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 mechanism of the thermal decomposition of Fe2(SO4)3 in air has been studied at different temperatures (520-700 °C) using mainly 57Fe Mössbauer spectroscopy. Iron(III) oxides with corundum (), bixbyite (), spinel () and orthorhombic () structures were identified as solid products of this conversion. A significant influence of the heating temperature on the decomposition mechanism and on the phase composition of reaction products was found.  相似文献   

2.
Cuprous oxide (Cu2O) thin films have been deposited onto fluorine doped tin oxide (FTO) glass substrates by using electrochemical route. The structural, morphological, and chemical composition of the deposited films have been studied by using X-ray diffraction (XRD), Scanning electron microscopy (SEM) and Energy dispersive x-ray spectroscopy (EDAX) techniques respectively. The optical studies have been carried out by using UV-Vis spectroscopy. The effect of potential, pH and bath temperature onto absorption and band gap of Cu2O thin films have been studied. The highest sensitivity 6.25 mA·mM·cm- 2 is observed for the thin films which shows glucose concentration 7 mM in 0.1 M NaOH solution. The results indicates Cu2O is promising material for glucose sensor with high sensitivity, high stability, and repeatability.
Graphical abstract The surface morphology of Cu2O thin films was found to be tip-truncated octahedral. The films were  prepared by electrodeposition. The Cu2O thin films were used to construct low cost, highly sensitive and stable glucose sensor.
  相似文献   

3.
The viability of Lewis‐acid ionic liquids for the synthesis of low‐valent bismuth compounds is demonstrated. At room temperature, elemental bismuth and bismuth(III) cations synproportionate in the ionic liquid [BMIM]Cl/AlCl3 ([BMIM]+: 1‐n‐butyl‐3‐methylimidazolium) within minutes. The existence of bismuth polycations in the dark colored solution was proven by Raman spectroscopy. Dark‐red crystals of Bi5(AlCl4)3 were isolated from the ionic liquid and characterized by Raman spectroscopy and X‐ray crystallography (rhombohedral space‐group , a = 1187.1(2) pm, c = 3012.0(3) pm). The method allows the synthesis of bismuth cluster compounds under milder conditions than in high‐temperature melts and more conveniently and environmental friendly than in liquid SO2 with strongly oxidizing, toxic agents like SbF5 or AsF5.  相似文献   

4.
The oxidation of elemental palladium with oleum (65 % SO3) in the presence of barium carbonate in torch‐sealed glass ampoules at 180 °C leads to yellow single crystals of the heteroleptic palladate Ba2[Pd(HS2O7)2(S3O10)2] (triclinic, P ; Z=1; a=884.18(3), b=927.68(3), c=938.77(4) pm; α=60.473(1), β=80.266(2), γ=87.746(2)°). The crystal structure shows the Pd2+ ions in a square‐planar coordination of oxygen atoms of two hydrogendisulfate as well as of two trisulfate anions. The compound is the first example of the rarely seen S3O102? and HS2O7? anions acting as ligands in a complex anion and, moreover, the first heteroleptic polysulfatometallate known so far. The complex formation leads to a stabilization of the trisulfate anion relative to its uncoordinated congener. Ba2[Pd(HS2O7)2(S3O10)2] has been further characterized by vibrational spectroscopy and quantum chemical calculations. Thermal analyses by means of thermogravimetric/differential thermal analysis (TG/DTA) measurements show that the compound decomposes to yield elemental palladium and BaSO4.  相似文献   

5.
Effect of B3+ (%B2O3) percentage ratio in silica gel on the nature and the mechanism of defect formation and yields of paramagnetic radiation centers (PC) have been studied. The nature of paramagnetic radiation centers of electron hole type has been revealed and their behavior has been investigated as a function of temperature changes. It has been found that the interaction between paramagnetic centers formed at SiO2/B2O3 surface with absorbed water molecules (=1) leads to the formation of H and OH radicals. In addition the experiment shows that increased B3+ (%B2O3) percentage ratio in silica gel results in higher yields of paramagnetic centers and H-radicals. Possible mechanism of formation, localization and annihilation of non-equilibrium charge carriers (NCC) and their energy transfer to the adsorbed phase has been suggested.  相似文献   

