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.
The self-termination rates of the benzyl radical (C6H5---CH2) and para-substituted benzyl radicals (X---C6H4---CH2) were studied in aqueous solutions. The Arrhenius parameters and activation energies were determined in the temperature range 275.5–328 K. The kinetic activation energies of these radicals were close to the dynamic activation energy of the solvent, indicating that the termination rate is controlled by diffusion. The values for the rate constants (2kt (109 dm3 mol−1 s−1)) and the activation energies (E (kJ mol−1)) were 5.94±0.52 and 14.69±0.61 for CH3O---C6H4---CH2, 4.52±0.2 and 17.65±1.16 for CH37z.sbnd;C6H4---CH2, 3.07±0.45 and 17.58±0.97 for H---C6H4---CH2, 4.13±0.81 and 19.10±1.20 for Cl---C6H4---CH2 and 4.17±0.44 and 14.62±0.52 for NO2---C6H4---CH2.  相似文献   

2.
Pulse radiolysis of epicatechin in aqueous solution has been done to investigate the reactions of epicatechin derived phenoxy radical (EpO) at neutral pH. EpO was generated by N3 reacting toward EpOH, the rate constant was measured to be 3 × 108 dm3 mol−1 s−1. The biomolecular termination of EpO is rather slow ((2k < × 106 dm3 mol−1 s−1) and results in products exhibiting strong visible absorption around 450 nm. No reactions have been observed for EpO with O2 and O2 in the time scale of pulse radiolysis (0.01 s), suggesting the bimolecular rate constant are less than 104 and 5 × 106 dm3 mol−1 s−1, respectively.  相似文献   

3.
Pulse radiolysis technique has been employed to study the reactions of oxidizing (OH, N3) and reducing radicals (eaq, CO2√−, acetone ketyl radical) with 2-hydroxy-3-methoxybenzaldehyde (o-vanillin) at different pH. Hydroxyl radicals react mostly by addition reaction forming radical adducts (λmax=420 nm) and the oxidation is only a minor process even in the alkaline region. The reaction with azide radicals produced phenoxyl radicals (λmax=340 nm), which are formed on fast deprotonation of solute radical cation. Using PMZ√+/PMZ and ABTS√−/ABTS2− as the reference couple, different methods are employed to determine the one-electron reduction potential of o-vanillin and the average value is estimated to be 1.076±0.004 V vs. NHE at pH 6. The phenoxyl radicals of o-vanillin were able to oxidize ABTS2− quantitatively. The eaq is observed to react with o-vanillin with rate constant value of 2×1010 dm3 mol−1 s−1. CO2√− and acetone ketyl radical are also observed to react with o-vanillin by electron transfer mechanism and showed the formation of transient absorption bands with λmax at 350 and 390 nm at pH 4.5 and 9.7, respectively. The pKa of the one-electron reduced species was determined to be 8.1. The results indicate that the aldehydic group is the most preferred site for electron addition.  相似文献   

4.
Reactions of OH radicals and some one-electron oxidants with 2-aminopyridine (2-AmPy) and 3-aminopyridine (3-AmPy) were studied in aqueous solutions using pulse radiolysis technique. The OH adduct of 2-AmPy at pH 9 has an absorption maximum at 360 nm along with a weak absorption band in the visible region and was found to be reactive with oxygen. The rate constant for its reaction with O2 was determined to be 1.0×108 dm3 mol−1 s−1. At pH 4 also, the OH adduct of 2-AmPy has an absorption band at 360 nm. However, there are differences in the absorption at other wavelengths. From the plot of ΔOD vs. pH at 340 nm, the pKa of the OH adduct was determined to be 6.5. Among the specific oxidants, only SO4−√ radicals were able to oxidize 2-AmPy. In the case of 3-aminopyridine (3-AmPy), the transient species formed by OH radical reaction at pH 9 has an absorption maximum at 410 nm with shoulder bands on both the sides. Its absorption spectrum at pH 4 was different indicating the existence of a pK value for the OH adduct. pKa of 3-AmPy-OH radical adduct species was evaluated to be 5.7. This adduct species was also found to be reactive with oxygen (k=7.6×106 dm3 mol−1 s−1). Specific one-electron oxidants like N3, Br2−√ C2−√ and SO4−√ were able to oxidize 3-AmPy indicating that it is easier to oxidize 3-AmPy as compared to 2-AmPy.  相似文献   

