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 390 毫秒
1.
The chromaticity of poly(o-anisidine) (POAN) doped with different acids (HA), HA-doped POAN, has been studied by the spectrophotometric technique and the results were substantiated by molecular mechanics (MM+) calculations. The observed absorbance decrease (λ around 720 nm, dark green coloration) with increasing concentration of the inorganic oxidizing agent (KMnO4) can be attributed to the oxidative hydrolysis mechanism. The oxidative hydrolysis constant (K h) is highly dependent on the strength of the acid used. The HClO4-doped POAN matrix has the ability to store about 128.878 kJ⋅g−1 chromogenic energy (CE) at the wavelength 720 nm in a condensed lightweight form. MM+ calculations suggest that the potential energy (PE) in kJ⋅mol−1 of the optimum molecular geometric (OMG) structure of the HClO4-doped POAN matrix is at least two (2.052) times more stable than the OMG of the base form (POAN-EB) of the POAN matrix. Kinetic parameters of the oxidative hydrolysis reaction of the HA-doped POAN matrix were deduced from absorbance variations with time. The results of computer-oriented kinetic analysis indicate that the rate-controlling step for HA-doped POAN oxidative hydrolysis is governed by the Ginstling-Bronshein equation that represents three-dimensional diffusion (D4). Activation parameters for the oxidative hydrolysis of the HClO4-doped POAN matrix were computed and discussed.  相似文献   

2.
The incorporation of 9,10-anthraquinone-1,5-disulfonate (AQS2) into the protonated form of poly(o-toluidine) (POT), produced by oxidative polymerization of the cationic form of the monomer or by doping the basic form (POT-EB) by anion-exchange has been studied by FTIR and UV-VIS spectroscopy and mass spectrometry. The presence of sulfur and the absence of chlorine proven by elemental analysis of the polymer product confirmed the substitution of the chloride anion with AQS2 in the matrix. Molecular mechanics (MM+) calculations suggest that the optimal geometric structure (OMG) of AQS2-doped POT is at least three (3.92) times more stable than that of the parent chloride-doped POT (HCl-doped POT). The increase of the absorbance at about 840 nm associated with the increasing concentration of AQS2 revealed the insertion of AQS2 into the POT chain. This observation could be explained by the diffusion of AQS2 in the polymer chain. Kinetic parameters of the oxidative polymerization of the cationic form of o-toluidine (o-T-HClO4) in the presence of different amounts of AQS2 were deduced on the basis of absorbance variations. The results of computer-oriented kinetic analysis indicate that the rate-controlling step of the o-T polymerization is governed by the Ginstling-Bronstein equation representing the three-dimensional diffusion (D4). Activation parameters of the oxidative polymerization of protonated o-T in the presence of varying amount of AQS2 were computed and discussed.  相似文献   

3.
Protonation of poly(o‐toluidine) base form (POT‐EB) with 5‐sulfosalicylic acid (SSA) was proved experimentally and computationally. Molecular mechanics (MM+) calculations showed that the potential energy (PE) of the optimum molecular geometric structure of SSA‐doped POT is 4.703 × 103 kcal mol?1 or at least three orders of magnitude higher than the PE of the molecular geometric structure of the same matrix. These calculations indicate that the optimization of this matrix is necessary for understanding the stability. Dark green coloration (λ ~800 nm) after addition of SSA into POT‐EB matrix (dark blue, λ ~600 nm) revealed that the SSA was working as a protonating agent to convert POT base form (POT‐EB) to salt form (SSA‐doped POT). The change of the dark green color of SSA‐doped POT to dark brown (λ ~500 nm) after addition of oxidant (K2CrO4) was due to the highest oxidized form of the matrix obtained (the quinoid one), which undergoes a hydrolysis reaction to produce p‐hydroquinone (H2Q) by a mechanism similar to Schiff‐base hydrolysis. Kinetic parameters of the oxidation reaction were deduced employing a computer‐aided kinetic analysis of the absorbance (A) at ~800 nm against the hydrolysis time (t) data. The results obtained indicate that the rate controlling process may be governed by the Ginstling–Brounshetin equation for three‐dimensional diffusion (D4). The proposed mechanism for the oxidation of SSA‐doped POT matrix is also supported by MM+ calculations. Activation parameters for the rate of the oxidation process of acid‐doped POT matrix have been computed and discussed. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 35: 260–272, 2003  相似文献   

