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.
Thin films of AgSbS2 are important for phase‐change memory applications. This solid is deposited by various techniques, such as metal organic chemical vapour deposition or laser ablation deposition, and the structure of AgSbS2(s), as either amorphous or crystalline, is already well characterized. The pulsed laser ablation deposition (PLD) of solid AgSbS2 is also used as a manufacturing process. However, the processes in plasma have not been well studied. We have studied the laser ablation of synthesized AgSbS2(s) using a nitrogen laser of 337 nm and the clusters formed in the laser plume were identified. The ablation leads to the formation of various single charged ternary AgpSbqSr clusters. Negatively charged AgSbS, AgSb2S, AgSb2S, AgSb2S and positively charged ternary AgSbS+, AgSb2S+, AgSb2S, AgSb2S clusters were identified. The formation of several singly charged Ag+, Ag, Ag, Sb, Sb, S ions and binary AgpSr clusters such as AgSb, Ag3S?, SbS (r = 1–5), Sb2S?, Sb2S, Sb3S (r = 1–4) and AgS, SbS+, SbS, Sb2S+, Sb2S, Sb3S (r = 1–4), AgSb was also observed. The stoichiometry of the clusters was determined via isotopic envelope analysis and computer modeling. The relation of the composition of the clusters to the crystal structure of AgSbS2 is discussed. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

2.
The characteristic fragmentations of a pTyr group in the negative ion electrospray mass spectrum of the [M–H]? anion of a peptide or protein involve the formation of PO (m/z 79) and the corresponding [(M‐H)?–HPO3]? species. In some tetrapeptides where pTyr is the third residue, these characteristic anion fragmentations are accompanied by ions corresponding to H2PO and [(M‐H)?–H3PO4]? (these are fragmentations normally indicating the presence of pSer or pThr). These product ions are formed by rearrangement processes which involve initial nucleophilic attack of a C‐terminal ‐CO [or ‐C(?NH)O?] group at the phosphorus of the Tyr side chain [an SN2(P) reaction]. The rearrangement reactions have been studied by ab initio calculations at the HF/6‐31+G(d)//AM1 level of theory. The study suggests the possibility of two processes following the initial SN2(P) reaction. In the rearrangement (involving a C‐terminal carboxylate anion) with the lower energy reaction profile, the formation of the H2PO and [(M‐H)?–H3PO4]? anions is endothermic by 180 and 318 kJ mol?1, respectively, with a maximum barrier (to a transition state) of 229 kJ mol?1. The energy required to form H2PO by this rearrangement process is (i) more than that necessary to effect the characteristic formation of PO from pTyr, but (ii) comparable with that required to effect the characteristic α, β and γ backbone cleavages of peptide negative ions. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

3.
Measurements of the translational energy loss accompanying the charge-stripping reactions M++N→M2++N+e and M2++N→M3++N+e have been performed for C, C and C, C respectively. The energy nesessary to remove the second electron from Buckminsterfullerene was determined, Q=IE(C→C=12.25±0.5 eV.  相似文献   

4.
The interaction of the palladium(II) complex [Pd(hzpy)(H2O)2]2+, where hzpy is 2‐hydrazinopyridine, with purine nucleoside adenosine 5′‐monophosphate (5′‐AMP) was studied kinetically under pseudo‐first‐order conditions, using stopped‐flow techniques. The reaction was found to take place in two consecutive reaction steps, which are both dependent on the actual 5′‐AMP concentration. The activation parameters for the two reaction steps, i.e. ΔH = 32 ±2 kJ mol?1, ΔS = ?168 ±7 J K?1 mol?1, and ΔH = 28 ± 1 kJ mol?1, ΔS = ?126 ± 5 J K?1 mol?1, respectively, were evaluated and suggested an associative mode of activation for both substitution processes. The stability constants and the associated speciation diagram of the complexes were also determined potentiometrically. The isolated solid complex was characterized by C, H, and N elemental analyses, IR, magnetic, and molar conductance measurements. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 42: 132–142, 2010  相似文献   

