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 507 毫秒
1.
The reaction pathways and energetics for the reaction of methane with CaO are discussed on the singlet spin state potential energy surface at the B3LYP/6-311+G(2df,2p) and QCISD/6-311++G(3df,3pd)//B3LYP/6-311+G(2df,2p) levels of theory. The reaction of methane with CaO is proposed to proceed in the following reaction pathways: CaO + CH4 → CaOCH4 → [TS] → CaOH + CH3, CaO + CH4 → OCaCH4 → [TS] → HOCaCH3 → CaOH + CH3 or [TS] → CaCH3OH → Ca + CH3OH, and OCaCH4 → [TS] → HCaOCH3 → CaOCH3 + H or [TS] → CaCH3OH → Ca + CH3OH. The gas-phase methane–methanol conversion by CaO is suggested to proceed via two kinds of important reaction intermediates, HOCaCH3 and HCaOCH3, and the reaction pathway via the hydroxy intermediate (HOCaCH3) is energetically more favorable than the other one via the methoxy intermediate (HCaOCH3). The hydroxy intermediate HOCaCH3 is predicted to be the energetically most preferred configuration in the reaction of CaO + CH4. Meanwhile, these three product channels (CaOH + CH3, CaOCH3 + H and Ca + CH3OH) are expected to compete with each other, and the formation of methyl radical is the most preferable pathway energetically. On the other hand, the intermediates HCaOCH3 and HOCaCH3 are predicted to be the energetically preferred configuration in the reaction of Ca + CH3OH, which is precisely the reverse reaction of methane hydroxylation.  相似文献   

2.
The molecular geometries of three structurally related compounds have been determined by electron diffraction in the gas phase. Acetylacetone, which exists primarily as the enol tautomer, was found to have a planar symmetric ring with the following rg values: C1-C2 = 1.405± 0.005 Å, C2-C4 = 1.510± 0.005 Å, C-O = 1.287± 0.006 Å, C-H = 1.090± 0.010 Å, ∠C2C1C3 = 118.3 ± 1.8°, ∠C1C3O1 = 123.2± 1.7°, ∠C1C3C5 = 122.0± 1.2°, and ∠CCH = 110.2 ± 2.1°. A model in which the enol proton is in the ring plane located symmetrically between the oxygen atoms is in best agreement with the diffraction data. The structure of trifluoroacetylacetone is similar to that of acetylacetone. The rg values for this compound are C1-C3 = 1.4164 ± 0.006 Å, C3-C5 = 1.511 ± 0.021 Å, C2-C4 = 1.536 ± 0.018 Å, C-O = 1.270 ± 0.008 Å, C-H = 1.088 ± 0.039 Å, C-F = 1.340 ± 0.005 Å, ∠C2C1C3 = 117.2 ± 1.8°, ∠C1C2O2 = 123.6 ± 1.7°, ∠C1C3C5 = 118.1 ± 2.3°, ∠C1C2C4 = 123.0 ± 1.4°, ∠CCH = 109.0± 4.8°, and ∠CCF = 110.6± 0.8°. The rg values for the trifluoroacetone are: C1-C2 = 1.481 ± 0.019 Å, C1-C3 = 1.562 ± 0.011 Å, CO = 1.207 ± 0.006 Å, C-H = 1.089 ± 0.024 Å, C-F = 1.339 ± 0.003 Å, ∠C2C1O = 122.0 ± 1.1°, ∠C3C1O = 116.8 ± 0.7 °, ∠CCH = 105.0 ± 2.2 °, and ∠CCF = 110.7 ± 0.3°. The significance of the error estimates is discussed briefly.  相似文献   

3.
The present paper mainly studied the phase formation and reaction pathway of the Al–Ti–Si system in detail by thermal analysis combined with XRD and SEM observations. The phase formation sequence in Al–Ti–Si system from starting mixtures to final products with increasing temperature can be described as following: Al(l) + Ti(s) + Si(s) → (Al–Si)(l) + Ti(s) + Si(s) → Ti(Al,Si)3(s) + Si(s)Ti5(Si,Al)3 + Al(l). More importantly, the solubility of Si in Ti(Al,Si)3 decreased gradually while that of Al in Ti5(Si,Al)3 increased with temperature increasing, suggesting the transportation of Si atoms from intermediate aluminides Ti(Al,Si)3 to final stable silicides Ti5(Si,Al)3 and hence further confirming the formation of Ti5(Si,Al)3 at the expense of Ti(Al,Si)3.  相似文献   

