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.
Two polymeric carboxylato-bridged manganese(II) complexes, $ {}_{\infty }^{1} $ ∞ 1 [Mn(bpe)(NBA)2] (1) and $ {}_{\infty }^{1} $ ∞ 1 [Mn(bpe)(MBA)2] (2) (bpe = 1,2-di(4-pyridyl)ethylene; HNBA = m-nitrobenzoic acid; HMBA = m-methyl-benzoic acid), were synthesized and characterized. Compounds 1 and 2 are isostructural. X-ray diffraction studies show that the title compounds possess a double chain structure. The chains in 1 and 2 are assembled into 2-D layers via C–H···O hydrogen bonds interactions. Furthermore, owing to the offset face-to-face π–π stacking interactions, the adjacent 2-D layers embed into each other to form a 3-D supramolecular framework. Variable-temperature (2–300 K) magnetic susceptibility measurements show the presence of weak antiferromagnetic interactions between the high-spin Mn(II) (S = 5/2) ions through a carboxylato bridge with the best fit parameters for 1 being J = ?0.13 cm?1, zJ′ = ?0.098 cm?1 and for 2, J = ?0.33 cm?1, zJ′ = ?0.001 cm?1.  相似文献   

2.
Both (R)-2-Chloromandelato coordination polymers, [Cu(Clma)2] n 1 and [Mn(H2O)(Clma)2] n 2, crystallizing in the polar chiral monoclinic space group P21 represent the first two examples of metal complexes with (R)-2-Chloromandelic acid (HClma) ligand. Through the (R)-2-Chloromandelate Clma? ions, the penta- and hexa-coordinated metal atoms in 1 and 2 are linked into chains, 1 {[Cu(Clma)](Clma)2/2} n and 1 {[Mn(H2O)(Clma)](Clma)2/2} n , respectively, and further assembled into 3D supramolecular architectures via hydrogen bonding and close packing interactions. They exhibit ferroelectricity (Pr = 0.03 and 0.01 μC cm?2, Ec = 5.09 and 0.52 kV cm?1, Ps = 0.10 and 0.17 μC cm?2, respectively). The magnetic behaviors of 1 and 2 can be interpreted by means of a 1D chain model, where the magnetic exchanges are transmitted via carboxylate groups between metal atoms, and the best fit results in J = 2.97, ?0.15 cm?1 for 1 and 2, suggesting ferromagnetic and antiferromagnetic interactions, respectively.  相似文献   

3.
A single crystal of excessively Pb2+-exchanged zeolite Y (|Pb 15.5 2+ (Pb4O4(Pb 16/19 2+ Pb 3/19 4+ )4)4.75|[Si117Al75O384]-FAU) was prepared by exchange of Na–Y (|Na75|[Si117Al75O384]-FAU, Si/Al = 1.56) with an aqueous stream 0.05 M Pb(C2H3O2)2 at 294 K, followed by vacuum dehydration at 1 × 10?6 Torr and 693 K. Its structure was determined at 100 K, by X-ray diffraction techniques in the cubic space group Fd $ \overline{3} $ 3 ¯ m and was refined to the final error indices R 1/wR 2 = 0.0639/0.1323. About 53.5 Pbn+ ions per unit cell occupy three different equipoints; 26 are at site I′, 19 are at site II, and the remaining 8.5 are at another site II. Three Pb4+ ions at some of the positions must have higher oxidation states due to elevated dehydration temperature; Pb(IIa) is supposed to coexist with Pb2+ and Pb4+ ions assuming the charge balance of the zeolite framework. A distorted Pb4O4 cube, alternating Pb2+ at Pb(I′) and O2? at O(5), coordinates with four Pb2+ and/or Pb4+ ions through its oxygen atoms to give a [Pb 4 2+ O 4 2? (Pb 16/19 2+ Pb 3/19 4+ )4]176/19+ cluster in 4.75 of eight sodalite cavities per unit cell in zeolite Y.  相似文献   

