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.
Herein, the reaction between CO2 and piperidine, as well as commercially available functionalised piperidine derivatives, for example, those with methyl‐, hydroxyl‐ and hydroxyalkyl substituents, has been investigated. The chemical reactions between CO2 and the functionalised piperidines were followed in situ by using attenuated total reflectance (ATR) FTIR spectroscopy. The effect of structural variations on CO2 absorption was assessed in relation to the ionic reaction products identifiable by IR spectroscopy, that is, carbamate versus bicarbonate absorbance, CO2 absorption capacity and the mass‐transfer coefficient at zero loading. On absorption of CO2, the formation of the carbamate derivatives of the 3‐ and 4‐hydroxyl‐, 3‐ and 4‐hydroxymethyl‐, and 4‐hydroxyethyl‐substituted piperidines were found to be kinetically less favourable than the carbamate derivatives of piperidine and the 3‐ and 4‐methyl‐substituted piperidines. As the CO2 loading of piperidine and the 3‐ and 4‐methyl‐ and hydroxyalkyl‐substituted piperidines exceeded 0.5 moles of CO2 per mole of amine, the hydrolysis of the carbamate derivative of these amines was observed in the IR spectra collected. From the subset of amines analysed, the 2‐alkyl‐ and 2‐hydroxyalkyl‐substituted piperidines were found to favour bicarbonate formation in the reaction with CO2. Based on IR spectral data, the ability of these amines to form the carbamate derivatives was also established. Computational calculations at the B3LYP/6‐31+G** and MP2/6‐31+G** levels of theory were also performed to investigate the electronic/steric effects of the substituents on the reactivity (CO2 capture performance) of different amines, as well as their carbamate structures. The theoretical results obtained for the 2‐alkyl‐ and 2‐hydroxyalkyl‐substituted piperidines suggest that a combination of both the electronic effect exerted by the substituent and a reduction in the exposed area of the nitrogen atom play a role in destabilising the carbamate derivative and increasing its susceptibility to hydrolysis. A theoretical investigation into the structure of the carbamate derivatives of these amines revealed shorter N? C bond lengths and a less‐delocalised electron distribution in the carboxylate moiety.  相似文献   

2.
Summary The conformational free energies (-Go) of a number of 4-substituted piperidines and piperidinium salts have been determined by the J-value method. For the 4-substituted piperidines (R=Me, Phenyl, CO2Et, Br, OH, F) the relative conformer energies are almost identical to those of the analogous cyclohexanes.The methyl and phenyl compounds showed no change in the couplings on protonation, implying no change in the conformer energies. In constrast, in the remaining compounds with polar 4-substituents an almost constant stabilisation of the axial conformer of ca. 0.7–0.8 kcal mol-1 was observed on protonation. In three cases (R=F, OH and Br) the conformational preferences is reversed on protonation and the axial form is favoured.The conformer energies of both the free bases and the piperidinium salts can be quantitatively predicted by molecular mechanics calculations using the COSMIC force-field, in which the electrostatic interactions are calculated by a simple Coulombic model with the partial atomic charges in the molecules given by the CHARGE2 routine, and an effective dielectric constant of five. The precise agreement obtained demonstrates conclusively that the electrostatic interactions between the substituents and the protonated nitrogen are the cause of the conformational changes on protonation, and that these can be modelled successfully using existing force-fields.For Part 15, see Ref. 1.  相似文献   

3.
Electrochemical behaviour of self-assembled monolayer of 2-dimethylaminoethanethiol (DMAET) on gold electrode, with and without integration with SWCNT-poly(m-aminobenzene sulphonic acid) (SWCNT-PABS) has been probed. It is proved that the so-called electric field induced protonation–deprotonation process, hitherto observed for the –COOH terminated SAMs, is also observed for the –N(H)+(CH3)2 terminated. The surface pKa of DMAET was estimated as 7.6, smaller than its solution pKa of 10.8. It is also shown that SWCNT-PABS is irreversibly attached to the DMAET SAM.  相似文献   