4.
Two types of ammonium uranyl nitrate (NH4)2UO2(NO3)4·2H2O and NH4UO2(NO3)3, were thermally decomposed and reduced in a TG-DTA unit in nitrogen, air, and hydrogen atmospheres. Various intermediate phases produced by the thermal decomposition and reduction process were investigated by an X-ray diffraction analysis and a TG/DTA analysis. Both (NH4)2UO2(NO3)4·2H2O and NH4UO2(NO3)3 decomposed to amorphous UO3 regardless of the atmosphere used. The amorphous UO3 from (NH4)2UO2(NO3)4·2H2O was crystallized to γ-UO3 regardless of the atmosphere used without a change in weight. The amorphous UO3 obtained from decomposition of NH4UO2(NO3)3 was crystallized to α-UO3 under a nitrogen and air atmosphere, and to β-UO3 under a hydrogen atmosphere without a change in weight. Under each atmosphere, the reaction paths of (NH4)2UO2(NO3)4·2H2O and NH4UO2(NO3)3 were as follows: under a nitrogen atmosphere: (NH4)2UO2(NO3)4·2H2O → (NH4)2UO2(NO3)4·H2O → (NH4)2UO2(NO3)4 → NH4UO2(NO3)3 → A-UO3 → γ-UO3 → U3O8, NH4UO2(NO3)3 → A-UO3 → α-UO3 → U3O8; under an air atmosphere: (NH4)2UO2(NO3)4·2H2O → (NH4)2UO2(NO3)4·H2O → (NH4)2UO2(NO3)4 → NH4UO2(NO3)3 → A-UO3 → γ-UO3 → U3O8, NH4UO2(NO3)3 → A-UO3 → α-UO3 → U3O8; and under a hydrogen atmosphere: (NH4)2UO2(NO3)4·2H2O → (NH4)2UO2(NO3)4·H2O → (NH4)2UO2(NO3)4 → NH4UO2(NO3)3 → A-UO3 → γ-UO3 → α-U3O8 → UO2, NH4 UO2(NO3)3 → A-UO3 → β-UO3 → α-U3O8 → UO2.  相似文献   

5.
Extraction and carrier mediated transport through bulk liquid membrane and supported liquid membrane systems have wide applications in separation technology. This paper highlights the use of six noncyclic receptors (podands) having variations in chain length and end group for the removal of urea using liquid membrane system. These receptors R1, R2, R3, R4, R5, R6 are diethylene glycol dimethyl ether, diethylene glycol dibutyl ether, diethylene glycol dibenzoate, diethylene glycol, triethylene glycol and tetraethylene glycol respectively. The sequence of extraction and transport of urea by BLM system using various receptors is R2 > R3 > R1 > R4 > R5 > R6 and R6 ≈ R3 > R5 > R4 > R1 > R2 respectively. Receptor R2 containing butyl end group is best extractant while receptor R6 with flexible backbone is best carrier and this carrier efficiency is used to remove urea using BLM system from the feed phase by recyclization process up to 88.16%. The experimental results influenced by concentration of receptors and urea. Effect of time was also studied.  相似文献   

6.
《Tetrahedron》1987,43(1):109-114
The relative rates of the addition of CH3TiCl3 and CH3Ti (OCHMe2)3 to various cyclic ketones of different ring size have been determined. In contrast to CH3Li and CH3MgCl, significant rate differences were observed. For CH3TiCl3 (at 18°C) the following decreasing order in rate pertains, the numbers symbolizing the ring size: 6 > 5 > 7 > 15 > 8 > 12 > 9 > 11 > 10. For CH3Ti (OCHMe2)3 (at +22°C) it is slightly different: 6 > 5 > 7 > 8 > 15 > 9 > 12 > 13 > 11 > 10. Most of the results can be explained on the basis of Brown's hypothesis of I-strain. However, the 15-membered ring cyclopentadecanone represents the major “irregularity”.  相似文献   