4.
The formation dynamics of calcium barium sulphoaluminate mineral with the composition of 2.75CaO·1.25BaO·3Al2O3·SO3 (C2.75B1.25A3 $\overline{\text{S}}$ S ¯ ) was studied. The results suggest that, under the preparative conditions, the formation of C2.75B1.25A3 $\overline{\text{S}}$ S ¯ mineral is controlled by a diffusion mechanism from 1,100 to 1,380 °C; and, the formation dynamics fits nicely with D 4 = 1 ? 2α/3 ? (1 ? α)2/3 = Kt. From 1,100 to 1,300 °C, the apparent activation energy is 227.45 kJ mol?1. From 1,300 to 1,380 °C, the apparent activation energy decreases to 175.94 kJ mol?1, making the formation of C2.75B1.25A3 $\overline{\text{S}}$ S ¯ mineral faster and easier.  相似文献   

5.
Methyl bromide was synthesized by reacting methane with oxygen and hydrogen bromide over Rh/SiO2 catalyst. The reaction started from the oxidation of HBr to form active bromine species (Br? radicals and Br* surface species), which in turn reacted with CH4 to form CH 3 ? radicals and $\hbox{CH}_{3}^{\ast}$ surface species. These CH 3 ? and $\hbox{CH}_{3}^{\ast}$ species reacted with the active bromine species to form CH3Br and CH2Br2. The presence of HBr inhibited the deep oxidation and the steam reformation of CH4 and therefore, guaranteed the high selectivity of CH3Br. In the presence of HBr, CO was formed from the oxidation and steam reformation of CH3Br, while CO2 was formed from the oxidation and steam reformation of CO over Rh/SiO2 at reaction temperature higher than 560 °C.  相似文献   

6.
The M1 phase of the MoV(Nb,Ta)TeO system is one of the most effective catalysts for the ammoxidation and selective oxidation of propane to acrylonitrile (AN) and acrylic acid, respectively. The active centers of the M1 phase reside on the ab planes of this crystalline material (i.e., the (001) lattice face). Early on we proposed that the thus located active centers contain all key catalytic elements strategically placed for the conversion of propane to AN. These seven element comprising active centers contain: five metal oxide octahedra (2 V 0.32 5+ /Mo 0.68 6+ , 1 V 0.62 4+ /Mo 0.38 6+ , 2Mo 0.5 6+ /Mo 0.5 5+ ) and two Te4+??oxygen sites. In this contribution we analyze the various compositional probabilities of the seven element active centers and their additional eight element surroundings and conclude that there are 32 possible compositional arrangements of this 15 element assembly. From the diverse structural arrangements, diverse catalytic properties can be assigned to the individual sites, leading to diverse propane reaction pathways. We conclude that there are 22% AN forming, 22% propylene, 10% waste and 46% inert sites. After normalization these sites are deemed to lead to the following product yields: 41% AN, 41% propylene and 18% waste. The highest experimentally attained AN yield from propane is 42%, employing M1 phase only, which coincides with the predicted value of a concerted mechanism. Higher AN yields are, however, anticipated, up to a lofty upper limit of 82%, by allowing also for a consecutive mechanism (C3° ?? C 3 = ??AN). This possibility can be rationalized on the basis of the existence of vicinal C3° ?? C 3 = /C 3 = ??AN sites whose presence is plentiful on the catalytically important ab planes of M1. The placement and efficiency of these sites is, however, not perfect; therefore the upper AN yield limit is not realized in practice. Our analysis of the elemental distribution at the active centers and their immediate surroundings provides us with new insights into the relationship between structure and catalytic reaction mechanisms of the M1 phase and might serve as a guide towards a redesign of the M1 composition, so as to attain higher AN yields from propane. It provides a challenging task for the synthetic chemist.  相似文献   

7.
8.
Electrochemical reduction of AlCl3 dissolved in acidic AlCl3-n-butyl-pyridinium chloride melt was studied by linear sweep voltammetry and chronopotentiometry at tungsten and platinum electrodes, in the Al2Cl 7 ? concentration range 0.3 to 0.5 M between 30 and 60°C. Al2Cl 7 ? bulk reduction was preceded by a nucleation (tungsten) or alloy formation phenomenon (platinum). The overall results agree rather well with the mechanism: $$\begin{gathered} 2AlCl_4^ - \rightleftarrows Al_2 Cl_7^ - + Cl^ - \hfill \\ 4Al_2 Cl_7^ - + 3e \rightleftarrows Al + 7AlCl_4^ - \hfill \\ \end{gathered} $$ The electrochemical reaction appeared quasi-reversible. Calculated values of the product of the transfer coefficient by the number of the electron exchanged in the rate determining step were in the range 0.45 to 0.7. Diffusion coefficients for Al2Cl 7 ? were calculated.  相似文献   

