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 62 毫秒
1.
PHTS materials with combined micro- and mesoporosity and with different morphologies were synthesized. PHTS or Plugged Hexagonal Templated Silica is a plugged variant of SBA-15 containing extra silica nanoparticles (plugs) inside the mesoporous channels. Utilising a low amount of TEOS and high stirring and aging temperatures, PHTS-particles with spherical morphology were synthesized for the first time. It was observed that the morphology changed gradually with increasing stirring temperature. It evolved from smooth rods (60C) to rough rods with a deposition of small particles on the surface (70C) and finally to spheres (80C). However, synthesis of materials at a stirring and aging temperature of 80C using larger amounts of silica source (TEOS) did not result in particles with spherical morphology. This demonstrates the profound impact of the amount of silica source, as well as the temperature. The results are explained by means of the cloud point of the surfactant and the balance between the rate of polymerization of the silica source and the rate of mesostructure formation. The prepared PHTS materials were characterized by XRD, N2 sorption and SEM.  相似文献   

2.
Copper nucleation and growth on titanium substrates from concentrated acidic copper sulfate solutions was studied at 45 and 65C and high cathodic potentials. Electrochemical experiments allied to SEM examination were performed to characterize the mechanism of nucleation and its evolution with time. Particular emphasis was given to the influence of potential and agitation on the initial stages of nucleation and growth. Results indicated that most of copper nucleation on titanium from a 83g dm–3 Cu2+ solution, at 65 C, is achieved in a matter of milliseconds. As expected, the initial stages of the copper nucleation and growth is strongly dependent of potential and temperature, and the influence of agitation is only evident at very high potentials. The calculated diffusion coefficients for Cu2+ at 45 and 65 C, under the experimental conditions, were found to be 9.16 × 10–6 and 1.62 × 10–5cm2 s–1, respectively.  相似文献   

3.
Large, high quality, perfect hexagonally shaped AlPO4-5 crystals have been crystallized using hydrothermal synthesis. The crystallization was carried out at 447 or 457 K with a crystallization time from one to four days and tripropylamine (TPA) molecules as template. The morphology of the crystals was observed with SEM. For the first time, the structure of the AlPO4-5 crystal together with TPA template molecules has been successfully refined in space group P6cc by single crystal X-ray diffraction. The parameters are a = b = 13.725(3) Å, c = 8.473(3) Å, and = 120.0. The size of some crystals is up to 2.2 mm in c direction, or up to 0.31 mm in a or b direction. All observed angles and distances are within acceptable ranges: P–O = 1.50–1.54 Å, Al–O = 1.62–1.72 Å, O–P–O = 107–111, O–Al–O = 103–113 and P–O–Al = 147–150.  相似文献   

4.
The corrosion inhibition of austenitic chromium–nickel steel by two Schiff bases, N-(1-toluidine)salicylaldimine and N-(2-hydroxyphenyl)salicylaldimine, was investigated in sulphuric acid medium. The effect of concentration and temperature on inhibition properties was determined. It was found that when the concentrations of inhibitor were increased the inhibition efficiencies () and surface coverage () increased. Some thermodynamic parameters such as free energy of adsorption, G ads, and enthalpy, H, were determined for the Schiff bases. Experimental results agree with the Temkin isotherm for N-(1-toluidine)salicylaldimine, but the Langmuir isotherm is more appropriate for N-(2-hydroxyphenyl)salicylaldimine.  相似文献   

5.
The electrodeposition of -nickel hydroxide is promoted by the simultaneous chemical corrosion of the electrode by an acidic nitrate bath. Chemical corrosion results in the formation of a poorly ordered layered phase which is structurally similar to -nickel hydroxide and provides nucleation sites for the deposition of the latter. Therefore under conditions which enhance corrosion rates such as low current density (<1.3 mA cm–2), high temperature (60 C), high nickel nitrate concentration ( 1M) and the resultant low pH (1.7), -nickel hydroxide electrodeposition is observed, while -nickel hydroxide forms under other conditions. Further, -nickel hydroxide deposition is more facile on an iron electrode compared to nickel or platinum.  相似文献   