5.
The kinetics of formation and dissociation of [V(H2O)5NCS]2+ have been studied, as a function of excess metal-ion concentration, temperature, and pressure, by the stopped-flow technique. The thermodynamic stability of the complex was also determined spectrophotometrically. The kinetic and equilibrium data were submitted to a combined analysis. The rate constants and activation parameters for the formation (f) and dissociation (r) of the complex are: k/M ?1 · S?1 = 126.4, k/s?1 = 0.82; ΔH /kJ · mol?1 = 49.1, ΔH/kJ · mol?1 = 60.6; ΔS/ J·K?1·mol?1= ?39.8, ΔSJ·K?1·mol?1 = ?43.4; ΔV/cm3·mol?1 = ?9.4, and ΔV/cm3 · mol?1 =?17.9. The equilibrium constant for the formation of the monoisothiocynato complex is K298/M ?1 = 152.9, and the enthalpy and entropy of reaction are ΔH0/kJ · mol?1 = ? 11.4 and ΔS0/J. K?1mol?1 = +3.6. The reaction volume is ΔV0/cm3· mol?1 = +8.5. The activation parameters for the complex-formation step are similar to those for the water exchange on [V(H2O)6]3+ obtained previously by NMR techniques. The activation volumes for the two processes are consistent with an associative interchange, Ia, mechanism.  相似文献   

6.
BaSO4 precipitated from mixed salt solutions by common techniques for SO isotopic analysis may contain quantities of H2O and NO that introduce errors in O isotope measurements. Experiments with synthetic solutions indicate that δ18O values of CO produced by decomposition of precipitated BaSO4 in a carbon reactor may be either too low or too high, depending on the relative concentrations of SO and NO and the δ18O values of the H2O, NO, and SO. Typical δ18O errors are of the order of 0.5 to 1‰ in many sample types, and can be larger in samples containing atmospheric NO, which can cause similar errors in δ17O and Δ17O. These errors can be reduced by (1) ion chromatographic separation of SO from NO, (2) increasing the salinity of the solutions before precipitating BaSO4 to minimize incorporation of H2O, (3) heating BaSO4 under vacuum to remove H2O, (4) preparing isotopic reference materials as aqueous samples to mimic the conditions of the samples, and (5) adjusting measured δ18O values based on amounts and isotopic compositions of coexisting H2O and NO. These procedures are demonstrated for SO isotopic reference materials, synthetic solutions with isotopically known reagents, atmospheric deposition from Shenandoah National Park, Virginia, USA, and sulfate salt deposits from the Atacama Desert, Chile, and Mojave Desert, California, USA. These results have implications for the calibration and use of O isotope data in studies of SO sources and reaction mechanisms. Published in 2008 by John Wiley & Sons, Ltd.  相似文献   

7.
The haloacetyl ions [XCH2CO+], X = Cl, Br, I, have been produced by the dissociative ionization of selected precursor molecules and identified via their fragmentation characteristics. Their heats of formation, Δ, were 708 ± 8, 750 ± 25 and 784 ± 10 kJ mol?1, respectively, all appreciably above ΔH for [CH3CO+], 653 kJ mol?1. The effect of halogen substitution α to carbonyl is discussed for neutral molecules and their molecular ions.  相似文献   

8.
The equilibrium constants of the reactions MBr2(s) + Al2Br6(sln) ? MAl2Br8(sln) M = Cr, Mn, Co, Ni, Zn, Cd have been measured at 298 K in toluene. Ni: 0.017 ± 0.0024, Co: 0.54 ± 0.07, Zn: 1.5 ± 0.2, Mn: 2.1 ± 0, 7, Cr: 2.2 ± 1, Cd: 7 ± 5. They are compared with literature values of the equilibrium constants of analogous reactions in the gas phase MX2(s) + Al2X6(g) ? MAl2X8(g), X = Cl, Br. For CoAl2Br8(sln) the temperature dependence of the equilibrium constant yielded ΔfH = ?9.4 ± 1 kJ mol?1 and ΔfS = ?39.5 ± 3 J mol?1 K?1 while literature values for CoAl2Br8(g) are ΔfH = 42.4 ± 2 kJ mol?1 and ΔfS = 42.9 ± 2 J mol?1 K?1. The solubility of Al2Br6 in toluene as well as its enthalpy of dissolution have been measured in order to evaluate ΔH° and ΔS° of the solvation of Al2Br6(g) and CoAl2Br8(g) in toluene by a thermodynamic cycle. Solvation of Al2Br6(g): ΔH = ?72.7 ± 1 kJ mol?1, ΔS = ?139.6 ± 4 J mol?1 K?1, solvation of CoAl2Br8(g): ΔH = ?124.5 ± 4kJ mol?1, ΔS = ?222 ± 9J mol?1 K?1. Thus, CoAl2Br8 interacts more strongly with the solvent toluene than Al2Br6 does.  相似文献   

