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 890 毫秒
1.
Solvent transports across the perfluorosulfonic acid-type membrane Flemion S were measured for aqueous electrolyte solutions under a temperature difference and under an osmotic pressure difference. H+, Li+, Na+, K+, NH 4 + , CH3NH 3 + , (CH3)2NH 2 + , (CH3)3NH+, (CH3)4N+, (C2H5)4N+, (n-C3H7)4N+ and (n-C4H9)4N+ were used as counterions. Water flux across the membrane in HCl solution is higher than that in the other electrolyte solutions because hydrogen ions can exchange with the hydrogen of the neighbor water molecules and contribute to the water transport across the membrane as a proton jump in conductivity. The direction of thermoosmosis across the membrane in HCl, NaCl, (CH3)4NCl and (C2H5)4NCl solutions was from the cold side to the hot side and that in LiCl, KCl, NH4Cl, CH3NH3Cl, (CH3)2NH2Cl and (n-C4H9)4NBr solutions was from the hot side to the cold side, although thermoosmosis across anion-exchange membranes always occurs toward the hot side.  相似文献   

2.
Reactions that proceed within mixed ethylene–methanol cluster ions were studied using an electron impact time-of-flight mass spectrometer. The ion abundance ratio, [(C2H4)n(CH3OH)mH+]/[(C2H4)n(CH3OH)m+], shows a propensity to increase as the ethylene/methanol mixing ratio increases, indicating that the proton is preferentially bound to a methanol molecule in the heterocluster ions. The results from isotope-labelling experiments indicate that the effective formation of a protonated heterocluster is responsible for ethylene molecules in the clusters. The observed (C2H4)n(CH3OH)m+ and (C2H4)n(CH3OH)m–1CH3O+ ions are interpreted as a consequence of the ion–neutral complex and intracluster ion–molecule reaction, respectively. Experimental evidence for the stable configurations of heterocluster species is found from the distinct abundance distributions of these ions and also from the observation of fragment peaks in the mass spectra. Investigations on the relative cluster ion distribution under various conditions suggest that (C2H4)n(CH3OH)mH+ ions with n + m ≤ 3 have particularly stable structures. The result is understood on the basis of ion–molecule condensation reactions, leading to the formation of fragment ions, $ {\rm CH}_2=\!=\mathop {\rm O}\limits^ + {\rm CH}_3 $ and (CH3OH)H3O+, and the effective stabilization by a polar molecule. The reaction energies of proposed mechanisms are presented for (C2H4)n(CH3OH)mH+(n + m ≤ 3) using semi-empirical molecular orbital calculations.  相似文献   

3.
IR photodissociation spectra of mass‐selected clusters composed of protonated benzene (C6H7+) and several ligands L are analyzed in the range of the C? H stretch fundamentals. The investigated systems include C6H7+? Ar, C6H7+? (N2)n (n=1–4), C6H7+? (CH4)n (n=1–4), and C6H7+? H2O. The complexes are produced in a supersonic plasma expansion using chemical ionization. The IR spectra display absorptions near 2800 and 3100 cm?1, which are attributed to the aliphatic and aromatic C? H stretch vibrations, respectively, of the benzenium ion, that is, the σ complex of C6H7+. The C6H7+? (CH4)n clusters show additional C? H stretch bands of the CH4 ligands. Both the frequencies and the relative intensities of the C6H7+ absorptions are nearly independent of the choice and number of ligands, suggesting that the benzenium ion in the detected C6H7+? Ln clusters is only weakly perturbed by the microsolvation process. Analysis of photofragmentation branching ratios yield estimated ligand binding energies of the order of 800 and 950 cm?1 (≈9.5 and 11.5 kJ mol?1) for N2 and CH4, respectively. The interpretation of the experimental data is supported by ab initio calculations for C6H7+? Ar and C6H7+? N2 at the MP 2/6‐311 G(2df,2pd) level. Both the calculations and the spectra are consistent with weak intermolecular π bonds of Ar and N2 to the C6H7+ ring. The astrophysical implications of the deduced IR spectrum of C6H7+ are briefly discussed.  相似文献   