7.
This investigation was undertaken to determine the antioxidant activity of a range of fullerenes C60 and C70 in order to rank them according to their comparative efficiency. The model reaction of initiated (2,2′- azobisisobutyronitrile, AIBN) cumene oxidation was used to determine rate constants for addition of radicals to fullerenes. Measurements of oxidation rates in the presence of different fullerenes showed that the antioxidant activity as well as the mechanism and mode of inhibition were different for fullerenes C60 and C70 and fullerene soot. All fullerenes - C60 of gold grade, C60/C70 (93/7, mix 1), C60/C70 (80 ± 5/20 ± 5, mix 2) and C70 operated as alkyl radical acceptora, whereas fullerene soot surprisingly retarded the model reaction by a dual mode similar to that for the fullerenes and with an induction period like many of the sterically hindered phenolic and amine antioxidants. For the C60 and C70 the oxidation rates were found to depend linearly on the reciprocal square root of the concentration over a sufficiently wide range thereby fitting the mechanism for the addition of cumylalkyl radicals to the fullerene core. This is consistent with literature data on the more ready and rapid addition of alkyl and alkoxy radicals to the fullerenes compared with peroxy radicals. Rate constants for the addition of cumyl radicals to the fullerenes were determined to be k(333K) = (1.9 ± 0.2) × 108 (C60); (2.3 ± 0.2) × 108 (C60/C70, mix 1); (2.7 ± 0.2) × 108 (C60/C70, mix 2); (3.0 ± 0.3) × 108 (C70), M−1 s−1. The increasing C70 constituent in the fullerenes leads to a corresponding increase in the rate constant.The fullerene soot inhibits the model reaction according to the mechanism of trapping of peroxy radicals; the oxidation proceeds with a pronounced induction period and kinetic curves are linear in semi-logarithmic coordinates.For the first time the effective concentration of inhibiting centres and inhibition rate constants for the fullerene soot have been determined to be fn[C60−soot] = (2.0 ± 0.1) × 10−4 mol g−1 and kinh = (6.5 ± 1.5) × 103 M−1 s−1 respectively.The kinetic data obtained specify the level of antioxidant activity for the commercial fullerenes and scope for their rational use in different composites. The results may be helpful for designing an optimal profile of composites containing fullerenes.  相似文献   

8.
Sabo M  Matúška J  Matejčík S 《Talanta》2011,85(1):400-405
This study deals with O2 generation in corona discharge (CD) in point to plane geometry for single flow ion mobility spectrometry (IMS) with gas outlet located behind the ionization source. We have designed CD of special geometry in order to achieve the high O2 yield. Using this ion source we have achieved in zero air conditions that up to 74% all negative ions were O2 or O2(H2O). It has been demonstrated that the non-electronegative nitrogen positively influences the efficiency of O2 generation in O2/N2 mixtures. The reduced ion mobility of 2.27 cm2 V−1 s−1 has been measured for O2/O2(H2O) ions in zero air. Additional ions detected in zero air (less than 200 ppb CO2) using the mass spectrometric and IMS technique were, NO2, N2O2 (2.37 cm2 V−1 s−1), NO3, N2O3 and N2O3(H2O). The CO3 and CO4 ions have been detected after the introduction of 5 ppm CO2 into zero air.  相似文献   

9.
Two pure strontium borates SrB2O4·4H2O and SrB2O4 have been synthesized and characterized by means of chemical analysis and XRD, FT-IR, DTA-TG techniques. The molar enthalpies of solution of SrB2O4·4H2O and SrB2O4 in 1 mol dm−3 HCl(aq) were measured to be −(9.92 ± 0.20) kJ mol−1 and −(81.27 ± 0.30) kJ mol−1, respectively. The molar enthalpy of solution of Sr(OH)2·8H2O in (HCl + H3BO3)(aq) were determined to be −(51.69 ± 0.15) kJ mol−1. With the use of the enthalpy of solution of H3BO3 in 1 mol dm−3 HCl(aq), and the standard molar enthalpies of formation for Sr(OH)2·8H2O(s), H3BO3(s), and H2O(l), the standard molar enthalpies of formation of −(3253.1 ± 1.7) kJ mol−1 for SrB2O4·4H2O, and of −(2038.4 ± 1.7) kJ mol−1 for SrB2O4 were obtained.  相似文献   