9.
Kinetics of the complex formation of chromium(III) with alanine in aqueous medium has been studied at 45, 50, and 55°C, pH 3.3–4.4, and μ = 1 M (KNO3). Under pseudo first-order conditions the observed rate constant (kobs) was found to follow the rate equation: Values of the rate parameters (kan, k, KIP, and K) were calculated. Activation parameters for anation rate constants, ΔH(kan) = 25 ± 1 kJ mol?1, ΔH(k) = 91 ± 3 kJ mol?1, and ΔS(kan) = ?244 ± 3 JK?1 mol?1, ΔS(k) = ?30 ± 10 JK?1 mol?1 are indicative of an (Ia) mechanism for kan and (Id) mechanism for k routes (‥substrate Cr(H2O) is involved in the k route whereas Cr(H2O)5OH2+ is involved in k′ route). Thermodynamic parameters for ion-pair formation constants are found to be ΔH°(KIP) = 12 ± 1 kJ mol?1, ΔH°(K) = ?13 ± 3 kJ mol?1 and ΔS°(KIP) = 47 ± 2 JK?1 mol?1, and ΔS°(K) = 20 ± 9 JK?1 mol?1.  相似文献   

10.
The reactions of dimethyl ether (CH3OCH3, DME) with O(3P) and H atoms have been studied at high temperatures by using a shock tube apparatus coupled with atomic resonance absorption spectroscopy (ARAS). The rate coefficients for the reactions CH3OCH3 + O(3P) → CH3OCH2 + OH (1) and CH3OCH3 + H → CH3OCH2 + H2 (2) were experimentally determined from the decay of O(3P) and H atoms as: These results show that DME can react with O(3P) atoms more easily than with H atoms. By combining these results with the previous lower temperature data, we obtained the following modified Arrhenius expressions applied over the wide temperature range between 300 and 1500 K: Both reactions of DME are faster than those of ethane, because the dissociation energy of the C? H bond in DME is smaller. Furthermore, the rate coefficients for reactions ( 1 ) and ( 2 ) were calculated with the transition‐state theory (TST). Structural parameters and vibrational frequencies of the reactants and the transition states required for the TST calculation were obtained from the MP2(full)/6‐31G(d) ab initio molecular orbital (MO) calculation. The energy barrier, E?0, was adjusted until the TST rate coefficient most closely matched the observed one. The fitting results of E(1) = 23 kJ mol?1 and E(2) = 34 kJ mol?1 were in agreement with the G2 energy barriers, within the expected uncertainty, demonstrating that the experimentally determined rate coefficients were theoretically valid. © 2006 Wiley Periodicals, Inc. 39: 97–108, 2007  相似文献   

11.
The substituted thiourea, 4‐methyl‐3‐thiosemicarbazide, was oxidized by iodate in acidic medium. In high acid concentrations and in stoichiometric excess of iodate, the reaction displays an induction period followed by the formation of aqueous iodine. In stoichiometric excess of methylthiosemicarbazide and high acid concentration, the reaction shows a transient formation of aqueous iodine. The stoichiometry of the reaction is: 4IO + 3CH3NHC(S)NHNH2 + 3H2O → 4I + 3SO + 3CH3NHC(O)NHNH2 + 6H+ (A). Iodine formation is due to the Dushman reaction that produces iodine from iodide formed from the reduction of iodate: IO + 5I + 6H+ → 3I2(aq) + 3H2O (B). Transient iodine formation is due to the efficient acid catalysis of the Dushman reaction. The iodine produced in process B is consumed by the methylthiosemicarbazide substrate. The direct reaction of iodine and methylthiosemicarbazide was also studied. It has a stoichiometry of 4I2(aq) + CH3NHC(S)NHNH2 + 5H2O → 8I + SO + CH3NHC(O)NHNH2 + 10H+ (C). The reaction exhibits autoinhibition by iodide and acid. Inhibition by I is due to the formation of the triiodide species, I, and inhibition by acid is due to the protonation of the sulfur center that deactivates it to further electrophilic attack. In excess iodate conditions, the stoichiometry of the reaction is 8IO + 5CH3NHC(S)NHNH2 + H2O → 4I2 + 5SO + 5CH3NHC(O)NHNH2 + 2H+ (D) that is a linear combination of processes A and B. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 193–203, 2000  相似文献   