4.
Formation constants have been measured by a solvent distribution method for the ion pairing of an arene sulfonate, methyl orange dye, with two series of quaternary ammonium ions: R4N+(R=Et,n-Pr,n-Bu, andn-Pent) and C6H5CH2R3N+ (R=Me, Et,n-Pr,n-Bu,n-Pent, andn-Hex). Ion pairing increases dramatically as the length of the R group increases beyond butyl. Using a hard-sphere model for contact ion pairs, it is estimated that coulombic attraction contributes about –kT to the binding free energy and decreases slightly with increasing size of R4N+. Other factors related to solvation effects, of which cosphere overlap predominates, contribute from –2kT to –7kT of binding energy. Plots of logK for association as a function of cation size show an inflection with decreasing slope between R=propyl and R=butyl. Possible causes for the inflection are considered.  相似文献   

5.
MINDO/3 calculations for singlet and triplet doubly charged benzene [C6H6]2+ are in satisfactory agreement with the experimentally determined values of the vertical double ionization energy of benzene; calculations for straight chain isomeric structures are consistent with the observed kinetic energy release on fragmentation to [C5H3]+ and [CH3]+. Symmetrical doubly charged benzene ions relax to a less symmetrical cyclic structure having sufficient internal energy to fragment by ring opening and hydrogen transfer towards the ends of the carbon chain. Fragmentation of [CH3C4CH3]2+ to [CH3C4]+ and [CH3]+ is a relatively high energy process (A), whereas both (B): [CH3CHC3CH2]2+ to [CHC3CH2]+ and [CH3]+ and (C): [CH3CHCCHCCH]2+ to [CHCCHCCH]+ and [CH3]+ may be exothermic processes from doubly charged benzene. Furthermore, the calculated energy for the reverse of process (A) is less than the experimentally observed kinetic energy released, whereas larger energies for the reverse of processes B and C are predicted. Heats of formation of homologous series [HCn]+, [CH3Cn]+, [CH2Cn?2CH]+, [CH3Cn?2CH2]+ and [CH2?CHCn?3CH2]+ with 1 < n < 6 are calculated to aid prediction of the most stable products of fragmentation of doubly charged cations. The homologous series [CH2Cn?2CH]+ is relatively stable and may account for ready fragmentation of doubly charged ions to [CnH3]+; alternatively the symmetrical [C5H3]+ ion [CHCCHCCH]+ may be formed. Dicoordinate carbon chains appear to be important stabilizing features for both cations and dications.  相似文献   

6.
An improved, high‐yield, one‐pot synthetic procedure for water‐soluble ligands functionalized with trialkyl ammonium side groups H2N(CH2)2NHSO2p‐C6H4CH2[NMe2(CnH2n+1)]+ ( [HL n ]+ ; n=8, 16) was developed. The corresponding new surface‐active complexes [(p‐cymene)RuCl( L n )] and [Cp*RhCl( L n )] (Cp*=η5‐C5Me5) were prepared and characterized. For n=16 micelles are formed in water at concentrations as low as 0.6 mM , as demonstrated by surface‐tension measurements. The complexes were used for catalytic transfer hydrogenation of ketones with formate in water. Highly active catalyst systems were obtained in the case of complexes bearing C16 tails due to their ability to be adsorbed at the water/substrate interface. The scope of these catalyst systems in aqueous solutions was extended from partially water soluble aryl alkyl ketones (acetophenone, butyrophenone) to hydrophobic dialkyl ketones (2‐dodecanone).  相似文献   

7.
A series of novel cationic gemini surfactants, p-[C n H2n+1N+(CH3)2CH2CH(OH)CH2O]2C6H4·2Cl? [A(n = 12), B(n = 14) and C(n = 16)], containing a spacer group with two flexible and hydrophilic groups (2-hydroxy-1,3-propylene) on both sides of a rigid and hydrophobic group (1,4-dioxyphenylene) has been synthesized by the reaction of hydroquinone diglycidyl ether with N,N-dimethylalkylamine and N,N-dimethylalkylamine hydrochloride. Their surface-active properties have been investigated by surface tension measurement. The critical micelle concentration (cmc) values of the synthesized cationic gemini surfactants are one order of magnitude lower than those of their corresponding monomeric surfactants (C n H2n + 1N+(CH3)3·Cl?). Both the cmc and surface tension at the cmc (γcmc) of A are lower than those of p-[C12H25N+(CH3)2CH2]2C6H4·2Cl? (D). The novel cationic gemini surfactants A and B also show good foaming properties.  相似文献   