6.
Niobium deposits were prepared from alkali chloride melts on nickel and AISI316 stainless steel substrates both by constant current and by pulse current methods. The influence of electrolysis conditions on the nature, morphology and purity of the deposits was investigated by X-ray diffraction, scanning electron microscopy and energy dispersive X-ray analysis. No metallic niobium was obtained at temperatures below 500 C. At temperatures between 550 and 650 C, the deposits were dendritic and non-adherent, whereas pure niobium layers could be obtained at 750 C. Detailed analysis showed that a large negative overpotential during the pulse current period lead to the presence of suboxides, such as Nb6O, in the metallic phase. Suitable electrolysis conditions gave pure oxygen-free niobium. Cross section analysis showed that on nickel a thin layer of niobium–nickel alloy such as NbNi3, was formed at the metal interface. In contrast no alloys were detected at the niobium-stainless steel interface, where homogenous adherent layers of thickness around 50 m were obtained.  相似文献   

7.
The structure, chemistry and morphology of commercially available carbon-supported and unsupported Pt–Ru catalysts are investigated by X-ray diffraction, energy-dispersive analysis by X-rays and electron microscopy. The catalytic activities of these materials towards electrooxidation of methanol in solid-polymer-electrolyte direct methanol fuel cells have been investigated at 90C and 130C with varying amounts of Nafion ionomer in the catalytic layer. The unsupported Pt–Ru catalyst exhibits higher performance with lower activation-control and mass-polarization losses in relation to the carbon-supported catalyst.On leave from the  相似文献   

8.
Corrosion rate data for several commercial carbon dioxide corrosion inhibitors have been fitted to the Temkin adsorption isotherm and van't Hoff equation, enabling a determination of the enthalpy of adsorption (H ad ). It has been found that experimentally determined H ad values for commercial inhibitor formulations can provide valuable insights into the behaviour of the inherent active components. The temperature sensitivities along with the mechanism of adsorption (physi- or chemisorption) and the minimum concentration required for activity of inhibitors is determinable by this approach. Additionally, FTIR has been used to identify those inhibitors that are tenaciously adsorbed to steel, and suitable for use in batch treatment applications.  相似文献   

9.
The growth kinetics of electrogenerated hydrogen, oxygen and chlorine gas bubbles formed at microelectrodes, were determined photographically and fitted by regression analysis to the equation;r(t)=t x , wherer(t) is the bubble radius at timet after nucleation, the growth coefficient, andx the time coefficient. The coefficientx was found to decrease from a short time (< 10 ms) value near unity, typical of inertia controlled growth, through 0.5, characteristic of diffusional control, to 0.3, expected for Faradaic growth, at long times (\s> 100 ms). The current efficiency for bubble growth increased with bubble lifetime, reflecting the decrease in local dissolved gas supersaturation. The pH dependency of the bubble departure diameter indicated that, in surfactant-free electrolytes, double layer interaction forces between the negatively charged hydrogen evolving cathode or positively charged oxygen/chlorine evolving anode and positively (pH \s< 2) or negatively (pH \s> 3) charged bubbles, were the determining factor. The effect of addition of an increasing concentration of cationic (DoTAB) or anionic (SDoS) surfactant was to progressively reduce the pH effect on departure diameter, due to surfactant adsorption on the bubble and, to a lesser extent, on the electrode.Nomenclature C coefficient [3] - D diffusion coefficient (m2 s–1) - I current (A) - P pressure (kN m–2) - R universal gas constant (8.314 J mol–1 K–1) - r bubble radius (m) - T absolute temperature (K) - t time (ms) - x time coefficient - zF molar charge (96 487z C mol–1) - growth coefficient (m s–0.33) - P Laplace excess pressure (kN m–2) - surface tension (mN m–1) - electrolyte density (kg m–3) - contact angle () Paper presented at the International Meeting on Electrolytic Bubbles organized by the Electrochemical Technology Group of the Society of Chemical Industry, and held at Imperial College, London, 13–14 September 1984.  相似文献   

10.
H. He  H.X. Dai  K.Y. Ngan  C.T. Au 《Catalysis Letters》2001,71(3-4):147-153
The physico-chemical properties of passivated -Mo2N have been investigated. The material showed high activities for NO direct decomposition: nearly 100% NO conversion and 95% N2 selectivity were achieved at 450C. The amount of O2 taken up by -Mo2N increased with temperature rise and reached 3133.9 molg–1 at 450C; we conclude that there formation of Mo2OxNy occurred. This oxygen-saturated -Mo2N material was catalytically active: NO conversion and N2 selectivity were 89 and 92% at 450C. We found that by means of H2 reduction at 450C, Mo2OxNy could be reduced back to -Mo2N and the oxidation/reduction cycle is repeatable; such a behaviour and the high oxygen capacity (3133.9 molg–1) of -Mo2N suggest that -Mo2N is a promising catalytic material for automobile exhaust purification.  相似文献   

