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.
Conformational analysis using molecular mechanics (MM) was performed for a determination of the stereochemistry of serricornin, the sex pheromone of the cigarette beetle (Lasioderma serricorne F.). An exhaustive conformational analysis using MM2 calculations with algorithms for covering torsional energy surfaces of flexible molecules furnishes coordinates and steric energies of all local energy minimum conformers of serricornin, both acyclic and the corresponding cyclic forms. These coordinates gave angles required for the calculation of vicinal H/H coupling constants (3 J HHs) of each energy minima by Altona's modified Karplus equation. The Boltzmann distributions of all local energy minima were calculated from their steric energies to furnish populations of each energy minimum conformer. Populationweighted averaged3 J HHs of four enantiomeric pairs, (S *,S *,S *)-, (S *,S *,R *)-, (S *,R *,S *)-, and (R *,S *,S *)-serricornins were calculated from the data above. The observed3 J HHs of the naturally occurring serricornin, both acyclic and cyclic forms, are fitted best to calcd.3 J HHs of (4S *, 6S *, 7S *)-acyclic and (3S *, 5S *, 6S *)-cyclic serricornin, respectively, among those of four enantiomeric pairs of serricornin.  相似文献   

2.
Three N-alkyl bis-quaternary ammonium salt surfactants were synthesized by using epichlorohydrin, trimethylamine hydrochloride, and N,N-dimethylalkyl amine as raw materials in a two-step manner. The products were characterized by 1H NMR and MS, confirming the successful synthesis of 2-Hydroxy-N1,N1,N3,N3-tetramethyl-N3-dodecylpropane-1,3-diammonium chloride (HPDDC), 2-Hydroxy-N1,N1,N3,N3-tetramethyl-N3-tetradecylpropane-1,3-diammonium chloride (HPTDC), and 2-Hydroxy-N1, N1, N3, N3-tetramethyl-N3-hexadecylpropane-1,3-diammonium chloride (HPHDC). Moreover, the influence of carbon chain length on surface-active properties, foaming properties, and paraffin liquid emulsion stability was investigated. Results indicated that critical micelle concentrations (cmc) decreased with increasing carbon chain length from 12 to 16, and the cmc and γcmc were lower than those of Dodecyl trimethyl ammonium chloride (DTAC). The products exhibited better foam properties and worse emulsifying performance than those of DTAC. The Krafft points of all products were determined to be below 0 °C. Moreover, the products also demonstrated outstanding antibacterial properties.  相似文献   

3.
Four polynomial expressions are obtained that provide a good approximation and an easy, rapid calculation of the chromatic coordinates and the chroma—L *, a *, b *, and C—for the illuminant C and the standard observer, for a virgin or extra virgin olive oil; absorbance is measured at only 480 and 670 nm. These are as follows: L *=0.556458(A480)2−2.51145A480+0.55504(A670)2−8.53016A670+98.4089; a *=0.177372(A480)2+2.1363A480+1.43254(A670)2−0.789231A670−13.9246; b *=−16.0277(A480)2+79.8932A480−5.06558(A670)2+3.36169A670+31.9405; C=−15.8439(A480)2+78.9312A480−5.26784(A670)2+3.56917A670+33.3927. These give acceptable results, making the method a practical alternative to the extremely laborious Commission Internationale d’Eclairage (CIE) L * a * b * system, by which 391 absorbance values must be measured individually, nanometer by nanometer, before applying more complex equations. The validity of the proposed method has been confirmed by comparison, using a set of 20 sample oils different from the set of 25 oils used to generate the order of the equations. The variations between the values provided by the proposed and standard methods, respectively, had a mean of 0.00 for each of the chromatic variables—L * , a * , b * , and C; SD were moderate (0.71, 0.52, 1.22, and 1.22, respectively); the root mean square and the R 2-terms also confirmed the validity of the method.  相似文献   