8.
η6-o-Chlorotoluene-η5-cyclopentadienyliron hexafluorophosphate undergoes nucleophilic substitution of the chlorine atom with anions generated (K2CO3/DMF) from methyl thioglycolate, diethyl malonate, dimethyl malonate, methyl acetoacetate and 2,4-pentanedione. The compounds prepared were o-CH3C6H4SCH2CO2CH3FeCp+PF6, o-CH3C6H4CH(CO2C2H5)2FeCp+PF6, o-CH3C6H4CH(CO2CH3)2FeCp+PF6, o-CH3C6H4CH(COCH3)CO2CH3FeCp+PF6 and o-CH3C6H4CH2COCH3FeCp+PF6 . Similarly, the reaction of diethyl malonate, dimethyl malonate, methyl acetoacetate anions and methylamine with η6-2,6-dichlorotoluene-η5-cyclopentadienyliron hexafluorophosphate yielded monosubstitution of one of the chloro groups. The complexes prepared in this study were η6-diethyl(3-chloro-2-methyl) phenylmalonate- η5-cyclopentadienyliron hexafluorophosphate, η6-dimethyl(3-chloro-2-methyl)phenylmalonate-η5-cyclopentadienyliron hexafluorophosphate, η6-methyl(3-chloro-2-methyl)phenylacetoacetate-η5-cyclopentadienyliron hexafluorophosphate and η6-3-chloro(2-methyl-N-methyl)aniline-η5-cyclopentadienyliron hexafluorophosphate. Reaction of η6-2,6-dichlorotoluene-η5-cyclopentadienyliron hexafluorophosphate with excess methanol as well as methyl thioglycolate in the presence of K2CO3 resulted in disubstitution of both chloro groups to yield new complexes, η6-2,6-dimethoxytoluene-η5-cyclopentadienyliron hexafluorophosphate and η6-methyl[(2-methylphenyl)1,3-dithio] diacetate-η55-cyclopentadienyliron hexafluorophosphate, respectively. Complexes o-CH3C6H4CH(CO2C2H5)2FeCp+PF6, o-CH3C6H4CH(CO2CH3)2FeCp+PF6 and o-CH3C6H4CH2 COCH3FeCp+ PF6 react with excess K2CO3 and benzyl bromide in refluxing methylene chloride to give 80–90% yields of complexes o-CH3C6H4C(CH2C6H5)(CO2C2H5)2FeCp+PF6, o-CH3C6H4C(CH2C6H5)(CO2CH3)2FeCp+PF6 and o-CH3C6H4CH(CH2C6H5)COCH3FeCp+PF6, respectively. Reaction of complex, o-CH3C6H4C(CH2C6H5)(CO2C2H5)2FeCp+PF6 with one molar equivalent of t-BuOK followed by acidic work-up gives o-(C2H5CO2CH2)C6H4CH(CO2C2H5)CH2C6H5FeCp+PF6. Similarly, reactions of complexes o-CH3C6H4C(CH2C6H5)(CO2C2H5)2FeCp+PF6 and o-CH3C6H4C(CH2C6H5)(CO2CH3)2FeCp+PF6 with t-BuOK in THF followed by alkylation with methyl iodide gave the new complexes, o-(C2H5O2C(CH3)CH)C6H4CH(CH2C6H5)CO2C2H5FeCp+PF6 and o-(CH3O2C(CH3)CH)C6H4CH(CH2C6H5)CO2CH3FeCp+PF6, respectively. Vacuum sublimation of the new complexes, o-CH3C6H4C(CH2C6H5)(CO2C2H5)2FeCp+PF6 and o-(C2H5O2CCH2)C6H4CH(CH2C6H5)CO2C2H5FeCp+PF6 gives o-CH3C6H4C(CH2C6H5)(CO2C2H5)2 and O-(C2H5O2CCH2)C6H4CH(CH2C6H5)CO2C2H5, respectively.  相似文献   