6.
The mechanism of the spin‐forbidden quenching process O(1D) + CO2(1Σ) → O(3P) + CO2(1Σ) was investigated by ab initio quantum chemistry methods. The calculations showed the singlet potential surface [O(1D)+CO2] is attractive where a strongly bound intermediate complex CO3 is formed in the potential basin without a transition state, whereas the complex CO3 that is formed on the triplet surface [O(3P)+CO3] must overcome a barrier. The complex channel was documented by searching minimum energy intersection points in the region of the bound complex CO3 and calculating spin‐orbit coupling at the point. A direct channel was proposed by a study of cross point of singlet and triplet PESs with different collision angles and calculations of spin‐orbit coupling at those cross points in a nonbound region of the [O(1D)+CO3] system. The mechanism of the energy transfer is discussed on the basis of the theoretical results. © 2005 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

7.
The interactions of Al2O3, CeO2, Pt/Al2O3, and Pt/CeO2 films with SO2, SO2 + H2O, SO2 + O2, and SO2 + O2 + H2O in the temperature range 300–673 K at the partial pressures of SO2, O2, and H2O equal to 1.5 × 102, 1.5 × 102, and 3 × 102 Pa, respectively, were studied using X-ray photoelectron spectroscopy. The formation of surface sulfite at T 473 K (the S 2p 3/2 binding energy (E b) is 167.5 eV) and surface sulfate at T 573 K (E b = 169.2 eV) was observed in the reactions of Al2O3 and CeO2 with SO2. The formation of sulfates on the surface of CeO2 occurred much more effectively than in the case of Al2O3, and it was accompanied by the reduction of Ce(IV) to Ce(III). The formation of aluminum and cerium sulfates and sulfites on model Pt/Al2O3 and Pt/CeO2 catalysts occurred simultaneously with the formation of surface platinum sulfides (E b of S 2p 3/2 is 162.2 eV). The effects of oxygen and water vapor on the nature and yield of sulfur-containing products were studied.  相似文献   

8.
Five coordination compounds of bismuth, lanthanum and praseodymium nitrate with the oxygen‐coordinating chelate ligand (iPrO)2(O)PCH2P(O)(OiPr)2 (L) are reported: [Bi(NO3)3(L)2] ( 1 ), [La(NO3)3(L)2] ( 2 ), [Pr(NO3)3(L)2] ( 3 ), [La(NO3)3(L)(H2O)] ( 4 ) and [Pr(NO3)3(L)(H2O)] ( 5 ). The compounds were characterized by means of single crystal X‐ray crystallography, 1H and 31P NMR spectroscopy in solution, solid‐state 31P NMR spectroscopy, IR spectroscopy, DTA‐TG measurements ( 1 , 2 and 4 ), conductometry and electrospray ionization mass spectrometry (ESI‐MS). In addition, DFT calculations for model compounds of 1 and 2 support our experimental work. In the solid state mononuclear coordination compounds were observed for 1 — 3 , whereas compounds 4 and 5 gave one‐dimensional hydrogen‐bonded polymers via water‐nitrate coordination. Despite of the similar ionic radii of bismuth(III), lanthanum(III) and praseodymium(III) for a given coordination number the bismuth and lanthanide compounds 1 — 3 are not isostructural. The bismuth compound 1 shows a 9‐coordinate bismuth atom whereas lanthanum(III) and praseodymium(III) atoms are 10‐coordinate in the lanthanide complexes 2 — 5 . The general LnO10 coordination motif in compounds 2 — 5 is best described as a distorted bi‐capped square antiprism. The BiO9 polyhedron might be deduced from the LnO10 polyhedron by replacing one oxygen ligand with a stereochemically active lone pair. The one‐to‐one complexes 4 and 5 dissociate in solution to give the corresponding one‐to‐two complexes 2 and 3 , respectively, and solvated Ln(NO3)3. In contrast to the lanthanides, the one‐to‐two bismuth complex 1 is less stable in CH3CN solution and partially dissociates to give solvated Bi(NO3)3 and (iPrO)2(O)PCH2P(O)(OiPr)2.  相似文献   