12.
Protonated benzene, C6H, has been studied extensively to understand the structure and energy of a protonated organic molecule in the gas phase. The formation of C6H is either through direct protonation of benzene, i.e., chemical ionization, or through fragmentation of certain radical cations produced from electron ionization or photon ionization. We report a novel observation of C6H as a product ion formed in the collision‐induced dissociation (CID) of protonated benzamide and related molecules produced via electrospray ionization (ESI). The formation of C6H from these even‐electron precursor ions during the CID process, which has not been previously reported, is proposed to occur from the protonated molecules via a proton migration in a five‐membered ring intermediate followed by the cleavage of the mono‐substituent C? C bond and concurrent formation of an ion‐molecule complex. This unique mechanism has been scrutinized by examining some deuterated molecules and a series of structurally related model compounds. This finding provides a convenient mean to generate C6H, a reactive intermediate of considerable interest, for further physical or chemical investigation. Further studies indicate that the occurrence of C6H in liquid chromatography/electrospray ionization tandem mass spectrometry (LC/ESI‐MS/MS) appears to be a rather common phenomenon for many compounds that contain ‘benzoyl‐type’ moieties. Hence, the observation of the C6H ion in LC/ESI‐MS/MS can be used as an informative fragmentation pathway which should facilitate the identification of a great number of compounds containing the ‘benzoyl‐type’ and similar structural features. These compounds are frequently present in food and pharmaceutical products as leachable impurities that require strict control and rapid elucidation of their identities. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

13.
  • 1. The anions CH3O‐CO and CH3OCO‐CO are both methoxide anion donors. The processes CH3O‐CO → CH3O + CO and CH3OCO—CO → CH3O + 2CO have ΔG values of +8 and ?68 kJ mol?1, respectively, at the CCSD(T)/6‐311++G(2d, 2p)//B3LYP/6‐311++G(2d,2p) level of theory.
  • 2. The reactions CH3OCOCO → CH3OCO + CO (ΔG = ?22 kJ mol?1) and CH3COCH(O)CO2CH3 → CH3COCH(O)OCH3 + CO (ΔG = +19 kJ mol?1) proceed directly from the precursor anions via the transition states (CH3OCO…CO2) and (CH3COCHO…CH3OCO), respectively.
  • 3. Anion CH3COCH(O)CO2CH3 undergoes methoxide anion transfer and loss of two molecules of CO in the reaction sequence CH3COCH(O)CO2CH3 → CH3CH(O)COCO2CH3 → [CH3CHO (CH3OCO‐CO)] → CH3CH(O)OCH3 + 2CO (ΔG = +9 kJ mol?1). The hydride ion transfer in the first step is a key feature of the reaction sequence.
Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

14.
The enthalpy of formation of tris(methylidene)-cyclopropane (“[3]radialene”, 1 ) has been determined as ΔH = 396 ± 12 kJ mol?1 from three fragmentation reactions of its molecular ion 1 + formed from 1 by photoionisation using synchrotron radiation. Comparative electron impact measurements using conventional mass spectrometry were also performed. A treatment of the latter data is described which leads to satisfactory agreement with the photoionization data. The experimental value of ΔH( 1 ) is compared with various theoretical estimates. The strain energy of 1 is calculated to be 226.3 kJ mol?1. Linear extrapolation of this quantity from the increase of strain in passing from cyclopropane to methylidenecyclopropane yields 282.4 kJ mol?1. The discrepancy between these values, already predicted by Dewar and Baird ten years ago from theoretical calculations, is discussed on the basis of maximum overlap considerations. The enthalpy of formation of bis(methylidene)cyclopropane is predicted to be ΔH= 309 kJ mol?1.  相似文献   