5.
Using N3 species as specific electron acceptor a defined ascorbate radical: AH↔A+H+max=360 nm, =3400 dm3 mol−1 cm−1) is observed. The attack of DMSO+ on vit.E results in a vit.E radical (k=1×109 dm3 mol−1 s−1; λmax=425 nm, =2400 dm3 mol−1 cm−1; 2k=4.7×108 dm3 mol−1 s−1). Vit.E-acetate leads to the formation of a radical cation (vit.E-ac+). β-carotene reacts also with DMSO+ forming a radical cation, β-car+ (k=1.75×108 dm3 mol−1 s−1; λmax=942 nm, =14 600 dm3 mol−1 cm−1), which probably leads to the formation of a dimer radical cation, (β-car)+2 (k=2.5×107 dm3 mol−1 s−1).

Using E.coli bacteria (AB1157) as a model system in vitro it was found that all three vitamins are rather efficient radiation protecting agents. They can also increase the activity of cytostatica, e.g., mitomycin C (MMC), by electron transfer process. The mixture of vit.E-ac and β-car acts contradictory, but adding vit.C to it a strong cooperative enhancement of the MMC activity is observed once again. A relationship between the pulse radiolysis and the radiation biological data is found and discussed. A possible explanation of the previously reported trials concerning the role of vit.E and β-car on the increased occurence of lung and other types of cancer in smokers and drinkers is presented.  相似文献   


6.
Twenty-two isomers/conformers of C3H6S+√ radical cations have been identified and their heats of formation (ΔHf) at 0 and 298 K have been calculated using the Gaussian-3 (G3) method. Seven of these isomers are known and their ΔHf data are available in the literature for comparison. The least energy isomer is found to be the thioacetone radical cation (4+) with C2v symmetry. In contrast, the least energy C3H6O+√ isomer is the 1-propen-2-ol radical cation. The G3 ΔHf298 of 4+ is calculated to be 859.4 kJ mol−1, ca. 38 kJ mol−1 higher than the literature value, ≤821 kJ mol−1. For allyl mercaptan radical cation (7+), the G3 ΔHf298 is calculated to be 927.8 kJ mol−1, also not in good agreement with the experimental estimate, 956 kJ mol−1. Upon examining the experimental data and carrying out further calculations, it is shown that the G3 ΔHf298 values for 4+ and 7+ should be more reliable than the compiled values. For the five remaining cations with available experimental thermal data, the agreement between the experimental and G3 results ranges from fair to excellent.

Cation CH3CHSCH2+√ (10+) has the least energy among the eleven distonic radical cations identified. Their ΔHf298 values range from 918 to 1151 kJ mol−1. Nevertheless, only one of them, CH2=SCH2CH2+√ (12+), has been observed. Its G3 ΔHf298 value is 980.9 kJ mol−1, in fair agreement with the experimental result, 990 kJ mol−1.

A couple of reactions involving C3H6S+√ isomers CH2=SCH2CH2+√ (12+) and trimethylene sulfide radical cation (13+) have also been studied with the G3 method and the results are consistent with experimental findings.  相似文献   


7.
Rate coefficients for the reactions of cyclohexadienyl (c-C6H7) radicals with O2 and NO were measured at 296 ± 2 K. The c-C6H7 radicals were detected selectively by laser-induced fluorescence. The rate coefficient for the reaction of c-C6H7 with O2, (4.4 ± 0.5) × 10−14 cm3 molecule−1 s−1, was independent of the bath-gas (He) pressure (13–80 Torr). In the reaction of c-C6H7 with NO, thermal equilibrium among c-C6H7, NO, and C6H7NO was observed. The forward and reverse reactions were in the falloff region, and the equilibrium constant was (1.5 ± 0.6) × 10−15 cm3 molecule−1.  相似文献   

8.
Free radical reactions of dehydrozingerone (DZ), a methoxy phenol, were studied at dfferent pHs with a variety of oxidants using nanosecond pulse radiolysis technique. Hydroxyl radical (OH) reaction with the phenolic form at pH 6 led mainly to the formation of an OH-adduct absorbing at 460 nm in addition to a minor oxidation product. On the other hand, at pH 10 with the deprotonated phenoxide ion, the only reaction observable was oxidation generating a phenoxyl radical absorbing at 360 nm. HPLC analysis indicated formation of two different products at pH 6 from addition and oxidation reactions, whereas at pH 10, only the oxidation product was detectable. Reactions of more specific secondary oxidizing radicals, N3√, Br√, Br2√ and Tl(II) with DZ gave rise to the phenoxyl radical over the entire pH range. DZ in the phenoxide ion form reacted with nitrogen dioxide and trichloromethyl peroxyl radicals with rate constants 6×108 and 8.8×108 dm3 mol−1 s−1 respectively leading to the phenoxyl radicals. The DZ phenoxyl radical reacted with trolox C (an analogue of -tocopherol) with a rate constant of 8.3×107 dm3 mol−1 s−1. One electron reduction potential of the DZ phenoxyl radical at pH 6 was determined to be +1.1 V vs NHE using N3√/N3 as the standard couple.  相似文献   