10.
Tren amine cations [(C2H4NH3)3N]3+ and zirconate or tantalate anions adopt a ternary symmetry in two hydrates, [H3tren]2·(ZrF7)2·9H2O and [H3tren]6·(ZrF7)2·(TaOF6)4·3H2O, which crystallise in R32 space group with aH = 8.871 (2) Å, cH = 38.16 (1) Å and aH = 8.758 (2) Å, cH = 30.112 (9) Å, respectively. Similar [H3tren]2·(MX7)2·H2O (M = Zr, Ta; X = F, O) sheets are found in both structures; they are separated by a water layer (Ow(2)-Ow(3)) in [H3tren]2·(ZrF7)2·9H2O. Dehydration of [H3tren]2·(ZrF7)2·9H2O starts at room temperature and ends at 90 °C to give [H3tren]2·(ZrF7)2·H2O. [H3tren]2·(ZrF7)2·H2O layers remain probably unchanged during this dehydration and the existence of one intermediate [H3tren]2·(ZrF7)2·3H2O hydrate is assumed. Ow(1) molecules are tightly hydrogen bonded with -NH3+ groups and decomposition of [H3tren]2·(ZrF7)2·H2O occurs from 210 °C to 500 °C to give successively [H3tren]2·(ZrF6)·(Zr2F12) (285 °C), an intermediate unknown phase (320 °C) and ZrF4.  相似文献   

11.
The phase relations in the pseudo-binary system SrO-Fe2O3 have been investigated in air up to 1150°C by means of powder X-ray diffraction and thermal analysis. Sr3Fe2O7−δ, SrFeO3−δ and SrFe12O19 are stable phases in the entire investigated temperature region, whereas Sr2FeO4−δ and Sr4Fe3O10−δ decompose above 930±10°C and 850±25°C, respectively. Sr4Fe6O13±δ is entropy-stabilized relative to SrFeO3−δ and SrFe12O19 above 775±25°C. Extended solid-solution SrxFeO3−δ was demonstrated. On the Fe-deficient side, the extent of solid solubility appeared to decrease gradually with temperature, whereas an abrupt decrease due to formation of Sr4Fe6O13±δ was observed above 775°C on the Sr-deficient side.  相似文献   

12.
Differential scanning calorimetry and high temperature oxide melt solution calorimetry are used to study enthalpy of phase transition and enthalpies of formation of Cu2P2O7 and Cu3(P2O6OH)2. α-Cu2P2O7 is reversibly transformed to β-Cu2P2O7 at 338–363 K with an enthalpy of phase transition of 0.15 ± 0.03 kJ mol−1. Enthalpies of formation from oxides of α-Cu2P2O7 and Cu3(P2O6OH)2 are −279.0 ± 1.4 kJ mol−1 and −538.8 ± 2.7 kJ mol−1, and their standard enthalpies of formation (enthalpy of formation from elements) are −2096.1 ± 4.3 kJ mol−1 and −4302.7 ± 6.7 kJ mol−1, respectively. The presence of hydrogen in diphosphate groups changes the geometry of Cu(II) and decreases acid–base interaction between oxide components in Cu3(P2O6OH)2, thus decreasing its thermodynamic stability.  相似文献   

13.
We have synthesized the composition x = 0.01 of the (Sr1-xLax)2(Ta1-xTix)2O7 solid solution, mixing the ferroelectric perovskite phases Sr2Ta2O7 and La2Ti2O7. Related oxide and oxynitride materials have been produced as thin films by magnetron radio frequency sputtering. Reactive sputter deposition was conducted at 750 °C under a 75 vol.% (Ar) + 25 vol.% (N2,O2) mixture. An oxygen-free plasma leads to the deposition of an oxynitride film (Sr0.99La0.01) (Ta0.99Ti0.01)O2N, characterized by a band gap Eg = 2.30 eV and a preferential (001) epitaxial growth on (001) SrTiO3 substrate. Its dielectric constant and loss tangent are respectively Epsilon' = 60 (at 1 kHz) and tanDelta = 62.5 × 10−3. In oxygen-rich conditions (vol.%N2 ≤ 15%), (110) epitaxial (Sr0.99La0.01)2(Ta0.99Ti0.01)2O7 oxides films are deposited, associated to a larger band gap value (Eg = 4.55 eV). The oxide films permittivity varies from 45 to 25 (at 1 kHz) in correlation with the decrease in crystalline orientation; measured losses are lower than 5.10−3. For 20 ≤ vol.% N2 ≤ 24.55, the films are poorly crystallized, leading to very low permittivities (minimum Epsilon' = 3). A correlation between the dielectric losses and the presence of an oxynitride phase in the samples is highlighted.  相似文献   