4.
Formation and protonation of aromatic anion radicals in 2-propanol were studied by kinetic spectrophotometric pulse radiolysis. All polycyclic hydrocarbons studied were found to react very rapidly with e?solv. Those with relatively high electron affinity were also reduced by (CH3)2CO?. The anion radicals formed undergo protonation by direct reaction with the alcohol molecule. The rate constants for this protonation vary from ≈ 6 × 105 s?1 for cis-stilbene and naphthalene down to 20 s?1 for perylene. The variations in rates are discussed in terms of changes in singlet energy separation (ΔES1 ← S0). The logarithm of the protonation rate constant for alternant hydrocarbons is linearly dependent on ΔES1 ← S0.  相似文献   

5.
This paper describes the effect of benzyl amine on the base-catalyzed transamination of α-keto esters. Among various benzyl amines examined, o-HOC6H4CH2NH2 was found to be highly effective for the reaction, affording a wide variety of α-amino esters in good yields. The o-OH group of the benzyl amine facilitates the transamination process likely via H-bond. Moderate enantiomeric excess was obtained for α-amino ester when a quinine derived catalyst was used.  相似文献   

6.
Thermogravimetry, differential thermal analysis, X-ray diffractometry and infrared spectroscopy showed that Ni(CH3COO)2·4H2O decomposes completely at 500°C, giving rise to a mixture of Nio and NiO. The results revealed that the compound undergoes dehydration at 160°C and melts at 310°C. The water thus released hydrolyses surface acetate groups, acetic acid being evolved into the gas phase. At 330°C, the anhydrous acetate is converted into NiCO3, releasing CH3COCH3 into the gas phase. The carbonate subsequently decomposes (at 365°C) to give NiO(s), CO2(g) and CO(g). On further heating up to 373°C, a mixture of Nio and NiO is formed. Other gas-phase products were detected at 400°C, viz. CH4 and (CH3)2CH=CH2, which were formed in surface reactions involving initial gas-phase products. Non-isothermal kinetic parameters (A and ΔE) were calculated on the basis of temperature shifts experienced in the various decomposition processes as a function of heating rate (2–20 deg·min?1).  相似文献   

7.
Abstract— –On in situ photolysis (Λ= 250–400 nm) of aqueous oxaloacetic acid solutions, between pH 5 and 10, the radicals -O2CCH2C(O->=C(O+)CH2CO2- and -O2CCH2C(OH)CO2 are identified. With acetone present, CH2CO2, CH3C(OH)CO-2, CH3C(O-)=C(O)CH2CO2 and -O2CCHCOCO2 are also observed. CO2 and CO are identified as reaction products. The experimental results are explained in terms of α-cleavage of the electronically excited keto-isomer dianion of oxaloacetic acid to yield O2CCH2CO and CO2-. -O2CCH2CO adds to the keto-isomer of oxaloacetic acid and to pyruvic acid, which is formed from oxaloacetic acid by thermal decarboxylation, to yield -O2CCH2C(O-)= C(O-)CH2C0- and CH3C(O-)=C(O)CH2CO-2, respectively, via a decarboxylase substitution reaction. CH2CO2 is derived from -O2CCH2CO by decarbonylation. CO2- is scavenged by oxaloacetic acid and pyruvic acid to yield O2CCH2C(OH)CO2 and CH3C(OH)CO2-, respectively.  相似文献   

8.
Decarbonation curve for the synthetic dolomite analogues; (Cd-dolomites) were determined for CdMg(CO3)2, CdMn(CO3)2 and CdZn(CO3)2 under CO2 pressure of up to 2.5 kbar. All the three double carbonates were completely disordered at the decomposition temperatures and hence the thermodynamic data (Standard enthalpy; ΔH f o, Standard free energy; ΔG f o) retrieved from the univariant decarbonation curve corresponds to the disordered phases. They are: The mixing enthalpies and free energies for the formation of the disordered 1∶1 solid solution phases are: The thermodynamic data (ΔH f o, ΔG f o and ΔH r o, ΔG r o) showed a positive correlation with the decomposition temperatures. The mixing energies of the disordered double carbonates also show a direct correlation with the cationic size differences.  相似文献   