9.
Two new procedures were employed for studying the reaction of hydrogen atoms with hydrogen peroxide. The absorption in the UV-range was observed either for an acidic aqueous solution containing only hydrogen peroxide or for a similar solution but also containing an aliphatic alcohol. From the increase in absorption of various alcohol radicals, a rate constant of 3.5×107 dm3 mol−1 s−1 was determined. In addition, the rate constant for the reaction of hydroxyl radicals with hydrogen peroxide was determined to be 3.0×107 dm3 mol−1 s−1.  相似文献   

10.
The one-electron oxidation of Mitomycin C (MMC) as well as the formation of the corresponding peroxyl radicals were investigated by both steady-state and pulse radiolysis. The steady-state MMC-radiolysis by OH-attack followed at both absorption bands showed different yields: at 218 nm Gi (-MMC) = 3.0 and at 364 nm Gi (-MMC) = 3.9, indicating the formation of various not yet identified products, among which ammonia was determined, G(NH3) = 0.81. By means of pulse radiolysis it was established a total κ (OH + MMC) = (5.8 ± 0.2) × 109 dm3 mol−1 s−1. The transient absorption spectrum from the one-electron oxidized MMC showed absorption maxima at 295 nm (ε = 9950 dm3 mol−1 cmt-1), 410 nm (ε = 1450 dm3 mol−1 cm−1) and 505 nm ( ε = 5420 dm3 mol−1 cm−1). At 280–320 and 505 nm and above they exhibit in the first 150 μs a first order decay, κ1 = (0.85 ± 0.1) × 103 s−1, and followed upto ms time range, by a second order decay, 2κ = (1.3 ± 0.3) × 108 dm3 mol-1 s−1. Around 410 nm the kinetics are rather mixed and could not be resolved.

The steady-state MMC-radiolysis in the presence of oxygen featured a proportionality towards the absorbed dose for both MMC-absorption bands, resulting in a Gi (-MMC) = 1.5. Among several products ammonia-yield was determined G(NH3) = 0.52. The formation of MMC-peroxyl radicals was studied by pulse radiolysis, likewise in neutral aqueous solution, but saturated with a gas mixture of 80% N2O and 20% O2. The maxima of the observed transient spectrum are slightly shifted compared to that of the one-electron oxidized MMC-species, namely: 290 nm (ε = 10100 dm3 mol−1 cm−1), 410 nm (ε = 2900 dm3 mol−1 cm−1) and 520 nm (ε = 5500 dm3 mol−1 cm−1). The O2-addition to the MMC-one-electron oxidized transients was found to be at 290 to 410 nm gk(MMC·OH + O2) = 5 × 107 dm3 mol−1 s−1, around 480 nm κ = 1.6 × 108 dm3 mol−1 s−1 and at 510 nm and above, κ = 3 × 108 dm3 mol−1 s−1. The decay kinetics of the MMC-peroxyl radicals were also found to be different at the various absorption bands, but predominantly of first order; at 290–420 nm κ1 = 1.5 × 103 s−1 and at 500 nm and above, κ = 7.0 × 103 s−1.

The presented results are of interest for the radiation behaviour of MMC as well as for its application as an antitumor drug in the combined radiation-chemotherapy of patients.  相似文献   


11.
The rate constant for the reaction between the sulphate radical (SO4√−) and the ruthenium (II) tris-bipyridyl dication (Ru(bipy)32+) is (3.3±0.2)×109 mol−1 dm3 s−1 in 1 mol dm−3 H2SO4 and (4.9±0.5)×109 mol−1 dm3 s−1 in 0.1 mol dm−3, pH 4.7 acetate buffer. The SO4√−radical produced by the electron transfer quenching of Ru(bipy)32+* by S2O82− reacts rapidly with both acetate buffer and chloride ions. These side reactions result in a reduction in the overall quantum yield of Ru(bipy)33+ production and reduced reaction selectivity when Ru(bipy)32+* is quenched by persulphate.  相似文献   