9.
The [C4H6O] ion of structure [CH2?CHCH?CHOH] (a) is generated by loss of C4H8 from ionized 6,6-dimethyl-2-cyclohexen-1-ol. The heat of formation ΔHf of [CH2?CHCH?CHOH] was estimated to be 736 kJ mol?1. The isomeric ion [CH2?C(OH)CH?CH2] (b) was shown to have ΔHf, ? 761 kJ mol?1, 54 kJ mol?1 less than that of its keto analogue [CH3COCH?CH2]. Ion [CH2?C(OH)CH?CH2] may be generated by loss of C2H4 from ionized hex-1-en-3-one or by loss of C4H8 from ionized 4,4-dimethyl-2-cyclohexen-1-ol. The [C4H6O] ion generated by loss of C2H4 from ionized 2-cyclohexen-1-ol was shown to consist of a mixture of the above enol ions by comparing the metastable ion and collisional activation mass spectra of [CH2?CHCH?CHOH] and [CH2?C(OH)CH?CH2] ions with that of the above daughter ion. It is further concluded that prior to their major fragmentations by loss of CH3˙ and CO, [CH2?CHCH?CHOH]+˙ and [CH2?C(OH)CH?CH2] do not rearrange to their keto counterparts. The metastable ion and collisional activation characteristics of the isomeric allenic [C4H6O] ion [CH2?C?CHCH2OH] are also reported.  相似文献   

10.
Crystals of brucinium 3,5‐dinitro­benzoate methanol solvate, C23H27N2O4+·C7H3N2O6·CH3OH, (I), brucinium 3,5‐dinitro­benzoate methanol disolvate, C23H27N2O4+·C7H3N2O6·2CH3OH, (II), and brucinium 3,5‐dinitro­benzoate trihydrate, C23H27N2O4+·C7H3N2O6·3H2O, (III), were obtained from methanol [for (I) and (II)] or ethanol solutions [for (III)]. The brucinium cations and 3,5‐dinitro­benzoate anions are linked by ionic N—H+⋯O hydrogen bonds. In the crystals of (I), (II) and (III), the brucinium cations exhibit different modes of packing, viz. corrugated ribbons, pillars and corrugated monolayer sheets, respectively. While in (III), the amide O atom of the brucinium cation participates in O—H⋯O hydrogen bonds, in which water mol­ecules are the donors, in (I) and (II), the amide O atom of the brucinium cation is involved in weak C—H⋯O hydrogen bonds and other brucinium cations are the donors.  相似文献   

11.
Study of carbonylation of benzyl bromide in a biphasic liquid-liquid system shows that the catalytic ion pair Bu4N+Co(CO)4? stays in the organic layer and that the carbonylated product C6H5CH2CO2? stays in the aqueous one.  相似文献   

12.
IntroductionInrecentyears ,bis(quaternaryammonium)surfac tantsorgeminisurfactants ,inwhichtwocationicsurfac tantmoietiesareconnectedwiththeammoniumheadgroupbyaploymethylenechain ,namely ,aspacerhavebecomeofinterestduetotheirexceptionalsurfaceactivityandrem…  相似文献   

13.
The mass spectra of several alkyl phenyl tellurides, C6H5TeR (R = CH3, CD3, C2H5, n-C3H7, i-C3H7 and n-C4H9) have been studied with special emphasis on the fragmentation patterns involving cleavage of the alkyl and aryl tellurium–carbon bonds. Each compound exhibited intense parent ions. The rearrangement ions [C6H6Te]+? and [C6H6]+? were found in the spectra of phenyl ethyl and higher tellurides. Two other rearrangement ions [HTe]+ and [C7H7]+ were observed in the spectrum of each compound. Examination of the mass spectrum of phenyl methyl-d3 telluride demonstrated that the [HTe]+ ions derive hydrogen from the phenyl group.  相似文献   

14.
In framework molecular cations and radical cations of adamantane C10H m q+ and also in polyhedral molecules and molecular ions C5H5 +, C6H6 2 +, B5H9, and B10H10 2 -, the charge density of valence electrons in the central areas of C n and B n cavities and faces is significant. In the molecule of adamantane C10H16, the valence electron density in central areas of the cavity and faces of the C10 framework is small as compared to the electron density along its edges C-C. These distinctions are due to the fact that, in the electronic structure of C n H q m cations and radical cations and also of B n H m molecules and molecular ions, there is an additional orbital interaction involving vacant valence orbitals of C+ or B (orbital-reduntant bonds); the absence of vacant valence orbitals of C atoms in neutral adamantane molecule excludes additional orbital interactions in excess of C-H and C-C.  相似文献   