9.
Based on the D6d-symmetrical C24H24 fullerene, its derivatives, which have been exohedrally functionalized by replacing one of its H atoms with –CH2OH, –CONH2, –COOH, and –COH, have been firstly calculated using the hybrid DFT-B3LYP functional in conjunction with 6-31G(d) basis sets. Present calculations show that the C24H23(COOH) is the most stable under the values of ΔH and De. The calculated HOMO–LUMO energy gaps of the derivatives with –CH2OH and –COOH groups are same as in the case of the not functionalized C24H24 cluster. In addition, the C24H23(CONH2) displays the largest dipole moment (3.32 D), and the vertical electron affinities of the functionalized derivatives are all very higher. These can promote new development in the fields of the bio-medical applications.  相似文献   

10.
The protonation of haloaromatics by [N2H]+ and [CO2H]+ has been studied by chemical ionization mass spectrometry. In general, the fragmentation reactions following protonation by [CO2H]+ are similar to those observed following protonation by [CH5]+, while the fragmentation reactions induced by protonation by [N2H]+ are intermediate between those observed on reaction with [CH5]+ and with [H3]+. These results are consistent with the conclusion that the fragmentation mode is determined by the protonation exothermicity since the proton affinity of CO2 is the same as that of CH4 while the proton affinity of N2 is intermediate between that of CH4 and H2.  相似文献   

11.
Standard thermodynamic values of proton ionisation of 3 substituted (Cl?, Br?, I?, C2H5? and CH2CN?) pyridine derivatives are determined at 25°C, in an aqueous medium of ionic strength 0.5 M KNO3.Free energies are deduced from equilibrium constants log K potentiometrically calculated. Enthalpies are obtained from calorimetric measurements.ΔGo, ΔHo and ΔSo values are discussed in the context of the results of our earlier studies on this subject.The Hammett plot corresponding to the thirteen systems examined gives the proton ionisation constant of a substituted pyridine from the equation log K = 5,48 – 5,94Σσ.Besides, a linear relationship is found between ΔGo and ΔHo, which confirms the observation made previously in other series of pyridine derivatives.  相似文献   

12.
The selective electrochemical oxidation of the phenol function in the case of hydroxymethyl phenol derivatives (o?, m?, p-hydroxybenzyl alcohol) leads to “reactive polymer” films of polyphenylene oxide substituted by CH2OH groups. The transformation of the hydroxyl function into an ester function by acetyl chloride indicates the reactivity of the CH2OH group. As for the family of carbonylated polyphenylene oxide films, reactivity is limited to the superficial layers of film. Average film thickness is between 50–100 nm; however with the ferrocene-ferricinium system acting as a redox catalyst, it can reach about 300 nm. This catalytic mechanism intervenes only when the oxidation potential of the ferrocene-ferricinium couple is very similar to that of the phenol derivative.  相似文献   

13.
Coordination and protolytic equilibria in the system vanadium(V)-malonic acid were studied calorimetrically at 298.15 K and I = 1.0(NaClO4). The heat effects of formation of the VO2CH2(COO) 2 ? complex and step protonation of the malonate ion were determined for the first time. The data obtained were used to calculate the thermodynamic characteristics (logK, ΔG, ΔH, and ΔS) of the equilibria studied.  相似文献   