9.
O(1D), produced from the photolysis of N2O at 2139 Å, reacts with N2O in accord with: We have used the method of chemical difference to obtain an accurate measure of k2/k3 = 0.59 ± 0.01. Furthermore, the quantum yield of production of O(3P), either on direct photolysis or on deactivation of O(1D) by N2O, is less than 0.02 and probably zero.  相似文献   

10.
A series of triarylantimony dichrysanthemate compounds of the type Ar3Sb(O2CR)2 [Ar=C6H5, 4-CH3C6H4, 3-CH3C6H4, -CH3C6H4, 4-ClC6H4; R=4-ClC6H4CH(i-Pr), cis-Cl2C:CH trans-Cl2C:CH ] have been synthesized and characterized by elemental analysis, infrared spectra, 1H NMR spectra and mass spectra. Some activities of these compounds in plant growth regulation have been determined. Their results indicate that the derivatives of cis-dichlorochrysanthemic acid and trans-dichlorochrysanthemic acid significantly promote rooting of excised cucumber cotyledons at 10 ppm. An X ray structure determination has been carried out as follows for Ph3Sb(O2CCHCMe2CMe2)2: orthorhombic, space group Pbcn, Z=4, structure solution with 2385 independent reflections, R=0.035. Lattice dimensions at 26 °C: a=15.616(3) Å, b=10.275(2) Å, c=20.201(5) Å, V=3241(2) Å3, ρ=1.302 g cm−3. © 1998 John Wiley & Sons, Ltd.  相似文献   

11.
A novel bidentate Schiff base ligand L (L = N-(4-amino-2-chloro-phenyl)-2-hydroxybenzaldehyde) and the subsequent octahedral manganese(III) Schiff base complex MnL 3 have been synthesized and characterized by, FT-IR spectroscopy and elemental analyses (CHN). Additionally, Schiff base ligand has been characterized by 1H NMR spectroscopy. Thermogravimetric analysis of the ligand and its metal complexes reveals their thermal stability and decomposition pattern. Thus, the MnL 3 complex has been investigated as a novel precursor for the facile preparation of Mn3O4 nanoparticles via solid-state thermal decomposition under aerobic conditions, at a temperature of ca. 450 °C The resulting Mn3O4 nanocrystals were characterized by FT-IR spectroscopy, X-ray powder diffraction (XRPD), transmission electron microscopy (TEM) and scanning electron microscopy (SEM). The XRPD studies reveal the characteristic diffraction peaks indexed to the Mn3O4 hausmanite structure, while the absence of additional peaks tends to clearly indicate the high purity of the sample. In addition, the TEM/SEM investigations displayed the nanoplate shape of the rather monodisperse crystalline Mn3O4 nanoparticles, with an average diameter of ca. 10 nm. The statistical distribution of the nanoparticles size has to be provided with an histogram graphic.  相似文献   

12.
The metastable decompositions of trimethylsilylmethanol, (CH3)3SiCH2OH (MW: 104, 1) and methoxytrimethylsilane, (CH3)3SiOCH3 (MW: 104, 2) upon electron ionization have been investigated by use of mass-analyzed ion kinetic energy (MIKE) spectroscopy and D labeling. The metastable ions of 1 ·+ decompose to give the fragment ions m/z 89 (CH 3 · loss) and 73 (·CH2OH loss), whereas those of 2 ·+ only yield the fragment ion m/z 89 (CH 3 · loss). The latter fragment ion is generated by loss of a methyl radical from the trimethylsilyl group via a simple cleavage reaction as shown by D labeling. However, the fragment ions m/z 89 and 73 from 1 ·+ are generated following an almost statistical exchange of the original methyl and methylene hydrogen atoms in the molecular ion as shown also by D labeling. This exchange indicates a complex rearrangement of the molecular ion of 1 ·+ prior to metastable decomposition for which as key step a 1,2-trimethylsilyl group migration from carbon to oxygen is suggested. A different behavior is also found between the source-generated m/z 89 ions from 1 ·+ which decompose in the metastable time region to give ions m/z 61 by loss of ethylene and those from 2 ·+ which decompose in the metastable region to yield ions m/z 59 by elimination of formaldehyde.  相似文献   

