coprecipitation method, then through the appropriate chimie douce exchange reactions the original chloride anions are replaced by a variety of oxo- and polyoxometallates, (CrO4)2−, (Cr2O7)2−, (V2O7)2−, (V10O28)6− and (Mo7O24)6−. On the other hand, the [Mg–Al] hydrotalcites, intercalated by (V2O7)2−, (V10O28)6− and [FeIII(CN)6]3− anions, are obtained by structure regeneration. This was done by rehydration of a commercial calcined material (Kyowa) in aqueous solutions containing the desired anion. The different materials have been fully characterised by conventional analytical techniques to evidence their lamellar properties and chemical nature. They were then tested in the catalytic reaction involving butan-1-ol and one or more units of ethylene oxide to make butyl-monoglycol ether (BMGE), di-glycol ether (BDGE), tri-glycol ether (BTGE), etc. The reactions were carried out between 80 and 120°C, temperature range in which no collapse of the lamellar structure is normally observed. In this paper it is shown that decavanadate exchanged LDHs proved to be very selective catalysts for the preparation of the monoglycol adduct, some samples achieving up to 100% selectivity in the screening tests.  相似文献   

8.
Mixed oxides, prepared via the thermal decomposition of layered double hydroxides (LDHs), were screened gravimetrically for their ability to adsorb carbonyl sulfide (COS). Based on promising results obtained for Ni/Mg/Al, Ni/Mg/Fe and Co/Mg/Al mixed oxides, a study was undertaken to optimize the composition of these materials for COS adsorption. To investigate the effect of the M(II):M(III) ratio, LDHs of the type [MzMgyAlx(OH)2](CO3)x/2·0.5H2O (where M = Ni or Co, and x + y + z = 1) were prepared at values of x corresponding to 0.33 and 0.20. Simultaneously, the elemental ratio of transition metal to magnesium (z/y) was varied. Mixed oxides obtained from the resulting LDHs were tested in fixed bed mode with a feed of 100 ppm COS in N2 to determine breakthrough capacity. In general Ni/Mg/Al mixed oxides showed the best performance, a composition with Ni/Mg/Al = 0.32/0.48/0.20 showing the best adsorption capacity. Treatment of the spent adsorbent under an atmosphere of 5% H2 in N2 at 450 °C was found to provide an effective means of restoring the adsorption capacity over two cycles of adsorption and regeneration, although after three such cycles, adsorption capacity decreased.  相似文献   

9.
《Applied Clay Science》2009,44(3-4):459-464
Nitrate forms of layered double hydroxides (LDHs) were synthesized based on the co-precipitation method under different synthesis conditions (aqueous ammonia solution or potassium hydroxide to control the pH of the solution; magnesium to aluminium ratios of 2:1 and 5:1). In a second step the lab procedure was up-scaled to pilot plant scale. The effects of the synthesis conditions on the structural and textural properties as well as on the anion exchange of the LDH products were investigated using different methods to evaluate the suitability as a soil conditioner. Using a Mg:Al ratio of 2:1 resulted in higher basal spacings compared to a Mg:Al ratio of 5:1, while the type of pH-controlling solution had no effect. Based on nitrogen adsorption isotherms the specific surface area of pores was calculated. The highest values for specific surface areas were found for LDHs synthesized using KOH at a Mg:Al ratio of 5:1. These products had a lower total nitrate adsorption capacity compared to LDHs synthesized at a lower Mg:Al ratio. However, nitrate exchangeability by counter anions for LDHs synthesized using KOH at a Mg:Al ratio 5:1 differed significantly (HCO3 > Cl > SO42−) indicating that these LDHs are preferable under a multi-anionic environment like the soil solution.  相似文献   

10.
Cr-doped Li3V2−xCrx(PO4)3/C (x = 0, 0.05, 0.1, 0.2, 0.5, 1) compounds have been prepared using sol–gel method. The Rietveld refinement results indicate that single-phase Li3V2−xCrx(PO4)3/C with monoclinic structure can be obtained. Although the initial specific capacity decreased with Cr content at a lower current rate, both cycle performance and rate capability have excited improvement with moderate Cr-doping content in Li3V2−xCrx(PO4)3/C. Li3V1.9Cr0.1(PO4)3/C compound presents an initial capacity of 171.4 mAh g−1 and 78.6% capacity retention after 100 cycles at 0.2C rate. At 4C rate, the Li3V1.9Cr0.1(PO4)3/C can give an initial capacity of 130.2 mAh g−1 and 10.8% capacity loss after 100 cycles where the Li3V2(PO4)3/C presents the initial capacity of 127.4 mAh g−1 and capacity loss of 14.9%. Enhanced rate and cyclic capability may be attributed to the optimizing particle size, carbon coating quality, and structural stability during the proper amount of Cr-doping (x = 0.1) in V sites.  相似文献   