15.
Solubility in ternary aqueous stratifying systems containing Catamine AB (alkylbenzyldimethylammonium chloride [C n H2n + 1N+(CH3)2CH2C6H5] · Cl, a cationic surfactant, where n = 10–18) and LiCl, NaCl, KCl, and NH4Cl inorganic salts was studied for the first time at 25°C. The boundaries of two-phase liquid equilibrium fields were determined. The studied stratifying systems were proposed for use in the liquid extraction of metal ions.  相似文献   

16.
The behaviour under electron impact (70 eV) which includes some rearrangement processes of some tetraorganodiphosphanedisulfides R2P(S)-P(S)R2 (R ? CH3, C2H5, n-C3H7, n-C4H9, C3H5, C6H5) and CH3RP(S)–P(S)CH3R (R ? C2H5, n-C3H7, n-C4H9, C6H5, C6H5, C6H5,CH2) is reported and discussed. Fragmentation patterns which are consistent with direct analysis of daughter ions and defocusing metastable spectra are given. The atomic composition of many of the fragment ions was determined by precise mass measurements. In contrast to compounds R3P(S) loss of sulphur is not a common process here. The first step in the fragmentation of these compounds is cleavage of one P–C bond and loss of a substituent R?. The second step is elimination of RPS leading to [R2PS]+ from which the base peaks in nearly all the spectra arise. The phenyl substituted compounds give spectra with very abundant [(C6H5)3P]+. and [(C6H5)2CH3P]+. ions respectively, resulting from [M]+. by migration of C6H5. Rearrangement of [M]+. to a 4-membered P-S ring system prior to fragmentation is suggested.  相似文献   

17.
The complexes [Pt(tBu3tpy){C?C(C6H4C?C)n?1R}]+ (n=1: R=alkyl and aryl (Ar); n=1–3: R=phenyl (Ph) or Ph‐N(CH3)2‐4; n=1 and 2, R=Ph‐NH2‐4; tBu3tpy=4,4’,4’’‐tri‐tert‐butyl‐2,2’:6’,2’’‐terpyridine) and [Pt(Cl3tpy)(C?CR)]+ (R=tert‐butyl (tBu), Ph, 9,9’‐dibutylfluorene, 9,9’‐dibutyl‐7‐dimethyl‐amine‐fluorene; Cl3tpy=4,4’,4’’‐trichloro‐2,2’:6’,2’’‐terpyridine) were prepared. The effects of substituent(s) on the terpyridine (tpy) and acetylide ligands and chain length of arylacetylide ligands on the absorption and emission spectra were examined. Resonance Raman (RR) spectra of [Pt(tBu3tpy)(C?CR)]+ (R=n‐butyl, Ph, and C6H4‐OCH3‐4) obtained in acetonitrile at 298 K reveal that the structural distortion of the C?C bond in the electronic excited state obtained by 502.9 nm excitation is substantially larger than that obtained by 416 nm excitation. Density functional theory (DFT) and time‐dependent DFT (TDDFT) calculations on [Pt(H3tpy)(C?CR)]+ (R= n‐propyl (nPr), 2‐pyridyl (Py)), [Pt(H3tpy){C?C(C6H4C?C)n?1Ph}]+ (n=1–3), and [Pt(H3tpy){C?C(C6H4C?C)n?1C6H4‐N(CH3)2‐4}]+/+H+ (n=1–3; H3tpy=nonsubstituted terpyridine) at two different conformations were performed, namely, with the phenyl rings of the arylacetylide ligands coplanar (“cop”) with and perpendicular (“per”) to the H3tpy ligand. Combining the experimental data and calculated results, the two lowest energy absorption peak maxima, λ1 and λ2, of [Pt(Y3tpy)(C?CR)]+ (Y=tBu or Cl, R=aryl) are attributed to 1[π(C?CR)→π*(Y3tpy)] in the “cop” conformation and mixed 1[dπ(Pt)→π*(Y3tpy)]/1[π(C?CR)→π*(Y3tpy)] transitions in the “per” conformation. The lowest energy absorption peak λ1 for [Pt(tBu3tpy){C?C(C6H4C?C)n?1C6H4‐H‐4}]+ (n=1–3) shows a redshift with increasing chain length. However, for [Pt(tBu3tpy){C?C(C6H4C?C)n?1C6H4‐N(CH3)2‐4}]+ (n=1–3), λ1 shows a blueshift with increasing chain length n, but shows a redshift after the addition of acid. The emissions of [Pt(Y3tpy)(C?CR)]+ (Y=tBu or Cl) at 524–642 nm measured in dichloromethane at 298 K are assigned to the 3[π(C?CAr)→π*(Y3tpy)] excited states and mixed 3[dπ(Pt)→π*(Y3tpy)]/3[π(C?C)→π*(Y3tpy)] excited states for R=aryl and alkyl groups, respectively. [Pt(tBu3tpy){C?C(C6H4C?C)n?1C6H4‐N(CH3)2‐4}]+ (n=1 and 2) are nonemissive, and this is attributed to the small energy gap between the singlet ground state (S0) and the lowest triplet excited state (T1).  相似文献   