12.
The photoinduced electron transfer reactions of the triplet state of rose bengal (RB) and several electron donors were investigated by the complementary techniques of steady state and time-resolved electron paramagnetic resonance (EPR) and laser flash photolysis (LFP). The yield of radicals varied with the light fluence rate, RB concentration and, in particular, the electron donor used. Thus for L-dopa (dopa, dihydroxyphenylalanine) only 10% of RB anion radical (RB√−) was produced, with double the yield observed with NADH (NAD, nicotinamide adenine dinucleotide) as quencher and more than three times the yield observed with ascorbate as quencher. Quenching of the RB triplet was both reactive and physical with total quenching rate constants of 4 × 108 mol−1 dm3 s−1 and 8.5 × 108 mol−1 dm3 s−1 for ascorbate and NADH respectively. The rate constant for the photoinduced electron transfer from ascorbate to RB triplet was 1.4 × 108 mol−1 dm3 s−1 as determined by Fourier transform EPR (FT EPR). FT EPR spectra were spin polarized in emission at early times indicating a radical pair mechanism for the chemically induced dynamic electron polarization. Subsequent to the initial electron transfer production of radicals, a complex series of reactions was observed, which were dominated by processes such as recombination, disproportionation and secondary (bleaching) reactions.

It was observed that back electron transfer reactions could be prevented by mild oxidants such as ferric compounds and duroquinone, which were efficiently reduced by RB√−.  相似文献   


13.
The kinetics of the reactions of hydroxyl radicals with aliphatic alcohols in aqueous solution were studied using pulse radiolysis. Based on the optical absorption observed in the UV region, the rate constants for the reaction of hydroxyl radicals with methanol, ethanol, 2-propanol and tert-butyl alcohol were determined to be 9.0×108, 2.2×109, 2.0×109 and 6.2×108 dm3 mol−1 s−1, respectively. The values obtained here by direct observation of the alcohol radicals basically confirm the values which were earlier determined indirectly by competition.  相似文献   

14.
The rate constants at which oxidizing and reducing radicals react with the dinuclear iron(III) complex Fe2O(ttha)2− were measured in neutral aqueous solution. The rate constants for reduction of the complex by ·CO2.− CH3.CHOH and O2.− were found to be comparable with rate constants previously measured in mononuclear iron(III) polyaminocarboxylate systems. Fe2O(ttha)2− reacts slowly with O2.− (k8 = (1.2 ± 0.2) × 104 dm3 mol−1 s−1) and, hence, is a relatively poor catalyst for the dismutation of superoxide radical. The hydrated electron reduces the complex at a diffusion-controlled rate in a process which consumes one proton: eaq + Fe2O(ttha)2− → Fe2III,IIO(ttha)3− The reduction by carbon-centered radicals produces a (III,II) mixed-valence complex with an absorption spectrum different from that of the Fe2(II,III) species produced from reduction by the hydrated electron. The oxidizing radicals .OH and ·CO3 appear to act as reductants of the complex via ligand oxidation rather than by oxidation of the Fe2IIIO core to Fe2III,IVO. In the former case ligand attack appears to occur mainly at the methylene carbon of a glycinate group. The decarboxylation product, CO2, was detected by its aquation reaction in the presence of a pH sensitive dye, bromthymol blue.  相似文献   

15.
The reactions of two triphenyl methane (TPM) dyes—crystal violet (CV+) and malachite green (MG+)—with N3 and OH radicals were studied by pulse radiolytic kinetic spectrophotometry. The rate constants for the reaction of the cationic dyes (D+) with N3 are (9.0±0.6)×109 and (3.0±0.2)×109 dm3 mol−1 s−1 respectively and those for the reaction with OH are obtained as (8.0±0.6)×109 and (1.1±0.1)×109 dm3 mol−1 s−1 respectively. The transient spectra resulting from the oxidation of the dyes were characterized. The time-resolved spectra indicate that the reaction with OH radicals initially generates an adduct which subsequently dissociates to form the radical dication D•2+. The D•2+ species decay by further reaction with the parent dye.  相似文献   