11.
A lithium–manganese oxide, Li x MnO2 (x=0.30.6), has been synthesized by heating a mixture (Li/Mn ratio=0.30.8) of electrolytic manganese dioxide (EMD) and LiNO3 in air at moderate temperature, 260 C. The formation of the Li–Mn–O phase was confirmed by X-ray diffraction, atomic absorption and electrochemical measurements. Electrochemical properties of the Li–Mn–O were examined in LiClO4-propylene carbonate electrolyte solution. About 0.3 Li in Li x MnO2 (x=0.30.6) was removed on initial charging, resulting in characteristic two discharge plateaus around 3.5V and 2.8V vs Li/Li+. The Li x MnO2 synthesized by heating at Li/Mn ratio=0.5 demonstrated higher discharge capacity, about 250mAh (g of oxide)–1 initially, and better cyclability as a positive electrode for lithium secondary battery use as compared to EMD.  相似文献   

12.
Summary In dependence on crystallization conditions three ranges with different crystal structure and heat of fusion were found by DSC,WAXS,and IR for unoriented PA 6.6 samples of densities between 1.10 and 1.17gcm–3: Range I:I triclinic, c I =1.225 gcm–3,H M I = 235 Jg–1. Range II:II triclinic, c II =1.165 gcm–3, H M II =185 Jg–1. Range III:Continuous variation from c I ,H M I to c II , H M II . a=1.095 gcm–3 is independent of crystallization. conditions. The transition between I and II is probably due to changes of the chain conformation.  相似文献   

13.
Summary Critical values of the polymer volume fraction 2,c and the interaction parameter c have been computed for the case that the equation for the chemical potential of solvent contains terms c 2 3 and c 2 4 in addition to 2 2 . For 0 c 1/3, the limits for infinite chain length are 2,c = 0 and c = 0.5. Quite different results are obtained for c > 1/3, 2,c being finite and c lower than 1/2. Conclusions for the estimation of the temperature and the entropy-of-dilution parameter are discussed.  相似文献   

14.
The lamellar morphology of a melt-miscible blend consisting of two crystalline constituents, poly(3-hydroxybutyrate) (PHB) and poly(ethylene oxide) (PEO) have been investigated by means of small angle X-ray scattering (SAXS). The blend was a crystalline/amorphous system when temperatures lay between the melting point of PEO (ca. T m PEO=60C) and that of PHB (ca. T m PHB=170C), while it became a crystalline/crystalline system below T m PEO. The crystalline microstructures of the blends were induced by two types of crystallization history, i.e. one-step and two-step crystallizations. In the one-step crystallization, the blends were directly quenched from the melt to room temperature to allow simultaneous PHB and PEO crystallization. The two-step crystallization involved first cooling to 70C to allow PHB crystallization for 72 h followed by cooling to room temperature (ca. 19C) to allow PEO crystallization. In the crystalline/crystalline state, two scattering peaks have been observed in the Lorentz-corrected SAXS profiles, irrespective of the crystallization histories, meaning that crystallization created separate PHB and PEO lamellar stack domains. One-step crystallization yielded lamellar stack domains containing almost pure PHB and PEO lamellae. Two-step crystallization generated almost pure PHB lamellar domains and the PEO lamellar domains with inserted PHB lamellae. In the crystalline/amorphous state, the composition dependence of the amorphous layer thickness (l a), the presence of zero-angle scattering, and the volume fraction of the PHB lamellar stack (s) revealed that both one-step and two-step crystallizations, generated the interfibrillar segregation morphology, where the extent of interfibrillar segregation of amorphous PEO increased with increasing PEO content.  相似文献   

15.
Summary By using -butyrolactone (-BL) as the reaction media, highly active catalysts--light rare earth chloride-epoxidy---BL-for the solution polymerization of -caprolactone, have been obtained for the first time. With these catalyst, PCL with molecular weight as his as 40x104(Mv) can be prepared at 60°C for 1.5 hr. The amount of epoxide in catalyst solution, catalyst aging temperature and time affect the catalyst activity significantly. The mechanism study shows that in -BL, the weakening of Ln-Cl bonds by the donation of coordinated -BL with Ln3+ and the homogenous effect promote the reaction between light rare earth chloride and epoxide. The produced rare earth alkoxide initiates CL polymerization via a coordination-insertion mechanism with Acyl-oxygen bond cleavage.  相似文献   