4.
Crossing experiments between two closely related moths, Ostrinia scapulalis and O. zealis, were conducted to gain insight into the genetic basis of the divergence of female sex pheromones. The sex pheromone of O. scapulalis comprises (E)-11- and (Z)-11-tetradecenyl acetates (E11 and Z11), and distinct genetic variation is found in the blend of components. This variation is largely controlled by a single autosomal locus with two alleles, AE(sca) and AZ(sca). E-type (AE(sca)AE(sca)) females produce a pheromone with amean E11:Z11 ratio of 99:1, whereas Z-type (AZ(sca)AZ(sca)) and I-type (AE(sca)AZ(sca)) females produce a pheromone with a mean of 3:97 and 64:36, respectively. O. zealis is distinctive in that it has a third pheromone component, (Z)-9-tetradecenyl acetate (Z9), in addition to E11 and Z11, and the typical blend ratio is 60:35:5 (Z9:E11:Z11). Our study revealed that Z9 production in O. zealis is mainly regulated by an autosomal recessive gene phr(zea), which is suggested to be involved in the chain-shortening of a pheromone precursor fatty acid, and linked to AE(zea), a gene corresponding to AE(sca) in O. scapulalis. A few mutations in a gene involved in pheromone production could explain the dramatic shift between a two-component pheromone communication system in O. scapulalis and a three-component system in O. zealis.  相似文献   

5.
The Faraday effects of Ge‐Ga‐Sb(In)‐S serial chalcogenide glasses were investigated at the wavelengths of 635, 808, 980, and 1319 nm, respectively. The compositional dependences were analyzed and associated influencing factors including the absorption edge, the concentration of Sb3+/In3+ ions, and the wavelength dispersion of refraction index were discussed. 80GeS2·20Sb2S3 composition glass was found to have the largest Verdet constant (V=0.253, 0.219, 0.149, and 0.065 min·G?1·cm?1 for wavelengths 635, 808, 980, and 1319 nm, respectively) in these glasses, which is larger than that of commercial diamagnetic glasses (Schott, SF 6, V=0.069 min·G?1·cm?1@633 nm, for example). Sb3+ ions with high polarizability possessing s2‐sp electron jumps involving 1S01P1, 3P0,1,2 transitions are responsible for large Verdet constant, and Becquerel rule is proved to be an effective guidance for estimating the Verdet constant and further optimizing the compositions in chalcogenide glasses.  相似文献   

6.
The phase relationships in binary, ternary, and more complex Me 2 +O–Me 2+O–Me 2 3+O3Me 4+O2–TiO2 systems (Me + = Li+, K+, Rb+, Cs+; Me 2+ = Mg2+, Sr2+, Ba2+, Zn2+; Me 3+ = Al3+, Fe3+, Ga3+; Me 4+ = Sn4+, Zr4+) are investigated in the concentration regions corresponding to the compositions of titanates with a tunnel structure: Li2(Me 2+,Me 3+) y (Me 4+,Ti)4O8 ramsdellites, (Me +,Me 2+) x (Me 2+,Me 3+) y (Me 4+,Ti)8O16 hollandites, and (Ba,Me 2+)2(Me 4+,Ti)9O20 phases. The homogeneity regions of the solid solutions with the above structures are determined, and their crystal chemical characteristics, phase transformations, and thermal and electrical properties are studied. The results obtained and the data available in the literature are analyzed and generalized. The general approaches to the prediction of changes in the structure and properties of the studied titanates with a variation in the chemical composition due to isomorphous substitutions in different structural positions of the crystal lattice are discussed.  相似文献   

7.
To introduce the 3‐[18F]fluoro‐2‐hydroxypropyl moiety into positron emission tomography (PET) radiotracers, we performed automated synthesis of (rac)‐, (R)‐, and (S)‐[18F]epifluorohydrin ([18F] 1 ) by nucleophilic displacement of (rac)‐, (R)‐, or (S)‐glycidyl tosylate with 18F? and purification by distillation. The ring‐opening reaction of (R)‐ or (S)‐[18F] 1 with phenol precursors gave enantioenriched [18F]fluoroalkylated products without racemisation. We then synthesised (rac)‐, (R)‐, and (S)‐ 2‐{5‐[4‐(3‐[18F]fluoro‐2‐hydroxypropoxy)phenyl]‐2‐oxobenzo[d]oxazol‐3(2H)‐yl}‐N‐methyl‐N‐phenylacetamide ([18F] 6 ) as novel radiotracers for the PET imaging of translocator protein (18 kDa) and showed that (R)‐ and (S)‐[18F] 6 had different radioactivity uptake in mouse bone and liver. Thus, (rac)‐, (R)‐, and (S)‐[18F] 1 are effective radiolabelling reagents and can be used to develop PET radiotracers by examining the effects of chirality on their in vitro binding affinities and in vivo behaviour.  相似文献   