16.
Excitation of solutions of Fe(bipy)2(CN)2 by a 266-nm laser pulse produces a hydrated electron and the oxidized complex, Fe(bipy)2 (CN)2+, in the primary photochemical step, in homogeneous aqueous solution as well as in aqueous solutions containing cetyltrimethylammonium bromide (CTAB) or sodium dodecyl sulfate (SDS) micelles. In all cases nascent hydrated electrons react with ground state Fe(bipy)2(CN)2 to form Fe(bipy)2(CN)2, and comparison of the decay constants in the three media (H2O: k = 2.8 × 1010 M−1 s−1; CTAB: k = 2.9 × 1010 M−1 s−1; SDS: k = 5.5 × 109 M−1 s−1), shows that the reaction is essentially unaffected by CTAB micelles but is much slower in SDS solution. Similar micellar effects were found for the back reaction between eaq and Fe(bpy)2(CN)2+. Rate constants for the scavenging of the photogenerated hydrated electrons by methyl viologen (MV2+) cations and NO3 anions were measured in the three systems, and the results indicate that for scavenging by MV2+ the rate constants are decreased in the micelle systems (k in H2O, 8.4 × 1010; CTAB, 3.5 × 1010 and SDS, 1.58 × 1010 M−1 s−1), whereas for NO3 the CTAB micelle decreases while the SDS micelle enhances the scavenging compared to water solution (k in H2O, 8.3 × 109; CTAB, 7 × 108; and SDS, 2.05 × 1010 M−1 s−1). For the comproportionation reaction between Fe(bipy)2(CN)2+ and Fe(bipy)2(CN)2 both micelles reduce the rate (k in H2O, 3.3 × 1010; CTAB, 2.3 × 1010; and SDS, 1.05 × 1010 M−1s−1), but while the reaction of Fe(bipy)2(CN)2+ with MV+ is increased in CTAB compared to water, it is slowed in SDS (k in H2O, 2.4 × 1010; CTAB, 8.9 × 1010; and SDS, 1.8 × 1010 M−1s−1). All effects observed in these microheterogeneous systems can be uniformly interpreted in terms of Coulombic interactions between the actual reactants and the charged surface of the micelles.  相似文献   

17.
The collisional quenching of electronically excited germanium atoms, Ge[4p2(1S0)], 2.029 eV above the 4p2(3P0) ground state, has been investigated by time-resolved atomic resonance absorption spectroscopy in the ultraviolet at λ = 274.04 nm [4d(1P10) ← 4p2(1S0)]. In contrast to previous investigations using the ‘single-shot mode’ at high energy, Ge(1S0) has been generated by the repetitive pulsed irradiation of Ge(CH3)4 in the presence of excess helium gas and added gases in a slow flow system, kinetically equivalent to a static system. This technique was originally developed for the study of Ge[4p2(1D2)] which had eluded direct quantitative kinetic study until recently. Absolute second-order rate constants obtained using signal averaging techniques from data capture of total digitised atomic decay profiles are reported for the removal of Ge(1S0) with the following gases (kR in cm3 molecule−1 s−1, 300 K): Xe, 7.1 ± 0.4 × 10−13; N2, 4.7 ± 0.6 × 10−12; O2, 3.6 ± 0.9 × 10−11; NO, 1.5 ± 0.3 × 10−11; CO, 3.4 ± 0.5 × 10−12; N2O, 4.5 ± 0.5 × 10−12; CO2, 1.1 ± 0.3 × 10−11; CH4, 1.7 ± 0.2 × 10−11; CF4, 4.8 ± 0.3 × 10−12; SF6, 9.5 ± 1.0 × 10−13; C2H4, 3.3 ± 0.1 × 10−10; C2H2, 2.9 ± 0.2 × 10−10; Ge(CH3)4, 5.4 ± 0.2 × 10−11. The results are compared with previous data for Ge(1S0) derived in the single-shot mode where there is general agreement though with some exceptions which are discussed. The present data are also compared with analogous quenching rate data for the collisional removal of the lower lying Ge[4p2(1D2)] state (0.883 eV), also characterized by signal averaging methods similar to that described here.  相似文献   