15.
Ternary chalcogenide As‐S‐Se glasses, important for optics, computers, material science and technological applications, are often made by pulsed laser deposition (PLD) technology but the plasma composition formed during the process is mostly unknown. Therefore, the formation of clusters in a plasma plume from different glasses was followed by laser desorption ionization (LDI) or laser ablation (LA) time‐of‐flight mass spectrometry (TOF MS) in positive and negative ion modes. The LA of glasses of different composition leads to the formation of a number of binary AspSq, AspSer and ternary AspSqSer singly charged clusters. Series of clusters with the ratio As:chalcogen = 3:3 (As3S, As3S2Se+, As3SSe), 3:4 (As3S, As3S3Se+, As3S2Se, As3SSe, As3Se), 3:1 (As3S+, As3Se+), and 3:2 (As3S, As3SSe+, As3Se), formed from both bulk and PLD‐deposited nano‐layer glass, were detected. The stoichiometry of the AspSqSer clusters was determined via isotopic envelope analysis and computer modeling. The structure of the clusters is discussed. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

16.
Dibromomethylsulfoniumsalts — Preparation and Crystal Structure The salts CH3SBrA? (A? = SbCl, AsF) were prepared by various routes and characterized by their Ramanspectra. CH3SBrAsF crystallized in the monoclinic space group P21/c with a = 770,5(4) pm, b = 942,4(12) pm, c = 1329,3(14) pm, β = 100,28(6)°, Z = 4. Distances and bond angles in the cation are as expected.  相似文献   

17.
Nitrogen (N) and oxygen (O) isotope ratios of NO are often used to trace dominant NO pollution sources in water. Both the silver nitrate (AgNO3) method and the bacterial denitrification method are frequently used analytical techniques to determine δ15N‐ and δ18O‐NO in aqueous samples. The AgNO3 method is applicable for freshwater and requires a concentration of 100–200 µmol of NO for isotope determination. The bacterial denitrification method is applicable for seawater and freshwater and for KCl extracts of soils with a NO concentration as low as 1 µmol. We have carried out a thorough method comparison using 42 real surface water samples having a wide range of δ15N‐ and δ18O‐NO values and NO concentrations. Various correction pairs using three international references and blanks were used to correct raw δ15N‐ and δ18O‐NO values. No significant difference between the corrected data was observed when using various correction pairs for each analytical method. Both methods also showed excellent repeatability with high intraclass correlation coefficients (ICC). The ICC of the AgNO3 method was 0.992 for δ15N and 0.970 for δ18O. The ICC of the bacterial denitrification method was 0.995 for δ15N and 0.954 for δ18O. Moreover, a positive linear relationship with a high correlation coefficient (r ≥ 0.88) between the two methods was found for δ15N‐ and δ18O‐NO. The comparability of the methods was assessed by the Bland‐Altman technique using 95% limits of agreement. The average difference between results obtained by the bacterial denitrification and the AgNO3 method for δ15N was ?1.5‰ with 95% limits of agreement ?3.6 and +0.5‰. For δ18O this was +2.0‰, with 95% limits of agreement ?3.3 and +7.3‰. We found that for δ15N and for δ18O, 97% of the differences fell within these 95% limits of agreement. In conclusion, the AgNO3 and the bacterial denitrification methods are highly correlated and statistically interchangeable. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

18.
Some newly synthesized 10B nido‐carborane derivatives, i.e., 7,8‐dicarba‐nido‐undecaborane monoanions ([7‐Me‐8‐R‐C2B9H10]K+, R = H, butyl, hexyl, octyl and decyl), have been fully characterised and examined by electrospray ionization and Fourier transform ion cyclotron resonance mass spectrometry with liquid chromatographic separation (LC/ESI‐FTICR‐MS). These boron‐containing compounds exhibit abundant molecular ions ([M]?) at m/z 140.22631 [CB9H14]?, m/z 196.28883 [CB9H22]?, m/z 224.32032 [CB9H26]?, m/z 252.35133 [CB9H30]? and m/z 280.38354 [CB9H34]? at the normal tube lens voltage setting of ?90 V, which was an instrumental parameter value selected in the tuning operation. Additional [M–nH2]? (n = 1?4) ions were observed in the mass spectra when higher tube lens voltages were applied, i.e., ?140 V. High‐resolution FTICR‐MS data revealed the accurate masses of fragment ions, bearing either an even or an odd number of electrons. Collision‐induced dissociation of the [M–nH2]? ions (n = 0–4) in the quadrupole linear ion trap (LTQ) analyzer confirmed the loss of hydrogen molecules from the molecular ions. It is suggested that the loss of H2 molecules from the alkyl chain is a consequence of the stabilization effect of the nido‐carborane charged polyhedral skeleton. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