8.
The model of pseudocrosslink is extended to polymer rheology. There exist various sizes b of 4–16 between transition points A and B. Each link is connected with a chain having length nb and relaxation time τb. nb is equal to b2 and τb is proportional to nb2. A, B, and C (polymer terminal) divide the stress-relaxation spectrum into four zones. In the AB zone, successive dissociation of links occurs from a small size to a large one and rigidity G is decreased with time t as G ∝? t?0.5 and viscosity η is increased as G ∝? t0.5. In the BC zone, dissociation of the B link proceeds along a molecule of length n in a mode of squeezing of molecule and η becomes constant, but G still decreases due to increase of unperturbed end-to-end distance of chain and G ∝? t?0.5. However, dynamic elasticity becomes constant due to a small amplitude. At high shear, links are loosened and G and η are much decreased. Beyond C, molecule flows and η increases as η ∝? γ?0.8 n3.5, but high shear rate γ diminishes the effect of n due to extension of the molecule. Extensional viscosity η* is affected by the change of shape as η* ∝? t1.5 and gives an overshoot. Under load, creep occurs and it is proportional to t1/2–1/3. © 1995 John Wiley & Sons, Inc.  相似文献   

9.
This paper reports on the preparation and spectral properties of europium (Eu3+) and terbium (Tb3+) ions doped cadmium lead boro tellurite (CLBT) glasses. For reference glasses, physical properties have been evaluated. From the [measurements of X-ray diffraction (XRD), glass amorphous nature of these [glasses has been studied. From the emission spectra of Eu3+: CLBT glasses, five [transitions (5 D 07 F 0, 7 F 1, 7 F 2, 7 F 3 and 7 F 4) at 579, 591, 613, 652 and 701 nm are observed with λexci = 392 nm (7 F 05 L 6) and in the case of Tb3+: CLBT glasses, four emission transitions 5 D 4 → (7 F 6, 7 F 5, 7 F 4 and 7 F 3) are observed at 489, 543, 584 and 621nm respectively monitered with λexci = 376 nm (7 F 65 G 6).  相似文献   

10.
Sodium ions spiked with 22Na as a tracer were migrated by electromigration and electro-osmosis in the water-saturated compacted Na-montmorillonite at dry densities 1.0×103 kg m−3, under an electric potential gradient. Dissolved helium was also migrated by electro-osmosis in the montmorillonite. After migration, concentration profiles of the sodium ions and helium were obtained by γ-spectrometry and mass-spectrometric methods, respectively. From the profiles of both chemical species, not only migration due to electrokinetic phenomena but also mechanical dispersion was observed in the montmorillonite. The dispersion coefficients, Di, and apparent migration rates, Uia, of 22Na and helium were found in the compacted Na-montmorillonite at 1.0×103 kg m−3. The migration of helium in the montmorillonite under an electric potential gradient reflects that of water because helium migrates as an electrically neutral species. The parameters DHem, UHea, and αHe correspond to those of water. The mechanical dispersion coefficients, DNam, of 22Na+ ions are much smaller than those of water obtained by helium. The dispersivity parameters, αNa, for 22Na+ obtained from these DNa and UNaa values are 10−5 m and those for water (αHe) are 10−3 m. This indicates that 22Na+ ions migrate in different spaces than water in the compacted montmorillonite under a potential gradient. This finding suggests that the migration of Na+ ions occurs in the interlayer and/or on the outer surfaces of the montmorillonite; whereas dissolved helium migrates in the pore water.  相似文献   

11.
Bare, cylindrical, explosive charges produce secondary shock waves in the direction of least presented area. Whilst the source of these shock waves was explored in the 1940’s, no attempt was made to predict them. This paper describes the detonation of bare, cylindrical charges of PE4 (RDX binder 88/12 %), mass 0.2 to 0.46 kg and with a length to diameter ratio of 4 to 1. High speed camera footage showed (i) the formation of the separate, primary, shock waves from the sides and ends of the charge, (ii) Mach reflection of these separate shock waves, giving rise to reflected, secondary shock waves, and (iii) the secondary shock waves catching and merging with the primary shock wave. In the axial direction, the secondary shock wave’s peak overpressure and impulse exceeded that of the primary shock wave for scaled distances, Z=R/M1/3 ≥3.9 m kg−1/3, where M is the mass in kg and R the distance from the charge in m. It was found possible to predict the primary peak overpressure, P, at all distances in the axial direction, for a constant length to diameter ratio, using P=3075 Z−3−1732 Z−2+305 Z−1. Close in the primary peak overpressure is proportional to M/R3 in the axial direction. It was not possible to predict the secondary peak overpressure with the data obtained. The total impulse from both shock waves, I, in the axial direction can be predicted using I=746(M2/3/R)3−708(M2/3/R)2+306(M2/3/R).  相似文献   