13.
This paper reports on a procedure for the synthesis of the compound [RuNO(NH3)3Cl(H2O)]Cl2. The complex was studied by IR spectroscopy and X-ray phase and X-ray diffraction analyses. Crystal data for RuCl3N4O2H11: a = 13.151(2), b = 6.852(1), c = 10.361(1) , V = 933.6(2) 2, space group Pna2 1 , Z = 4, d calc = 2.181 g/cm3, d exp = 2.178 g/cm3. The structure consists of the [RuNO(NH3)3Cl(H2O)]2+ complex cations and Cl anions. The compound crystallizes as small orange red isometric orthorhombic crystals well soluble in water and insoluble in concentrated hydrochloric acid and organic solvents and is stable when stored in air.  相似文献   

14.
Radioactive as well as inactive cobalt precipitated from an aqueous solution onto the surface of -Fe2O3 was investigated by the Mössbauer spectroscopy of57Fe. The experiment performed with the57Co-doped sample showed that some of the cobalt atoms diffused into the -Fe2O3 lattice owing to 1 h of heat treatment carried out at 300 °C.  相似文献   

15.
The cationic metal cage complex (1,3,6,8,10,13,16,19-octaazabicyclo[6.6.6]cicosane)cobalt(III), Co(sep)3+ has been investigated as a potential pillaring reagent for Na+-magadiite (Na1.7Si14O27.9(OH)1.9 · 7.6 H2O) a synthetic layered sodium silicate. Reaction of Na+-magadiite with aqueous solutions of Co(sep)Cl3 at 25°C resulted in the binding of Co(sep)3+ cations to the external crystalline surface of the layered silicate. In contrast, an intercalated product exhibiting a 17.6 Å basal spacing was generated by reaction at 100°C.29Si MAS NMR and FT-IR spectroscopy indicate that Co(sep)3+ intercalated reaction products retain the magadiite layer structure. Moreover, scanning electron micrographs of the reaction products showed retention of the original particle morphology, suggesting a topotactic intercalation. However, during intercalation, some of the Co(sep)3+ was found to undergo an unusual demetalation reaction leaving a combination of Co(II) and Co(sep)3+ between the layers. Nitrogen surface area analysis showed that only a small amount of microporous surface existed in the Co(sep)3+ intercalated derivative, suggesting that most of the interlayer space is stuffed with cobalt species.  相似文献   

16.
The hydrolysis of the [Pt(dien)H2O]2+ and [Pd(dien)H2O]2+ complexes has been investigated by potentiometry at 298 K, in 0.1 mol dm–3 aqueous NaClO4. Least-squares treatment of the data obtained indicates the formation of mononuclear and -hydroxo-bridged dinuclear complexes with stability constants: log 11 = –6.94 for [Pt(dien)OH]+, log 11 = –7.16 for [Pd(dien)OH]+, and also log 22 = –9.37 for [Pt2(dien)2(OH)2]2+ and log 22 = –10.56 for [Pd2(dien)2(OH)2]2+. At pH values > 5.5, formation of the dimer becomes significant for the PtII complex, and at pH > 6.5 for the PdII complex. These results have been analyzed in relation to the antitumor activity of PtII complexes.  相似文献   

17.
Ru(P)2(TaiMe)Cl2 [P = PPh3; TaiMe = 1-methyl-2-(p-tolylazo)imidazole] possesses a cis-RuCl2 configuration. Dechlorination was carried out in acetone solution by Ag+ and the solvated species reacted with N,O-[oxinate (ox), -picolinate (pic)], O,O-[salicylate (sa), 3-formylsalicylate (3-fsa), 5-formylsalicylate (5-fsa)] and S,S-[xanthate (xan), dithiocarbamate (dtc)] chelators to yield [Ru(P)2(TaiMe)(N,O/O,O/S,S)] (ClO4) n (n = 1 for N,O- and S,S-chelators; n = 0 for O,O-chelators). The complexes were characterised by microanalysis, molar conductance, i.r., u.v.–vis. and 1H-n.m.r. data. Isomeric structures in some complexes were identified by 1H-n.m.r. spectra. The electronic spectra show a high intensity ( 104) m.l.c.t. transition in the visible region together with a weak shoulder ( 103) at longer wavelength. Redox studies exhibit a RuIII/RuII couple and quasireversible azo reduction.  相似文献   