4.
3,3-Dinitroazetidinium (DNAZ) salt of perchloric acid (DNAZ·HClO4) was prepared, it was characterized by the elemental analysis, IR, NMR, and a X-ray diffractometer. The thermal behavior and decomposition reaction kinetics of DNAZ·HClO4 were investigated under a non-isothermal condition by DSC and TG/DTG techniques. The results show that the thermal decomposition process of DNAZ·HClO4 has two mass loss stages. The kinetic model function in differential form, the value of apparent activation energy (E a) and pre-exponential factor (A) of the exothermic decomposition reaction of DNAZ·HClO4 are f(α) = (1 − α)−1/2, 156.47 kJ mol−1, and 1015.12 s−1, respectively. The critical temperature of thermal explosion is 188.5 °C. The values of ΔS , ΔH , and ΔG of this reaction are 42.26 J mol−1 K−1, 154.44 kJ mol−1, and 135.42 kJ mol−1, respectively. The specific heat capacity of DNAZ·HClO4 was determined with a continuous C p mode of microcalorimeter. Using the relationship between C p and T and the thermal decomposition parameters, the time of the thermal decomposition from initiation to thermal explosion (adiabatic time-to-explosion) was evaluated as 14.2 s.  相似文献   

5.
Nitrogen‐rich heterocyclic bases and oxygen‐rich acids react to produce energetic salts with potential application in the field of composite explosives and propellants. In this study, 12 salts formed by the reaction of the bases 4‐amino‐1,2,4‐trizole (A), 1‐amino‐1,2,4‐trizole (B), and 5‐aminotetrazole (C), upon reaction with the acids HNO3 (I), HN(NO2)2 (II), HClO4 (III), and HC(NO2)3 (IV), are studied using DFT calculations at the B97‐D/6‐311++G** level of theory. For the reactions with the same base, those of HClO4 are the most exothermic and spontaneous, and the most negative ΔrGm in the formation reaction also corresponds to the highest decomposition temperature of the resulting salt. The ability of anions and cations to form hydrogen bonds decreases in the order NO3?>N(NO2)2?>ClO4?>C(NO2)3?, and C+>B+>A+. In particular, those different cation abilities are mainly due to their different conformations and charge distributions. For the salts with the same anion, the larger total hydrogen‐bond energy (EH,tot) leads to a higher melting point. The order of cations and anions on charge transfer (q), second‐order perturbation energy (E2), and binding energy (Eb) are the same to that of EH,tot, so larger q leads to larger E2, Eb, and EH,tot. All salts have similar frontier orbitals distributions, and their HOMO and LUMO are derived from the anion and the cation, respectively. The molecular orbital shapes are kept as the ions form a salt. To produce energetic salts, 5‐aminotetrazole and HClO4 are the preferred base and acid, respectively.  相似文献   

6.
Potentiometric ion sensors were prepared from the conjugated polymer poly(3-octylthiopene) (POT). The influence of additional membrane components, including silver 7,8,9,10,11,12-hexabromocarborane (AgCB11H6Br6) and potassium tetrakis[3,5-bis(trifluoromethyl)phenyl]borate (KTpFPB) as lipophilic salts, and [2.2.2]p,p,p-cyclophane as silver ionophore, was studied. The membrane components were dissolved in chloroform and membranes were prepared by solution casting on glassy carbon disk electrodes. For comparison, POT-based potentiometric sensors were also prepared by galvanostatic electrosynthesis of POT from the 3-octylthiophene monomer. All the POT-based ion sensors fabricated by solution casting show Nernstian or slightly sub-Nernstian response to Ag+, even those based only on POT without any additional membrane components. The potentiometric response of electrochemically polymerized POT depends on the film thickness and the doping anion incorporated in the conducting polymer during polymerization. It is of particular importance that chemically synthesized undoped POT (without any additives) shows a sensitive and selective potentiometric response to Ag+ ions although UV-vis results show that POT remains in its undoped form, i.e., POT is not oxidized by Ag+. This indicates that undoped POT can exhibit good sensitivity and selectivity to Ag+ also in the absence of metallic silver in the polymer film. In this case, the potentiometric response is related to interactions between Ag+ and the conjugated polymer backbone. Presented at the 4th Baltic Conference on Electrochemistry, Greifswald, 13–16, 2005  相似文献   