11.
Uptake of cesium and strontium cations by potassium-depleted phlogopite   总被引:2,自引:1,他引:1  
Phlogopite mica was equilibrated with 1.0 N sodium chloride (NaCl)–0.2 N sodium tetraphenylborate (NaTPB)–0.01 M disodium ethylenediaminetetraacetic acid (EDTA) solution at room temperature resulting in an almost complete removal (92%) of the mica's interlayer K. X-ray powder diffraction analysis provides additional evidence that hydrated Na+ ions had almost completely replaced the interlayer K+. Following equilibration, the c-axis spacing of the mica increased from 10.0 Å to approximately 12.2 Å. Cesium and Sr ion exchange isotherms indicate that K-depleted phlogopite is highly selective for both elements, the Cs+ exchange capacity is 1.26 meq/g or 65% of the theoretical cation exchange capacity and the Sr2+ exchange capacity is 1.94 meq/g or 100% of the theoretical exchange capacity of the mica. Kielland plots indicated that the mica was selective for Cs+ when the equivalent exchange capacity of Cs+ in the exchanger phase (Cs) was < 0.66 and selective for Sr2+ when Sr < 0.41. At equivalent fractions greater than these levels, layer collapse and/or steric effects limit the diffusion of these ions into the interlayers of the mica. Analysis of the Cs+ equilibrated mica utilizing XRD indicated that a collapse of the c-axis spacing had occurred. Based on the high selectivity of < 45-μm K-depleted phlogopite for Sr2+ and Cs+, this material may prove useful as an inorganic ion exchanger for these radioactive isotopes.  相似文献   

12.
Mg–Al layered double hydroxide (Mg–Al LDH) was modified with organic acid anions using a coprecipitation technique, and the uptake of heavy metal ions from aqueous solution by the Mg–Al LDH was studied. Citrate·Mg–Al LDH, malate·Mg–Al LDH, or tartrate·Mg–Al LDH, which had citrate3− (C6H5O73−), malate2− (C4H4O52−), or tartrate2− (C4H4O62−) anions intercalated in the interlayer, was prepared by dropwise addition of a mixed aqueous solution of Mg(NO3)2 and Al(NO3)3 to a citrate, malate, or tartrate solution at a constant pH of 10.5. These Mg–Al LDHs were found to take up Cu2+ and Cd2+ rapidly from an aqueous solution at a constant pH of 5.0. This capacity was mainly attributable to the formation of the citrate–metal, malate–metal, and tartrate–metal complexes in the interlayers of the Mg–Al LDHs. The uptake of Cu2+ increased in the order malate·Mg–Al LDH < tartrate·Mg–Al LDH < citrate·Mg–Al LDH. The uptake of Cd2+ increased in the order malate·Mg–Al LDH < tartrate·Mg–Al LDH = citrate·Mg–Al LDH. These differences in Cu2+ and Cd2+ uptake were attributable to differences in the stabilities of the citrate–metal, malate–metal, and tartrate–metal complexes. These results indicate that citrate3−, malate2−, and tartrate2− were adequately active as chelating agents in the interlayers of Mg–Al LDHs.  相似文献   

13.
Bozenna Pisarska   《Desalination》2008,230(1-3):298-304
Flow rates of co-ions across membranes (Cl and SO42− ions across Neosepta CMX cation exchange membranes, K+ and Mg2+ ions across Neosepta AMX anion exchange membranes) in the title electrodialytic metathesis have been determined. It was established that the flow of co-ions causes contamination of all solutions flowing through the electrodialyzer. For instance, Cl ions penetrate into anolyte (K2SO4 solution), chlorine is found among anode reaction products and deteriorates operating conditions and induces corrosive processes. A quantitative contamination assessment of two products (potassium sulphate and sodium chloride) was carried out.  相似文献   