12.
LaScO3:xBi3+,yTb3+,zEu3+ (x = 0 − 0.04, y = 0 − 0.05, z = 0 − 0.05) phosphors were prepared via high-temperature solid-state reaction. Phase identification and crystal structures of the LaScO3:xBi3+,yTb3+,zEu3+ phosphors were investigated by X-ray diffraction (XRD). Crystal structure of phosphors was analyzed by Rietveld refinement and transmission electron microscopy (TEM). The luminescent performance of these trichromatic phosphors is investigated by diffuse reflection spectra and photoluminescence. The phenomenon of energy transfer from Bi3+ and Tb3+ to Eu3+ in LaScO3:xBi3+,yTb3+,zEu3+ phosphors was investigated. By changing the ratio of x, y, and z, trichromatic can be obtained in the LaScO3 host, including red, green, and blue emission with peak centered at 613, 544, and 428 nm, respectively. Therefore, two kinds of white light-emitting phosphors were obtained, LaScO3:0.02Bi3+,0.05Tb3+,zEu3+ and LaScO3:0.02Bi3+,0.03Eu3+,yTb3+. The energy transfer was characterized by decay times of the LaScO3:xBi3+, yTb3+, zEu3+ phosphors. Moreover absolute internal QY and CIE chromatic coordinates are shown. The potential optical thermometry application of LaScO3:Bi3+,Eu3+ was based on the temperature sensitivity of the fluorescence intensity ratio (FIR). The maximum Sa and Sr are 0.118 K−1 (at 473.15 K) and 0.795% K−1 (at 448.15 K), respectively. Hence, the LaScO3:Bi3+,Eu3+ phosphor is a good material for optical temperature sensing.  相似文献   

13.
5‐Aminotetrazolium nitrate was synthesized in high yield and characterized using Raman and multinuclear NMR spectroscopy (1H, 13C, 15N). The molecular structure of 5‐aminotetrazolium nitrate in the crystalline state was determined by X‐ray crystallography: monoclinic, P 21/c, a=1.05493(8) nm, b=0.34556(4) nm, c=1.4606(1) nm, β=90.548(9)°, V=0.53244(8) nm3, Z=4, ϱ=1.847 g cm−3, R1=0.034, wR2 (all data)=0.090. The thermal stability of 5‐aminotetrazolium nitrate was determined using differential scanning calorimetry; the compound decomposes at 167 °C. The enthalpy of combustion (ΔcombH) of 5‐aminotetrazolium nitrate ([CH4N5]+[NO3]) was determined experimentally using oxygen bomb calorimetry: ΔcombH([CH4N5]+[NO3])=−6020±200 kJ kg−1. The standard enthalpy of formation (ΔfH°) of [CH4N5]+[NO3] was obtained on the basis of quantum chemical computations at the electron‐correlated ab initio MP2 (second order Møller‐Plesset perturbation theory) level of theory using a correlation consistent double‐zeta basis set (cc‐pVTZ): ΔfH°([CH4N5]+[NO3](s))=+87 kJ mol−1=+586 kJ kg−1. The detonation velocity (D) and the detonation pressure (P) of 5‐aminotetrazolium nitrate were calculated using the empirical equations by Kamlet and Jacobs: D([CH4N5]+[NO3])=8.90 mm μs−1 and P([CH4N5]+[NO3])=35.7 GPa.  相似文献   

14.