14.
This study was carried out to investigate the characteristics of an oxidative-dissolution of fission products (FP) when uranium (U) is dissolved in a Na2CO3–H2O2 carbonate solution. Simulated FP-oxides which contained 12 components were added to the solution to examine their dissolution behaviors. It was found that H2O2 was an effective oxidant to minimize the dissolution of FP. For the 0.5 M Na2CO3–0.5 M H2O2 solution, such elements as Re, Te, Cs, and MoO2 were dissolved with yields of 98 ± 2%, 98 ± 2%, 93 ± 2%, and 26 ± 3%, respectively, for 2 h. Among these components, Re, Te, and Cs were completely dissolved within 10–20 min without regard to the concentrations of Na2CO3, and H2O2 due to their high solubility in the carbonate solution with and without H2O2. However, MoO2 was very slowly dissolved and its yield was 29 ± 3% for 4 h. The pH of the dissolved solution revealed the greatest influence on the dissolution yields of the FP, exhibiting the most effective pH condition in the range of 10–12 in order to create a considerable suppression of the co-dissolution of FP during the oxidative-dissolution of U.  相似文献   

15.
Ti/IrO2(x) + MnO2(1-x) anodes have been fabricated by thermal decomposition of a mixed H2IrCl6 and Mn(NO3)2 hydrosolvent. Cyclic voltammetry (CV) and polarization curve have been utilized to investigate the electrochemical behavior and electrocatalytic activity of Ti/IrO2(x) + MnO2(1-x) anodes in 0.5 M NaCl solution (pH = 2). Ti/IrO2+MnO2 anode with 70 mol% IrO2 content has the maximum value of q*, indicating that Ti/IrO2(0.7) + MnO2(0.3) anode has the most excellent electrocatalytic activity for the synchronal evolution of Cl2 and O2 in dilute NaCl solution. Tafel lines displayed two distinct linear regions with one of the slope close to 62 mV dec−1 in the low potential region and the other close to 295 mV dec−1 in the high potential region. Electrochemical impedance spectroscopic is employed to investigate the impedance behavior of Ti/IrO2(x) + MnO2(1-x) anodes in 0.5 M NaCl solution. It is observed that as the R ct, R s and R f values for Ti/IrO2(0.7) + MnO2(0.3) anode become smaller, electrocatalytic activity of Ti/IrO2(0.7) + MnO2(0.3) anode becomes better than that of other Ti/IrO2 + MnO2 anodes with different compositions. Ti/IrO2(0.7) + MnO2(0.3) anode fabricated at 400 °C has been observed to possess the highest service life of 225 h, whereas the accelerated life test is carried out under the anodic current of 2 A cm−2 at the temperature of 50 °C in 0.5 M NaCl solution (pH = 2).  相似文献   

16.
Deficiency in the A sublattice of perovskite-type Sr1– y Fe0.8Ti0.2O3–δ (y=0–0.06) leads to suppression of oxygen-vacancy ordering and to increasing oxygen ionic conductivity, unit cell volume, thermal expansion, and stability in CO2-containing atmospheres. The total electrical conductivity, predominantly p-type electronic in air, decreases with increasing A-site deficiency at 300–700 K and is essentially independent of the cation vacancy concentration at higher temperatures. Oxygen ion transference numbers for Sr1– y Fe0.8Ti0.2O3–δ in air, estimated from the faradaic efficiency and oxygen permeation data, vary in the range from 0.002 to 0.015 at 1073–1223 K, increasing with temperature. The maximum ionic conductivity was observed for Sr0.97Fe0.8Ti0.2O3–δ ceramics. In the system Sr0.97Fe1– x Ti x O3–δ (x=0.1–0.6), thermal expansion and electron-hole conductivity both decrease with x. Moderate additions of titanium (up to 20%) in Sr0.97(Fe,Ti)O3–δ result in higher ionic conductivity and lower activation energy for ionic transport, owing to disordering in the oxygen sublattice; further doping decreases the ionic conduction. It was shown that time degradation of the oxygen permeability, characteristic of Sr(Fe,Ti)O3–δ membranes and resulting from partial ordering processes, can be reduced by cycling of the oxygen pressure at the membrane permeate side. Thermal expansion coefficients of Sr1– y Ti1– x Fe x O3–δ (x=0.10–0.60, y=0–0.06) in air are in the range (11.7–16.5)×10–6 K–1 at 350–750 K and (16.6–31.1)×10–6 K–1 at 750–1050 K. Electronic Publication  相似文献   