14.
The adsorption of sulphate, hydrogenchromate and dihydrogenphosphate anions on surfactant-modified clinoptilolite (SMC) was investigated. The SMCs were prepared by the adsorption of cis-1-aminoctadecen-9 (oleylamine) on both modified and unmodified natural clinoptilolite tuff. The properties of the modified clinoptilolite samples, such as cation type, structure of the zeolite framework and ECEC value, determined the mechanism of oleylamine adsorption, and consequently anion adsorption on the external clinoptilolite surface. According to the strength of the anion adsorption, two groups of SMCs could be distinguished: strong and weak anion adsorbents. Strong anion adsorbents were obtained by oleylamine adsorption on H+-clinoptilolites by protonation of the –NH2 groups. This mechanism of oleylamine adsorption resulted in the surface precipitation mechanism of anion adsorption being the dominant mechanism. The oleylamine derivatives of Ca- and Na-clinoptilolite were weak anion adsorbents. Oleylamine is adsorbed on Ca- and Na-clinoptilolite by hydrogen bonding, thus yielding insufficient adsorption sites for anions. Hydrogenchromate and dihydrogenphosphate anions were nevertheless adsorbed on these SMCs by interaction with oleylamine. The experiments of anion adsorption on various oleylamine loaded SMCs confirmed the existence of two types of anion adsorption sites and showed that excess oleylamine did not significantly influence the anion adsorption in the investigated concentration range. The kinetic results showed that SO42− and H2PO4 adsorptions were slow processes while HCrO4 adsorption was completed in a few minutes.  相似文献   

15.
Thin films of Mg2−xAlxNi alloys have been prepared by magnetron sputtering, and the effects of partial substitution of Al for Mg on the electrochemical properties of the films were studied. EIS results indicate the rate-limiting process for the thin film hydride electrode is the charge transfer reaction during the process of total discharge. A theoretical model has been derived for the impedance of a thin film hydride electrode based upon the assumption that hydrogen diffusion is neglected in the electrode. The charge-transfer reaction rate at the electrode surface and hydrogen diffusivity in the Mg2−xAlxNi thin film hydride electrodes were observed to initially decrease then increase with increasing Al content. Results from capacitance measurements indicate n-type semiconductor properties for the corrosion layer during the charge–discharge process. Hydrogen atom and OH transfer became more difficult with increasing Al content until x = 0.3, after which a significant drop in the barrier resistance was observed.  相似文献   

16.
H2S and COS adsorption were studied on two calcined layered double hydroxides (LDHs), Mg0.75Al0.25(OH)2(CO3)0.125 and Mg0.65Al0.35(OH)2(CO3)0.175, using diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) and a chemisorption apparatus. Both demonstrated the ability to irreversibly adsorb H2S, corresponding to uptakes of 1.54 and 1.76 μmol/m2, respectively, but Mg0.75Al0.25 had a significantly larger capacity for COS, 1.62 μmol/m2 compared to 0.80 μmol/m2 for Mg0.65Al0.35. Analysis of the DRIFT spectra suggests the adsorption of H2S proceeds via the substitution of lattice oxygen with sulfur, resulting in the formation of H2O on the surface. COS adsorption is more complicated, although it appears that a similar substitution of lattice oxygen with sulfur occurs. This results in the formation of CO2 and subsequently bicarbonates and carbonates. The formation of hydrogen thiocarbonate is also involved, although this form is generally only observed in the later stages of adsorption and appears to form at the expense of bicarbonate. The Mg0.75Al0.25 LDH retained its ability to adsorb COS in the presence of propene.  相似文献   

17.
In this work, a series of Fe3−xTixO4 (0 ≤ x ≤ 0.78) was synthesized using a new soft chemical method. The synthetic Fe3−xTixO4 were characterized using X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), Mössbauer spectroscopy, thermogravimetric and differential scanning calorimetry (TG–DSC) analyses. The results showed that they were spinel structures and Ti was introduced into their structures.Then, decolorization of methylene blue (MB) by Fe3−xTixO4 in the presence of H2O2 at neutral pH values was studied using UV–vis spectra, dissolved organic carbon (DOC) and element C analyses. Furthermore, the degradation products remained in reaction solution after the decolorization were identified using ionic chromatography (IC), 13C nuclear magnetic resonance spectra (NMR), liquid chromatography and mass spectrometry (LC–MS). Although small amounts of MB were mineralized, the aromatic rings in MB were destroyed completely after the decolorization. Decolorization of MB by Fe3−xTixO4 in the presence of H2O2 was promoted remarkably with the increase of Ti content in Fe3−xTixO4 due to the enhancement of both adsorption and degradation of MB on Fe3−xTixO4.  相似文献   