The levels of PM 10 , PM 2.5 , and NO 2 were studied at a kerbsite and ambient site in Mumbai. Measurements were also made for eight inorganic ions (F ? , Cl ? , NO 3 ? , SO 2? 4 , Na + , K + , NH 4 + , Ca 2+ , and Mg 2+ ) in the PM 2.5 fraction. During the study period, PM 2.5 , PM 10 and NO 2 levels ranged between 11–91, 18–125, and 8–64 μ g m ? 3 at a ambient site whereas at the kerbsite the ranges were 10–176, 21–189, and 4–55 μ g m ?3 respectively. Average PM 2.5 values were 42 μ g m ? 3 at ambient and 69 μ g m ?3 at the kerbsite. The measured ions accounted for about 50% of the PM 2.5 mass. Non-sea-salt (nss) sulfate contributed 91% and 85% of the ionic mass at the ambient and kerbsite sites respectively. Due to biomass sources of K, only about 5% of K + was from seas salt. The average equivalent ratio of NH 4 + to nss- SO 2 4 ? , and NO 3 ? was over 1, indicating high source strength of ammonia.  相似文献   

15.
An ALGOL computer program has been devised to manipulate light-scattering data from the Brice-Phoenix photometer. The input consist of experimental values of the galvanometer deflections and filter factors used for each concentration c and angle of measurement θ. These are transformed to the appropriate variables in the fundamental equation including the particle scattering factor, viz: c/Qθ = (W/K*)M?w?1[1 + (16/3) × π2n12λ〈S2〉 sin2 (θ/2)] + (W/K*)2A2c + (W/K*)3A3c2 in which Qθ is a corrected from of the Rayleigh ratio and (W/K*) is a composite constant term for the instrument and polymer–solvent system. By writing X?ij for the variable c/Qθ at θi and cj, a function X is found by least squares to fit X?ij, thus X = l + m sin (θ/2) + ncj + bcj2. The equations arising from minimizing ΣiK=1 ΣjL=1 (Xij ? X?ij)2 are solved by the computer to yield the best-fitting coefficients l, m, n, and b. These can then be related simply to the molecular weight, root-mean-square radius of gyration, second and third virial coefficients, respectively. The final portion of the program is designed to check the fit of these coefficients. It yields a table of the differences between all experimental c/Qθ values and the coressponding ones obtained by inserting the derived l, m, n, and b into the fundamental equation. The procedure has been tested satisfactorily by using a well-standardized sample of polystyrene in toluene at 30°C. and a wavelength of 436 mμ.  相似文献   

16.
The kinetics of the polymerization of dimethyldiallylammonium chloride (DMDAAC) and acrylamide (AM) with different monomer molar ratios initiated by an ammonium persulfate–sodium bisulfate redox complex in an aqueous solution were studied. The polymerization rate (Rp) equation, the activation energy (Ea), and the reactivity ratio were measured. The results show that when the nDMDAAC:nAM values were 1 : 9, 2 : 8, 3 : 7, 4 : 6, and 5 : 5, the copolymerization rate equation were Rp1 = k[M]2.61[IO]0.51[IR]0.52, Rp2 = k[M]2.70[IO]0.50[IR]0.53, Rp3 = k[M]2.73[IO]0.50[IR]0.56, Rp4 = k[M]2.77[IO]0.51[IR]0.59, and Rp5 = k[M]2.84[IO]0.51[IR]0.61 (where [M] is the total monomer concentration, [IO] is the oxidant concentration, and [IR] is the reductant concentration), respectively when the temperature was 45°C. The Ea values were Ea1 = 79.10 kJ/mol, Ea2 = 81.39 kJ/mol, Ea3 = 85.15 kJ/mol, Ea4 = 88.88 kJ/mol, and Ea5 = 90.61 kJ/mol in the temperature range 35–55°C, respectively. The reactivity ratios of DMDAAC and AM were rDMDAAC = 0.14 and rAM = 6.11 when the temperature was 45°C. The structure of PDA was characterized by Fourier transform infrared spectroscopy and 1H-NMR. The results of the kinetic parameters explained the differences in the copolymerization rate and intrinsic viscosity of PDA with different cationicities. © 2012 Wiley Periodicals, Inc. J Appl Polym Sci, 2012  相似文献   