17.
A novel gelling method was studied to stabilize phase change material Na2HPO4 · 12H2O with amylose grafted sodium acrylate. Gelled Na2HPO4 · 12H2O shows stable heat storage performance prepared at optimized conditions: 2.7mass/mass% sodium acrylate, 0.4 mass/mass% amylose, 0.05–0.09 mass/mass% N, N′-methylenebisacrylamide, 0.05–0.09 mass/mass% K2S2O8 and Na2SO3 (mass ratio 1:1), at 50 °C. Na2HPO4 · 12H2O was dispersed in gel network as tiny crystals less than 0.1 mm. Melting points were in the range 35.4 ± 2 °C. Short-term thermal cycling proves the effectiveness of the novel method for eliminating phase separation in the gelled salt. Adiabatic calorimetric measurement of heat capacities shows two phase transitions, which correspond to melting of Na2HPO4 · 12H2O and freezable bond water in gel, respectively. Heat of fusion of pure Na2HPO4 · 12H2O was determined as 260.9 J g−1. Distribution of extra water is: free water:freezable water:nonfreezing water = 0:0.85:0.15.  相似文献   

18.
We prepared Pt catalysts supported on various metal oxides, viz., ZrO2, CeO2, TiO2, yttria-stabilized zirconia (YSZ), SiO2, SiO2–Al2O3, and γ-Al2O3, using an incipient wetness method and applied them to propane combustion. In the cases of ZrO2-, CeO2-, and TiO2-supported Pt catalysts, supports with different surface areas were also used. The Pt dispersion in Pt catalysts supported on metal oxides increased with increasing surface area of the support for the same metal oxide. Pt catalysts on supports with lower surface areas (ZrO2, CeO2, and TiO2) showed higher catalytic activities for propane combustion than did Pt catalysts on supports with higher surface areas. The catalytic activity decreased in the following order: Pt/ZrO2 (2) > Pt/CeO2 (9) > Pt/TiO2 (1) = Pt/SiO2 (350) > Pt/ZrO2 (18) = Pt/YSZ > Pt/TiO2 (330) > Pt/SiO2–Al2O3 (350) > Pt/ZrO2 (73) > Pt/γ-Al2O3 (180) > Pt/CeO2 (160). The catalytic activity is inversely proportional to the amount of O2 chemisorbed up to the reaction temperature. It can be concluded that metallic Pt is essential for propane combustion and is maintained for the Pt catalysts with large Pt metal particles, which can be prepared by using a support with a low surface area.  相似文献   

19.

Abstract  

Long-term annealing of La2NiO4+δ single crystals at 1,273 K in air leads to the formation of nickel-rich Ruddlesden–Popper phases at the single-crystal surfaces. Both the composition and the morphology of these phases depend on the surface orientation; whereas only crystallites of La4Ni3O10−δ were observed on (001) surfaces, both La3Ni2O7−δ and La4Ni3O10−δ were formed on (100) surfaces. The formation of the nickel-rich RP phases is believed to be due to surface segregation of nickel or evaporation of a volatile lanthanum species. The crystallographic (001) planes inside the La3Ni2O7−δ and La4Ni3O10−δ crystallites were found to be oriented in the direction of preferential crystallite growth, which indicates that the diffusion of lanthanum and nickel cations is faster along the crystallographic (001) planes than perpendicular to these planes.  相似文献   

20.
Comprehensive two-dimensional chromatography (LC × LC) using combinations of two columns (C18 × CN and C18 × NH2) was employed with electrospray (ESI) mass spectrometry to analyze platycosides from root extract. Based on the capability of the C18, CN and NH2 columns to separate the platycosides, the orthogonality in two-dimensional space according to each combination of columns was predicted from the correlation coefficients between the retention times of the 17 compounds separated by the independent CN and C18 columns, and NH2 and C18 columns. The expected distribution of the peaks was also compared with the two-dimensional plots obtained by practical separation in an LC × LC system. The increased peak capacities using C18 × NH2 allowed three minor components and five isomers of the platycosides to be newly separated, which were not identified with 1D-LC using the individual C18 column, whereas the combination of C18 × CN did not result in any improvement of the separation performance.  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