9.
Kinetic analysis of the Li|Li+ interphase in an electrolyte based on N-metyl-N-propylpyrrolidinium bis(trifluoromethanesulfon)imide ionic liquid (MPPyrrTFSI) and lithium bis(trifluoromethanesulfon)imide salt (LiTFSI) was performed. Li|electrolyte|Li and LiC6|electrolyte|Li cells were galvanostatically charged/discharged in order to form solid electrolyte interphase (SEI) protecting layer. SEM images showed that the surface of both Li and LiC6 anodes was covered with small particles. The fitting procedure of electrochemical impedance data taken at different temperatures gave three resistances (R el, R SEI, R ct) and hence, three lnR = f(T ?1) straight lines of different slopes. Specific conductivity and activation energy of the conduction process of the liquid electrolyte, were ca. σ = 2.5 mS cm?1 (at T = 25.0 °C) and $ E_{\text{el}}^{\# } $  = 15 kJ mol?1. Activation energy for the conduction process in the SEI layer was ca. 56 kJ mol?1 in the case of the metallic lithium and 62 kJ mol?1 for the graphite anode. Activation energy of the charge transfer process, $ E_{\text{ct}}^{\# } $ , for Li and LiC6 anodes was 71 and 65 kJ mol?1, respectively. Analysis of literature data for different electrolytes suggests that the $ E_{\text{ct}}^{\# } $ value for Li+ reduction may be approximated by 57 ± 5 kJ mol?1. Activation energy for the diffusion processes in the graphite electrode, detected from the Warburg coefficient, was ca 74 kJ mol?1.  相似文献   

10.
Small water soluble Ru-nanoparticles (ca. 2–5 nm) stabilized by lignins were synthesized by reduction of RuCl3 using H2. For comparison purposes, small Ru-NPs (ca. 2.0 nm) with narrow size distribution were also synthesized using polyvinylpyrrolidone (PVP) as stabilizer and H2 and NaBH4 as reducing agents. All these Ru-NPs were active catalysts in Fischer–Tropsch reaction. Interestingly, CO2 was detected as by-product demonstrating that the water gas shift reaction is taking place under these conditions. The Ru-NPs stabilized by lignins were less active (up to 0.49 molCO mol Ru ?1  h?1) that those stabilized by PVP (up to 3.35 molCO mol Ru ?1  h?1), exhibiting also higher CO2 production. Several reaction parameters were optimized such as the stirring rate, reducing method, polymer/Ru ratio and size of the Ru-NPs.  相似文献   

11.
The metal iodides reduce partially the host coordination polymer of the type $ ^{ 3}_{\infty } \left[ {\left( {{\text{Me}}_{ 3} {\text{Sn}}} \right)_{ 3} {\text{Fe}}\left( {\text{CN}} \right)_{ 6} } \right] $ , I, to give new host–guest supramolecular coordination polymers (SCP). The physical and chemical characteristics of the new products were studied by elemental analyses, X-ray powder diffraction, IR, UV/Vis, and solid state NMR spectra. The host–guest SCP are [Mx(Me3Sn)3Fe(1–x)IIIFe x II (CN)6]n M = Li+·2H2O, 1; Li+, 2; Na+, 3; K+, 4; Cu+, 5, [Li(Me3Sn)3FeII(CN)6]n, 6 and [(LiDEE)0.9(Me3Sn)3Fe o.1 III Fe o.9 II (CN)6]n, 7. The stoichiometry and nature of the guest depend on the type of the metal iodide and the reaction conditions. The polymeric nature of these SCP is due to the presence of trigonal bipyramidal configured structure which bridges between the single d-transition metal ions. The host–guest SCP containing the Li ions have been tested as electrodes to construct four different lithium-ion batteries.  相似文献   