16.
Summary The interaction of -cyclodextrin(-CD) with sodium 1-pyrenesulfonate(PS) was studied spectrophotometrically. -CD was found to cause much larger decrease in the absorption maxima of PS than -CD. The fluorescence spectra of PS in the presence of -CD showed excimer emission, while those of PS with -CD showed only monomer emission, indicating that -CD forms 12 (-CDPS) complexes in which two PS molecules are included in the -CD cavity in a face-to-face fashion. The binding isotherm showed a sigmoidal curve. The association constants were estimated by computer simulation of the binding curve. The 12 (CDPS) complex was found to be much more stable (K=106 M–1) than the 11 complex (K=1 M–1). At high concentration of -CD another -CD cooperates in binding two PS molecules, resulting in the formation of a 22 complex.  相似文献   

17.
CO interacts with extraframework alkali metal cations (M+=) of zeolites to form both M+CO and M+OC species. By using variabletemperature FTIR spectroscopy, these Cbonded and Obonded species were found to be in a temperaturedependent equilibrium. For the same cation, the difference in interaction energy depends upon the zeolite framework. Thus, for the equilibrium process ZNa+=CO ZNa+OC, where Z represents the zeolite framework, H 0 was found to take the values 3.8 and 2.4 kJ mol for CO/NaZSM5 and CO/NaY, respectively. The Cbonded species show always the highest cation–CO interaction energy.  相似文献   

18.
The effect of heat treatment of Ti and Ti–0.2 Pd alloys on their anodic oxidation was studied in deaerated 1% NaCl by means of anodic linear sweep voltammetry, SEM, TEM, EDS, optical microscopy and microhardness measurements. The specimens, as fabricated, consisted of -phase only. The -phase, intergranular or with a Widmansttäten type growth, was produced by heat treatment of the Ti–0.2 Pd alloys at the temperature range from 750 to 850 C. The -phase was transformed into the -phase during quenching. The current density against voltage curves for pure Ti and Ti–0.2 Pd, as fabricated or heat-treated, presented an initial plateau at about 1.5 V vs Ag/AgCl/KCl (3 M), an anodic peak at about 4.5 V and a current increase due to the pitting attack at about 10 V. The anodic peak was related to an oxide growth together with a solution electrolysis. Current spikes appeared at random from potentials about 8.3 V, which were related to film breakdown and repair events. The passive films of the alloys oxidized up to about 10 V presented oxidation bands parallel to the surface, with different oxygen content and microhardness, together with a structural transformation of the -phase under the titanium oxide layer. The similar behaviour of pure Ti and Ti–0.2 Pd alloys in front of pitting corrosion in chloride was due to such a structural transformation.  相似文献   

19.
The present investigation concerns the synthesis of statistical networks similar to vulcanized rubber. We have used liquid precursor polymers: polybutadiene (PB) and polyisoprene (PI) with telechelic siloxanes. The reactivity of the polydienes can be classified as follows: PB units 1,2 PI units 3,4 PI units 1,4 0. The network formation is a function of the ratio R = nb, of pendent double bonds/nb. of SiH. For R 1 the networks are rather hard and brittle, for 8 R 10 they are soft and elastic, whereas for R 16 the gels obtained are very soft and sticky.  相似文献   

20.
Thermal transition of PVA-borax aqueous gels with a PVA concentration of 60 g/L and a borax concentration of 0.28 M was investigated at temperatures ranging from 15 to 60C using static light scattering (SLS), dynamic light scattering (DLS), and dynamic viscoelasticity measurements. Three relaxation modes, i.e. two fast and one slow relaxation modes, were observed from DLS measurements. Two fast relaxation modes located around 10–3101 sec, with one fast mode (f1) being scattering vector q-dependent and the other fast mode (f2, with f2>f1) being q-independent. The f1 mode was attributed to the gel mode whilst the f2 mode could be due to the hydrodynamics of intra-molecular hydrophobic domains formed by uncharged segments of polymer backbones. The slow relaxation mode with relaxation time located around 101103 sec in DLS data was due to the motion of aggregated clusters and was observed only at temperatures above 40C. The amplitude and relaxation time of slow mode decrease as temperature is increased from 40 to 60C. At temperatures below 40C, no slow relaxation mode was observed. The SLS measurements showed PVA-borax-water system had fractal dimensions D f2.4 and D f2.0 as temperature was below and above 40C, respectively. The simple tilting test indicated gel behaviour for the PVA-borax aqueous system at temperatures below 40C with a creep flow after a long time exposure in the gravity field. But the dynamic viscoelasticity measurements demonstrated a solution behaviour for PVA/borax/water at temperatures below 40C, the critical gel point behaviour for G() and G() was not observed in this system as those reported for chemical crosslinked gels. These results suggest that the PVA-borax aqueous system is a thermoreversible weak gel.  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