18.
The kinetics of C6H5 reactions with n‐CnH2n+2 (n = 3, 4, 6, 8) have been studied by the pulsed laser photolysis/mass spectrometric method using C6H5COCH3 as the phenyl precursor at temperatures between 494 and 1051 K. The rate constants were determined by kinetic modeling of the absolute yields of C6H6 at each temperature. Another major product C6H5CH3 formed by the recombination of C6H5 and CH3 could also be quantitatively modeled using the known rate constant for the reaction. A weighted least‐squares analysis of the four sets of data gave k (C3H8) = (1.96 ± 0.15) × 1011 exp[?(1938 ± 56)/T], and k (n‐C4H10) = (2.65 ± 0.23) × 1011 exp[?(1950 ± 55)/T] k (n‐C6H14) = (4.56 ± 0.21) × 1011 exp[?(1735 ± 55)/T], and k (n?C8H18) = (4.31 ± 0.39) × 1011 exp[?(1415 ± 65)T] cm3 mol?1 s?1 for the temperature range studied. For the butane and hexane reactions, we have also applied the CRDS technique to extend our temperature range down to 297 K; the results obtained by the decay of C6H5 with CRDS agree fully with those determined by absolute product yield measurements with PLP/MS. Weighted least‐squares analyses of these two sets of data gave rise to k (n?C4H10) = (2.70 ± 0.15) × 1011 exp[?(1880 ± 127)/T] and k (n?C6H14) = (4.81 ± 0.30) × 1011 exp[?(1780 ± 133)/T] cm3 mol?1 s?1 for the temperature range 297‐‐1046 K. From the absolute rate constants for the two larger molecular reactions (C6H5 + n‐C6H14 and n‐C8H18), we derived the rate constant for H‐abstraction from a secondary C? H bond, ks?CH = (4.19 ± 0.24) × 1010 exp[?(1770 ± 48)/T] cm3 mol?1 s?1. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 36: 49–56, 2004  相似文献   

19.
From the reaction of 1‐HOCPh2‐2‐NMe2C6H4 ( 1 ), 1‐HOC(C6H11)2‐2‐NMe2C6H4 ( 2 ) and 1‐HOCPh2CH2‐2‐NMe2C6H4 ( 3 ) with n‐BuLi in diethyl ether, the solvent‐free chelated dimethylamino lithium alkoxides [1‐LiOCPh2‐2‐NMe2C6H4]2 ( 4 ), [1‐LiOC(C6H11)2‐2‐NMe2C6H4]2 ( 5 ) and [1‐LiOCPh2CH2‐2‐NMe2C6H4]2 ( 6 ) were obtained. The lithium alkoxides 4 – 6 were characterized by 1H, 7Li, and 13C NMR spectroscopy. Crystal structure determinations of 5 and 6 were carried out. Compounds 5 and 6 are examples of structurally characterized solvent‐free chelated dimethylamino lithium alkoxides and 6 is a rare example of this type containing a seven‐membered ring. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

20.
The decomposition of the [C6H5CO]+ ions produced from eight alkyl benzoates by electron impact has been studied. By calculating the heat of formation of [C6H5CO]+ ions from the appearance potential value, it is shown that the ions from C6H5COOR when R?H, CH3, C2H5 have some excess energy, and those where R = n-C3H7, iso-C3H7, n-C4H9, iso-C4H9, iso-C5H11 are produced with more excess energy. It is also shown that by taking this excess energy into account, there is a linear relationship between the heat of formation of the activated complex produced in the reaction [C6H5CO]+→[C6H5]+ + CO and the vibrational degree of freedom of the neutral fragment ? OR.  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