12.
Denitrification strongly depends on the availability of carbon source in constructed wetlands (CWs). In this study, several relevant carbon source extracting solutions made from hydrolyzate of selected wetland litters were added to CWs for nitrogen removal enhancement. The feasibility of supplying a carbon extracting solution to improve potential denitrification rate in horizontal subsurface flow constructed wetland was deeply investigated. Combinations of different hydraulic retention time (HRT, especially for 2-day and 4-day) with different influent COD/N ratios were designed to prove the enhancement on denitrification by carbon source supplement. In addition, specific denitrification rate (SDNR) was calculated for the comparison of the nitrogen removal at different COD/N ratios. The sequential operation results on total nitrogen (TN) and nitrate (NO 3 ? -N) removal efficiencies were obtained in CW system with an influent COD/N ratio of 4.0. The accumulation of nitrite (NO 2 ? -N) was found to be closely related to the removal of NO 3 ? -N. Meanwhile, no obvious accumulation of NO 2 ? -N was found when the removal of NO 3 ? -N was relatively high.  相似文献   

13.
WO3/TiO2 composite film was prepared by microarc oxidation technique and characterized by SEM, XRD, UV-vis spectra and Zeta-potential. The photocatalytic activity of WO3/TiO2 composite film was evaluated by examining the degradation of methyl orange. The influence of solution pH and inorganic anions on removal ratio of methyl orange was investigated. Removal ratio of methyl orange decreased with an increase of pH value in acidic solution, while it increased with the pH value in alkaline solution. The influence of added anions on the removal ratio is divided into two aspects. Addition of Cl? and SO 4 2? resulted in a decrease in photocatalytic removal ratio of methyl orange, while it was facilitated by PO 4 3? , HCO 3 ? and NO 3 ? .  相似文献   

14.
Ring opening polymerization (ROP) of 1,3,5-tri-n-hexyl,1,3,5-trimethylcyclotrisiloxane (D 3 Hexa ) and 1,3,5-tri-n-heptyl,1,3,5-trimethylcyclotrisiloxane (D 3 Hepta ) was promoted by acid-treated synthetic silica–alumina to obtain Gaussian homo asymmetric polysiloxanes. Mw was above 70?kg/mol, meaning that homo asymmetric bulky side-group polysiloxane chains with high molecular weight were obtained. The material was treated in an acidic medium to improve the contents of acid sites and successfully tested as an inorganic acidic catalyst for ROP of D 3 Hexa and D 3 Hepta cyclosiloxanes. The samples of poly(methylhexylsiloxane) (PMHS) and poly(methylheptylsiloxane) (PMHepS) obtained were structurally characterized mainly by 29Si NMR. All the experimental values including the refractive index increment (dn/dc), the second virial coefficient (A2), the square root of the mean square radius of gyration ( $ \langle {{\text{RMS}}_{\text{radius}}}^{ 2} \rangle^{ 1/ 2} $ ), the average molecular weight (Mw), the average molecular numeral (Mn), and the weight polydispersity (Mw/Mn) were obtained using a gel permeation chromatography/light scattering (GPC/LS) coupled system. The A2 experimental value for the two polymers (between 4 and 6.5?×?10?4?mol/mL?g2) indicated that toluene was a good solvent. In addition, PMHS and PMHepS $ \langle {{\text{RMS}}_{\text{radius}}}^{ 2} \rangle^{ 1/ 2} $ were greater than 30?nm, indicating that larger chains of high molecular weight were obtained.  相似文献   

15.
The photocatalytic mineralization of oxalic acid over SiO2-based materials was investigated in the 200–800 nm range. The photocatalytic activity was found to be strongly related to the morphology of SiO2 materials. The simple as well as the Pt-modified SiO2 particles having a predominant spherical shape exhibited null photocatalytic activity. In contrast, the tubular shaped SiO2 particles revealed an interesting photocatalytic activity, the rate of CO2 evolvement being 45 µmol g cat ?1  h?1. The initial activity was significantly enhanced (428 µmol CO2 g cat ?1  h?1) by platinum photodeposition on the outer and inner surface of tubular SiO2. The catalytic materials were characterized by TEM, UV–VIS and XPS to obtain rational explanations for the phocatalytic activity that was noticed. The experiments revealed that SiO2 tubes behave as efficient photooxidation microreactors. The morphology-dependent photocatalysis can be an efficient tool in future for the abatement of pollutants in liquid phase.  相似文献   