19.
Nitrogenous materials can be transferred out of the topsoil, either vertically to a greater depth, or in lateral pathways to surface waters, and they may also become transformed, with the potential of generating environmentally active agents. We measured the production of NO and N2O in two contrasting subsoils (70 to 90 cm): one poorly drained and the other freely drained and compared this with the topsoil (0 to 20 cm) of the corresponding soils. The soils were incubated aerobically in jars with subtreatments of either synthetic cattle urine or deionised water and sampled at intervals up to 34 days. 15N‐NO was used to determine the processes responsible for NO and N2O production. The headspace was analysed for the concentrations of N2O, NO and CO2 and 15N enrichment of N2O. The soil samples were extracted and analysed for NO, NO and NH, and the 15N enrichment of the extracts was measured after conversion into N2O and N2. The study demonstrated the potential for NO, N2O and NO to be generated from subsoils in laboratory incubations. Differences in these N dynamics occurred due to subsoil drainage class. In the freely drained subsoil the rates of NO and NO production were higher than those observed for the corresponding topsoil, with mean maximum production rates of 3.5 µg NO‐N g−1 dry soil on day 16 and 0.12 µg NO‐N g−1 dry soil on day 31. The calculated total losses of N2O‐N as percentages of the applied synthetic urine N were 0.37% (freely drained subsoil), 0.24% (poorly drained subsoil), 0.43% (freely drained topsoil) and 2.09% (poorly drained topsoil). The calculated total losses of NO‐N as percentages of the applied synthetic urine N were 1.53% (freely drained subsoil), 0.02% (poorly drained subsoil), 0.25% (freely drained topsoil) and 0.08% (poorly drained topsoil). Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

20.
Multipulse pulsed laser polymerization coupled with size exclusion chromatography (MP‐PLP‐SEC) has been employed to study the depropagation kinetics of the sterically demanding 1,1‐disubstituted monomer di(4‐tert‐butylcyclohexyl) itaconate (DBCHI). The effective rate coefficient of propagation, k, was determined for a solution of monomer in anisole at concentrations, c, 0.72 and 0.88 mol L?1 in the temperature range 0 ≤ T ≤ 70 °C. The resulting Arrhenius plot (i.e., ln k vs. 1/RT) displayed a subtle curvature in the higher temperature regime and was analyzed in the linear part to yield the activation parameters of the forward reaction. In the temperature region where no depropagation was observed (0 ≤ T ≤ 50 °C), the following Arrhenius parameters for kp were obtained (DBCHI, Ep = 35.5 ± 1.2 kJ mol?1, ln Ap = 14.8 ± 0.5 L mol?1 s?1). In addition, the k data was analyzed in the depropagatation regime for DBCHI, resulting in estimates for the associated entropy (?ΔS = 150 J mol?1 K?1) of polymerization. With decreasing monomer concentration and increasing temperature, it is increasingly more difficult to obtain well structured molecular weight distributions. The Mark Houwink Kuhn Sakurada (MHKS) parameters for di‐n‐butyl itaconate (DBI) and DBCHI were determined using a triple detection GPC system incorporating online viscometry and multi‐angle laser light scattering in THF at 40 °C. The MHKS for poly‐DBI and poly‐DBCHI in the molecular weight range 35–256 kDa and 36.5–250 kDa, respectively, were determined to be KDBI = 24.9 (103 mL g?1), αDBI = 0.58, KDBCHI = 12.8 (103 mL g?1), and αDBCHI = 0.63. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 1931–1943, 2007  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