17.
Akihisa Mori  Yukio Imanishi 《Polymer》1984,25(12):1837-1844
An acrylamide derivative having a mono-N-methylated cyclic dipeptide as a substituent, c-(NεAcrLys-Sar), was synthesized and polymerized by radical polymerization, giving homopolymers and copolymers with styrene, 4-vinylpyridine, and N-dodecylacrylamide. The relative reactivities of these monomers in the radical polymerization decreased in the following order: styrene > 4-vinylpyridine > N-dodecylacrylamidec-(Nε-AcrLys-Sar). Poly[c-(Nε-AcrLys-Sar)] was soluble in water and the usual organic solvents except diethylether and n-hexane. In mixed solvent, poly[c-(Nε-AcrLys-Sar)] interacted most strongly with Ba2+ among alkali and alkaline-earth metal cations. A cyclic dipeptide, which is equivalent to the side chain of poly[c-(Nε-AcrLys-Sar)], did not interact with these cations, indicating that the side chain ligand groups of poly[c-(Nε-AcrLys-Sar)] interacts specifically with Ba2+ by intramolecular cooperation. However, the copolymer of c-(Nε-AcrLys-Sar) with styrene interacts more strongly with larger alkali cations, and equally well with Ba2+ and Ca2+. These results indicate that the styrene units in the copolymer influence its solubility and regulate the intramolecular cooperation of the cyclic dipeptide ligand groups as spacer groups. The copolymer of c-(Nε-AcrLys-Sar) with 4-vinylpyridine interacted very strongly with Ba2+ as a result of the intramolecular cooperation of a pyridyl group and a cyclic dipeptide group. The formation constant of the ternary complex was determined to be 109 M−3, which is larger by 102 fold than that by a polymer carrying benzo-15-crown-5 in the side chain. The copolymer of c-(Nε-AcrLys-Sar) with N-dodecylacrylamide was found to bind methyl orange in aqueous solution by hydrophobic interaction. The copolymer of c-(Nε-AcrLys-Sar) with styrene was found to extract phenylalanine from an aqueous solution to a n-octanol phase. However, the extraction was not enantiomer-selective.  相似文献   

18.
The structure of the orthorhombic phase in the MoVNbTeO propane ammoxidation catalyst system has been characterized and refined using a combination of TEM, synchrotron X-ray powder diffraction (S-XPD), and neutron powder diffraction (NPD). This phase, designated as M1 by Ushikubo et al. [1], crystallizes in the orthorhombic space group Pba2 (No. 32) with a = 21.134(2) Å, b = 26.658(2) Å, and c = 4.0146(3) Å. The formula unit is Mo7.5V1.5NbTeO29. Bond valence sum calculations indicate the presence of d 1 metal sites neighbored by d 0 metal sites. The d 1 sites are occupied by a distribution of Mo5+ and V4+, whereas the d 0 sites are occupied by a distribution of Mo6+ and V5+. Out-of-center distortions in d 0 octahedra are consistent with the second-order Jahn–Teller effect and lattice effects. We argue that the V5+–O–V4+/Mo5+ moieties adjacent to Te4+ and Mo6+ sites in the [001] terminal plane provide a spatially isolated active site at which the selective ammoxidation of propane occurs.  相似文献   

19.
An ozone reactor was constructed using a tubular gas diffuser made of microporous stainless steel to significantly reduce gas bubble size and increase overall mass transfer area. Overall mass transfer coefficient, KLa [s ?1], was correlated with gas (G) and liquid (L) flow rates using KLa = ALαGβ , with A = 3.96 × 10 8 [s?1], α = 1.53, and β = 0.40, with L and G in [m 3s?1]. The reactor is essentially plug flow at high G or L. This system achieves one of the highest ozone mass transfer rates observed in the literature.  相似文献   

20.
In this research, five different vegetable oils were oxidized at four different temperatures (373, 383, 393, and 403 K) under Rancimat test conditions. An increasing rate of oxidation could be observed as temperature increased. The natural logarithms of the kinetic rate constant (k value) varied linearly with respect to temperature, with the temperature coefficients (TCoeff) ranging from 6.95×10–2 to 7.40× 10–2 K–1 for the vegetable oils. On the basis of the Arrhenius equation and the activated complex theory, frequency factors (A), activation energies (Ea), Q10 numbers, activation enthalpies (ΔH++), and activation entropies (ΔS++) for oxidative stability of the vegetable oils were calculated. The A, Ea, Q10, ΔH++, and ΔS++ values for the vegetable oils ranged from 6.38×103 to 28.03×103 h–1, from 86.86 to 92.42 kJ/mol, from 2.08 to 2.18, from 83.64 to 89.20 kJ/mol, and from –116.66 to –104.35 J/mol K, respectively.  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