16.
The reaction of K3[Cu(CN)4], Me3SnCl and 4-methylpyrimidine (mpym) at room temperature affords the 3D-octameric 3 [Cu8(CN)8(mpym)4], I, in water/acetonitrile and the 3D-host–guest 3 [Cu8(CN)8(mpym)4 dioxane], II. The X-ray single crystal diffraction of I reveals the formation of homometallic octameric building blocks consisting of three fused 11 membered rings. The 3D-network structure of the supramolecular coordination polymer (SCP), I, consists of two crossing sets of parallel corrugated interpenetrating CuCN chains, which are connected by mpym, hydrogen bonds, π–π stacking and cuprophilic interactions. Species II is isostructural with I as indicated by X-ray powder diffraction and spectroscopic measurements.  相似文献   

17.
Sulfated mixed oxides, SO 4 = /Ni–Al–O and SO 4 = /Zr–Al–O were evaluated for double bond isomerization (DBI) of 1-hexene using helium and hydrogen as carrier gases. The increase of temperature from 100 to 200 °C seems to favor the deprotonation pathway and contribute to increase the 1-hexene conversion for both catalysts and without regard of the carrier gas. The results indicate that temperature it is the main factor that contributes to improve both conversion and selectivity towards (cis + trans)-2-hexene, while the reductive atmosphere beneficiate only the SO 4 = /Ni–Al–O catalyst performance, as hydrogen prevents this catalyst from a fast deactivation.  相似文献   

18.
Synthetic hydroxylapatites are prepared with additives, such as Mg2+, CO 3 2? , and C2O 4 2? . An increase in the concentration of magnesium leads to the formation of struvite. In the Ca(NO3)2-(NH4)HPO4-Na2CO3-NH4OH-H2O system, an excess of carbonate ions leads to the formation of calcite. When the synthesis is performed using oxalate ions as additives, calcium oxalate does not form the inherent phase. Calcium oxalate monohydrate is synthesized with additives, such as CO 3 2? , HPO 4 2? , and SO 4 2? ions and urea, glycine, and glutamic acid. X-ray powder diffraction analysis has revealed that the composition of the CaC2O4 · H2O precipitate remains unchanged under these conditions and in the presence of the aforementioned additives.  相似文献   

19.
Photocatalytic degradation of ammonia on supported TiO2 nanoparticles was investigated. The TiO2 nanoparticles used as photocatalyst were coated on light expanded clay aggregate granules (LECA), which is a porous and light weight support. Photocatalytic reaction activity of prepared catalyst was determined by ammonia degradation from water synthetically polluted with ammonia. Experiment results showed significantly high photocatalytic activity for the immobilized catalysts. The ammonia was removed more than 85% within 300 min of the process with optimum calcinations temperature 550 °C and pH 11. Kinetics of the photocatalytic reaction followed a pseudo-first order model. XRF, XRD and SEM analyses revealed a rather uniform coating of TiO2 on the support. By using floated TiO2/LECA as a photocatalyst in aqueous solution of NH 3 ? , the ammonia was photodegraded into N2 and H2 gases, while NO 2 ? and NO 3 ? were formed at very low concentrations.  相似文献   

20.
The purpose of this paper is to study the surface energetics of the polymer excipient cellulose acetate propionate (CAP) in the solid form. The net retention volumes, V N, for n-alkanes and polar solutes have been measured in the temperature range 353.15–403.15 K by inverse gas chromatography. The dispersive surface free energy, $ \gamma_{\text{S}}^{\text{d}} $ , and Lewis acid–base parameters $ K_{\text{a}} $ and $ K_{\text{b}} $ , have been determined using V N values. The $ \gamma_{\text{S}}^{\text{d}} $ values are decreased linearly with increase of temperature. The $ \gamma_{\text{S}}^{\text{d}} $ value at 353.15 K is 24.50 ± 1.54 mJ/m2, and the temperature gradient was found to be ?0.287 mJ/m2/K1. The $ K_{\text{a}} $ and $ K_{\text{b}} $ values are 0.410 ± 0.021 and 1.708 ± 0.388, respectively, which suggest that the CAP solid surface contain relatively more basic sites. The K a and K b values of CAP are compared with the similar values obtained on the cellulose acetate butyrate solid surface.  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