14.
Reactions of tetra-n-butylammonium 2,4-dinitrophenyl hydrogen phosphate, (ArPH)?(R4N)+, in aprotic and protic solvents, in the absence and in the presence of alcohols or water, ROH, are compared with analogous reactions of the salt in the presence of hindered and unhindered amines, e.g. diisopropylethyl amine and quinuclidine. Similar studies are performed with the acid, ArPH2, in the presence of variable amounts of amines. The release of phenol and the fate of the phosphorus compounds are followed by 1H and 31P NMR spectrometry. In the absence of free unhindered amine, reactions of the monoanion are relatively slow, sensitive to steric hindrance in the alcohol, and incapable of producing t-butyl phosphate from t-butanol; reactions of the dianion are relatively fast, insensitive to steric hindrance in the alcohol, and produce t-butyl phosphate. In the presence of free unhindered amine, reactions of the monoanion are relatively fast but still sensitive to steric hindrance in the alcohol, and hence do not produce t-butyl phosphate. The intermediate CH(CH2CH2)3+NP(O)(OH)O? is detected in the presence of quinuclidine. Reactions of the dianion in the presence of unhindered amines are analogous to those observed in the presence of hindered amines. The uncatalyzed and the nucleophilic amine-catalyzed reactions of the monoanion are assumed to proceed via oxyphosphorane, P(5), intermediates. The dianion reactions, which are not susceptible to nucleophilic catalysis, are assumed to proceed via the monomeric metaphosphate ion intermediate, PO3?. Significant effects related to solvent properties are observed in these reactions.  相似文献   

15.
Equilibrium vapor pressures were determined at temperatures between 294 K and 353 K for the NiCl2-CH3CN system.Compositions studied ranged from molar ratios of CH3CN to NiCl2 of 0.27 to 1.90. Three stoichiometric compounds were identified: NiCl2(CH3CN)2, NiCl2(CH3CN), NiCl2(CH3CN)0.88. Per mole of gaseous CH3CN the values of ΔHo and ΔSo were calculated to be 52.0 ±0.4 kJ mol?1 and 149.0±1.3 J mol?1 K?1 for the decomposition of NiCl2(CH3CN)2, and 25.9 ±0.8 kJ mol?1 and 58.6± 2.1 J mol?1 K?1 for the decomposition of NiCl2(CH3CN). Below a composition of NiCl2(CH3CN)0.88 the phase diagram is complex and could not be interpreted in terms of specific stoichiometric compounds.  相似文献   

16.
The pressure dependence of the melting temperature of six aliphatic polyesters belonging to two different homologous series, poly(x-succinate) and poly(x-adipate) having even number of methylene groups (2,4,6) in the alkylene segment (x) was investigated by high pressure differential thermal analysis (HP-DTA) up to 500 MPa. The phase diagrams of these polyesters were newly determined except for poly(ethylene adipate). The dTm/dpo at atmospheric pressure was obtained from the quadratic equation and the trend of dTm/dpo with respect to the number of methylene groups in the monomer unit in each homologous series is discussed. Amorphous densities at 25 °C, expansion and compressibility coefficients in the melt of these polyesters are also reported. The entropy of fusion (ΔSm), enthalpy of fusion (ΔHm), volume change on melting (ΔVm), conformational entropy (ΔSor) and volume entropy (ΔSv) were correlated with respect to the number of methylene groups in the alkylene segment. ΔVm and ΔSv displayed a similar trend as that of dTm/dpo while ΔSm, ΔSor and ΔHm showed an increasing trend. The influence of these parameters on dTm/dpo is discussed in the context of the Clapeyron equation.  相似文献   

17.
Two fluorescent, water-soluble bis-naphthalenophane isomers with six carboxylate arms, abbreviated as (bis-dtpa14nap)H6 and (bis-dtpa15nap)H6, were synthesized, which consist of two 1,4- or 1,5-substituted naphthalene rings interlinked by two diethylenetriaminepentaacetic (DTPA) chains through amide-linkages. Both DTPA-based macrocycles exhibit intense excimer and monomer emission bands, which sensitively respond to pH in three protonation steps; more sensitive is the 1,4-naphthyl isomer. The full pH-emission profiles have confirmed that the mono-protonated anion (bis-dtpanap)H5 ? is the major protonation species at the physiological pH. Fluorometric titrations at pH 7.2 have proven that the 1,4-naphthalenophane anion forms 1:1-complexes with cationic phenethylamine (formation constant, 5700 M?1) and histamine (3000 M?1), excluding tryptamine cation, whereas the 1,5-isomer does not react with any of the three amines. The primary binding forces are electrostatic interactions between the CH2CO2 ? arms of 1,4-naphthalenophane and the CH2CH2NH3 + chain of an aromatic amine. The resulting ion-pair is stabilized by encapsulation of the guest molecule in 1,4-napthalenophane cavity, while the 1,5-isomer cannot encapsulate. NMR studies have demonstrated that 1,4-napthalenophane has a higher freedom in reorientation of naphthalene rings. Such geometrical properties controlled by selection of naphthalene units are the feature of the new naphthalenophanes, and are responsible for the pH?emission profiles and the complexation.  相似文献   