18.
Pentaerythrityl tetraethylenediamine (PETEDA) dendrimer was synthesized from pentaerythrityl tetrabromide and ethylenediamine. Its molecular structure was characterized by elemental analysis, Fourier transform infrared resonance (FT-IR) and hydrogen nuclear magnetic resonance (1H NMR) spectroscopy. The composite membranes for selectively permeating CO2 were prepared by using PETEDA-PVA blend polymer as the active layer and polyethersulfone (PES) ultrafiltration membrane as the support layer and their permselectivity was tested by pure CO2 and CH4 gases and the gas mixture containing 10 vol.% CO2 and 90 vol.% CH4, respectively. For pure gases, the membrane containing 78.6 wt% PETEDA and 21.4 wt% PVA in the blend has a CO2 permeance of 8.14 × 10−5 cm3 (STP) cm−2 s−1 cmHg−1 and CO2/CH4 selectivity of 52 at 143.5 cmHg feed gas pressure. While feed gas pressure is 991.2 cmHg, CO2 permeance reaches 3.56 × 10−5 cm3 (STP) cm−2 s−1 cmHg−1 and CO2/CH4 selectivity is 19. For the gas mixture, the membrane has a CO2 permeance of 6.94 × 10−5 cm3 (STP) cm−2 s−1 cmHg−1 with a CO2/CH4 selectivity of 33 at 188.5 cmHg feed gas pressure, and a CO2 permeance of 3.29 × 10−5 cm3 (STP) cm−2 s−1 cmHg−1 with a CO2/CH4 selectivity of 7.5 at a higher feed gas pressure of 1164 cmHg. A possible gas transport mechanism in the composite membranes is proposed by investigating the permeating behavior of pure gases and the gas mixture and analyzing possible reactions between CO2/CH4 gases and the PETEDA-PVA blend polymer. The effect of PETEDA content in the blend polymer on permselectivity of the composite membranes was investigated, presenting that CO2 permeance and CO2/CH4 selectivity increase and CH4 permeance decreases, respectively with PETEDA content. This is explained by that with increasing PETEDA content, the carrier content increases, and the crystallinity and free volume of the PETEDA-PVA blend decrease that were confirmed by the experimental results of X-ray diffraction spectra (XRD) and positron annihilation lifetime spectroscopy (PALS).  相似文献   

19.
The oxidation reaction of 2-aminophenol (OAP) to 2-aminophenoxazin-3-one (APX) initiated by 2,2,6,6-tetramethyl-1-piperidinyloxyl (TEMPO) has been investigated in methanol at ambient temperature. The oxidation of OAP was followed by electronic spectroscopy and the rate constants were determined according to the rate law −d[OAP]/dt=kobs[OAP][TEMPO]. The rate constant, activation enthalpy and entropy at 298 K are as follows: kobs (dm3 mol−1 s−1)=(1.49±0.02)×10−4, Ea=18±5 kJ mol−1, ΔH=15±4 kJ mol−1, ΔS=−82±17 J mol−1 K−1. The results of oxidation of OAP show that the formation of 2-aminophenoxyl radical is the key step in the activation process of the substrate.  相似文献   

20.
Excess molar enthalpies HE and excess molar volumes VE have been measured, as a function of mole fraction x1, at 298.15 K and atmospheric pressure for the five liquid mixtures (x11,4-C6H4F2 + x2n-ClH2l+2), l = 7, 8, 10, 12 and 16. In addition, HE and excess molar heat capacities CPE at constant pressure have been determined for the two liquid mixtures (x1C6F6 + x2n-ClH2l+2), l = 7 and 14, at the same temperature and pressure. The instruments used were flow microcalorimeters of the Picker design (the HE version was equipped with separators) and a vibrating-tube densimeter, respectively.

The excess enthalpies of the five difluorobenzene mixtures are all positive and quite large; they increase with increasing chain length l of the n-alkane from HE(x1 = 0.5)/(J mol−1) = 1050 for l = 7 to 1359 for l = 16. The corresponding excess volumes VE are all positive and also increase with increasing l: VE(x1 = 0.5)/(cm3 mol−1) = 0.650 for l = 7 and 1.080 for l = 16. Interestingly, the excess enthalphies of the corresponding mixtures with hexafluorobenzene are only about 5% larger, whereas the excess volumes of (x1C6F6 + x2n-ClH2l+2) are roughly twice as large as those of their counterparts in the series containing 1,4-C6H4F2. Specifically, at 298.15 K HE(x1 = 0.5)/(J mol−1) = 1119 for (x1C6F6 + x2n-C7H16) and 1324 for (x1C6F6 + x2n-C14H30), and for the same mixtures VE(x1 = 0.5)/(cm3 mol−1) = 1.882 and 2.093, respectively. The excess heat capacities for both systems are negative and of about the same magnitude as the excess heat capacities of mixtures of fluorobenzene with the same n-alkanes (Roux et al., 1984): CPE(x1 = 0.5)/(J K−1 mol−1) = −1.18 for (x1C6F6 + x2n-C7H16), and −2.25 for (x1C6F6 + x2n-C14H30). The curve CPE vs. (x1 for x1C6F6 + x2n-C14H30) shows a sort of “hump” for x1 0.5, which is presumed to indicate emerging W-shape composition dependence at lower temperatures.  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