18.
Mg–Al hydrotalcite-like materials (Mg(1−x)Alx(OH)2(CO3)x/2·nH2O with x=0.33, 0.25 and 0.20) were used as supports for the immobilization of Fe(III) tetrasulfonated phthalocyanine (FePcTs). Electron paramagnetic resonance (EPR) and X-ray absorption (XANES) spectroscopies were used to characterize the solids after FePcTs adsorption on the hydrotalcite-like materials (HTs). An adsorption study was carried out by monitoring in situ the FePcTs UV/visible electronic spectra during its addition to suspensions of HTs. The results showed that the HT composition controls the nature of adsorbed species: for HTs with higher Al3+ contents (x=0.33 and 0.25), the FePcTs was adsorbed mainly in the μ-oxo complex form whereas for HT with x=0.20, non-oxo-bridged FePcTs dimeric species prevailed. The heterogeneous catalytic studies of the HT-FePcTs materials in the oxidation of catechol, using hydrogen peroxide as oxidant, showed an enhanced catalytic activity and longevity, compared to the homogeneous counterpart. The catechol conversion was 63%, 73% and 88% for the materials containing HTs with x=0.20, 0.25 and 0.33, respectively. Therefore, the reactivity of HT-FePcTs materials was improved when the Al3+ content in the HTs increases. These catalytic tests associated to the adsorption studies showed that the μ-oxo complex of the FePcTs, mainly adsorbed on the HTs with x=0.25 and 0.33, seems to be the active species in catechol oxidation. These findings suggest that a cooperative effect take place in the HT-FePcTs materials, showing that HTs do not act as an inert support.  相似文献   

19.
Cation exchange mechanism and rate of Cs+ exchange were investigated in < 2 μm and 20–2 μm particle size fractions of K-depleted phlogopite (Na-phlogopite). The K-depleted phlogopite was prepared from a natural phlogopite by a potassium removal method using sodium tetraphenylborate (NaTPB) at room temperature. X-ray diffraction (XRD) patterns revealed that interlayer K+ ions were completely replaced with sodium ions after the potassium removal treatment. Ion exchange isotherms and kinetics were determined for Na+ → Cs+ exchange with two particle size fractions. The isotherms indicated that both particle size fractions showed high selectivity for Cs+. Based on the isotherm tests, ΔGo values of < 2 μm and 20–2 μm particle fractions were − 6.83 kJ/mol and − 7.08 kJ/mol, respectively. Kinetics of Cs exchange revealed that the 20–2 μm particle size fraction of the K-depleted phlogopite took up more Cs+ ions than the < 2 μm particle size fraction. Various kinetic models were applied to describe Na+ → Cs+ exchange process. Elovich model described the kinetic data of the < 2 μm particle size fraction well, while the modified first-order model or parabolic diffusion model described the data of the 20–2 μm particle size fraction well.  相似文献   

20.
Effects of electrolyte addition on photocatalytic activity of (Ga1−xZnx)(N1−xOx) modified with either Rh2−yCryO3 or RuO2 nanoparticles as cocatalysts for overall water splitting under visible light (λ > 400 nm) are investigated. The cocatalyst Rh2−yCryO3 is confirmed to selectively promote the photoreduction of H+, while RuO2 functions as both H2 evolution sites and as efficient O2 evolution sites. The activity of Rh2−yCryO3-loaded (Ga1−xZnx)(N1−xOx) is found to be suppressed in the presence of Cl, which undergoes oxidation by photogenerated holes in the valence band of (Ga1−xZnx)(N1−xOx). Alkaline- and alkaline earth-metal cations in the reactant solution compensate the negative effect of Cl to a certain extent depending on the metal cation employed. Among the electrolytes examined, the addition of an appropriate amount of NaCl or A2SO4 (A = Li, Na, or K) to the reactant solution without pH control is found to increase activity by up to 75% compared to the case without additives. Direct splitting of seawater to produce H2 and O2 is also demonstrated using Rh2−yCryO3-loaded (Ga1−xZnx)(N1−xOx) catalyst under visible light.  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