18.
The chemisorption of CO2 by aqueous-hindered amines has been investigated experimentally and theoretically. Negative-ion ESI–MS analysis of solutions containing a sterically hindered amine and a source of 13CO2 reveals peaks corresponding to [M–H + 45]?. These ions readily lose 45 Da when subjected to collisional activation, and together with other key fragments confirms the generation of the 13C-labelled carbamate derivatives. The thermochemistry of the two key capture reactions: $$2.{\text{amine }} + {\text{ CO}}_{ 2} { \leftrightarrows }{\text{amine}} - {\text{CO}}_{ 2}^{ - } + {\text{ amine}} - {\text{H}}^{ + } {\kern 1pt} \quad 1:{\text{carbam}}$$ $${\text{amine }} + {\text{ CO}}_{ 2} + {\text{ H}}_{ 2} {\text{O}}{ \leftrightarrows }{\text{HCO}}_{ 3}^{ - } + {\text{ amine}} - {\text{H}}^{ + } \quad 2:{\text{ bicarb}}$$ at 298 K was modelled using composite chemistry methods, CCSD(T), DFT, and SM8 free energies of solvation. The aqueous reaction free energies (ΔG 298) for reaction 1 are predicted to be more negative than ΔG 298 for reaction 2 when amine = ammonia, 2-aminoethanol (MEA), 2-amino-2-methyl-1-propanol (AMP), 2-amino-2-hydroxymethyl-propane-1,3-diol (tris), and 2-piperidinemethanol (2-PM). For AMP, tris, and 2-PM, activation free energies ΔG 298 ? for reaction 1 (SM8 + CCSD(T)/6-311 ++G(d,p)//M08-HX/MG3S: 38–67 kJ mol?1) are smaller than the corresponding values for 2 (109–113 kJ mol?1). For 2-PM, the computed carbamate ΔG 298 ? (38 kJ mol?1) is comparable to the MEA value (45 kJ mol?1), whereas the primary amines with tertiary alpha carbons have slightly larger values (60–70 kJ mol?1). The organic amine values are much lower than the value for ammonia (93 kJ mol?1). The results indicate CO2 chemisorption proceeds via a carbamate intermediate for all aqueous primary and secondary amines. Hindered carbamates are susceptible to further chemical transformations following their formation.  相似文献   

19.
Eight diorganotin esters of Schiff base ligands formulated as [R2SnLY]2, where L1 is 4-NC5H4CON2C(CH3) CO2 with Y = H2O, R = Ph (1), PhCH2 (2), m-ClC6H4CH2 (3), and L2 is 2-HOC6H4CON2C(CH3) CO2 with Y = CH3OH, R = PhCH2 (4), o-ClC6H4CH2 (5), m-ClC6H4CH2 (6), o-FC6H4CH2 (7), p-FC6H4CH2 (8) have been prepared and characterized by elemental analysis, IR, 1H and 119Sn NMR spectra. The crystal structures of compounds 1, 2 and 4 have been determined by X-ray single crystal diffraction. Their structures show that the tin atoms of three compounds are all rendered seven-coordinated in distorted pentagonal bipyramid geometries with a planar SnO4N unit and two apical aryl carbon atoms. A comparison of the IR spectra of the ligands with those of the corresponding compounds, reveals that the disappearance of the bands assigned to carbonyl unambiguously conforms that the ligands coordinate with tin in the enol form.  相似文献   

20.
Olefinic amines of the ytpe CH2CH(CH2)nNHR undergo cyclization in acidic aqueous solution at 60°C in the presence of PtCl42?. The PtCl42? is regenerated at the end of the cyclization so that the reaction may be considered as catalytic. Both pyrrolidines and piperidines may be formed although the former are favored.  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