18.
A new protonated borophosphate (H3O)Mg(H2O)2[BP2O8]·H2O ( 1 ) was synthesized under mild hydrothermal conditions and characterized by single‐crystal X‐ray diffraction, FTIR spectroscopy and TG‐DTA. The compound crystallizes in the hexagonal system, space group P6(1)22 (No 178), a = 9.4462(7) Å, c = 15.759(2) Å, V = 1217.8(2) Å3, and Z = 6. There exist infinite helical $^1_\infty$ {[BP2O8]3–} ribbons built up from corner‐sharing PO4 and BO4 tetrahedra, which are connected by MgO4(H2O)2 leading to an infinite three‐dimensional open‐framework. The H3O+ ions are located at the free thread of the helical ribbons, whereas crystallized water occupy the channels of the helical ribbons. The dehydration of the compound occurs at a higher temperature which is presumably due to the anisotropic hydrogen bonds in the crystal structure. The luminescent properties of the compound were studied.  相似文献   

19.
Two novel heterometallic cubane-like and double cubane-like clusters, {MoCu3S3(S2COEt)}(O)(Ph3P)3 I and {Mo2Cu6S6(SCMe3)2}(O)2(Ph3P)4 II, were synthesized by reaction of {MoCu2S3}(O)(Ph3P)3 with CuS2COEt and CuSCMe3, respectively. ClusterI crystallized in the triclinic space group (2) witha=12.766(6) Å,b=22.904(5) Å,c=10.522(3) Å, =99.86(2)°, =109.68(2)°, =86.84(3)°,V=2854(2) Å3,Z=2,R=0.049 for 6622 observed reflections (I>5(I)) and 410 variables. ClusterII crystallized in the triclinic space group (2) with dimensionsa=14.212(4) Å,b=14.725(5) Å,c=12.396(8) Å, =110.32(4)°, =90.40(5)°, =62.88(2)°,V=2129(2) Å3,Z=1,R=0.039 for 6020 observed reflections (I>3(I)) and 461 variables. ClusterI consists of a neutral cubane-like molecule with the core {MoCu3S3(S2COEt)}2+, in which one corner of the cubane-like core is a novel triply bridging bidentate 1,1-dithiolato (xanthate, S2COEt) ligand. ClusterII is a double cubane-like one, in which two cubane-like cores {MoCu3S3(SCMe3)}2+ are connected by two Cu-S bonds of the triply bridging monothiolato (SCMe 3 ) ligand. Two different pathways of unit construction from a small heterometallic cluster {MoCu2S3}(O)(Ph3P)3 have been outlined. Comparisons of the selected bond lengths and bond angles for the cubane-like core {MoCu3S3 X} (X=Cl, Br, S2COEt, SCMe 3 ) are given. Spectroscopic properties of the title clusters are also reported.  相似文献   

20.
An explicit function has been derived for the potential-energy surface of the ground state of ClO3 with the six interatomic distances as variables. This surface is valid over all configurations of the atoms. The surface has been used to calculate classical trajectories for the reactions R1: O(3P2)+ClO(2Π3/2)→ O2(3∑)+Cl(2)P3/2 and R2: Cl(2P3/2)+O3(1A1)→ClO(2Π3/2)+O2(3∑). An appreciable fraction (~1/3) of the reactive trajectories for R1 go through a long-lived complex ClOO(2A″). The cross section decreases with increasing rotational state of the ClO; and 37% of the energy release is vibrational. The calculated rate constant at 298°K is 2.6 × 10?11 cm3/molecule sec. For reaction R2 there is no evidence of long-lived complexes. The product ClO is predominantly found in the backward-scattering direction. Most of the internal energy is carries off by ClO but O2 carried off most translational energy. An Arrhenius expression has been deduced from calculations at 220 and 300°K to give an A factor of 2.488 × 10?11 cm3/molecule sec and an activation energy of 1.543 kJ/mol.  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