7.
Kinetics of oxidation of pantothenic acid (PA) by sodium N‐chloro‐p‐toluenesulfonamide or chloramine‐T (CAT) in the presence of HClO4 and NaOH (catalyzed by OsO4) has been investigated at 313 K. The stoichiometry and oxidation products are same in both media; however, their kinetic patterns were found to be different. In acid medium, the rate shows first‐order dependence on [CAT]o, fractional‐order dependence on [PA]o, and inverse fractional‐order on [H+]. In alkaline medium, the rate shows first‐order dependence each on [CAT]o and [PA]o and fractional‐order dependence on each of [OH?] and [OsO4]. Effects of added p‐toluenesulfonamide and halide ions, varying ionic strength, and dielectric constant of medium as well as solvent isotope on the rate of reaction have been investigated. Activation parameters were evaluated, and the reaction constants involved in the mechanisms have been computed. The proposed mechanisms and the derived rate laws are consistent with the observed kinetics. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 37: 201–210, 2005  相似文献   

8.
The molar conductivities (Λ) of solutions of n-tetrabutylammonium tetraphenylborate (NBu4BPh4) in 3-pentanone have been measured in the temperature range from 283.15 to 329.15 K. The conductance data have been analyzed using the Lee-Wheaton conductivity equation with the distance parameter (a) set at Bjerrum’s pairing distance, and the limiting molar conductivities (Λo) and the association equilibrium constants (K A) have been derived. The limiting ion conductivities (λ_±o) have been evaluated according to the method of Krumgalz. The λ+ o values have been compared with λ+ o values calculated from the empirical equation of Gill. The thermodynamic functions, Gibbs energy (Δ G A o), enthalpy (Δ H A o) and entropy (Δ S A o) for the process of ion-pair formation as well as the activation energy of the ionic movement (ΔH ) have been evaluated. The obtained results are discussed in terms of ion-ion and ion-solvent interactions.  相似文献   

9.
Synthesis of poly(o-anisidine) doped with various protonic acids by using ammonium persulphate as oxidizing agent were carried out in aqueous acid media. Influences of protonic acids on the physicochemical properties were investigated. The various process parameters were optimized to obtain poly(o-anisidine) in the conducting salt phase form. The results are discussed with references to different protonic acids. It was observed that poly(o-anisidine) is highly soluble in organic solvents, such as m-cresol and N-methyl pyrrolidinone (NMP). The polymers were characterized by UV-Visible, FTIR, SEM, XRD and conductivity measurements. A result shows that, different types of dopant acids HCl, H2SO4 and HClO4 affect the morphology and electrical conductivity of the polymer. The electrical conductivity of the polymer follows the order HCl >H2SO4>HClO4. Thus the effect of dopant ion type and the size of its negative ions influence the physico-chemical properties. UV-Vis absorption spectra shows peaks at 740–783 nm with shoulder at 380–420 nm as characteristic peaks for the emeraldine salt (ES) phase of poly(o-anisidine) POA. The FTIR spectra show a broad and intense band at ~2800–3001 cm?1 and ~1159–1170 cm?1 that account for the formation of ES phase of the polymer. The X-ray diffraction spectra show a characteristic peak at 20–30o, 2θ range which reveals partial crystalline structure. The conductivity of the poly(o-anisidne) salt was found to be in the range of 10?3 to 10?2 S/cm. SEM studies of poly(o-anisidine) doped with HCl shows the continuous granular uniform morphology with sub-micrometer evenly distributed particles of size ~100–200 nm.  相似文献   

10.
Polymerization of o-methylaniline or o-toluidine (OT) was studied in aqueous acidic (HCl) medium with or without the use of the support of a water soluble polymer, polyacrylamide (PAAm). Poly(o- toluidine) (POT) produced with PAAm support was in the form of aqueous solution or dispersion that showed high stability and good processibility. High degree of dispersion or near solubility and storage stability of POT thus prepared are explained on the basis of establishment of hydrogen bonding between segments of POT being formed and the PAAm present in the medium thus resulting in a template effect. Studies by UV-visible spectroscopy and FTIR spectroscopy lend support to this view. POT-PAAm composites commonly show higher thermal stability than POT and the composites show DC-electrical conductivities in the range of 10−9–10−1 S · cm−1 depending on the POT content. Morphological analysis of the optically clear aqueous POT-PAAm solution by transmission electron microscopy (TEM) shows the presence of large clusters of PAAm-supported near-spherical POT nanoparticles in the aqueous PAAm solution. Scanning electron microscopic (SEM) analysis of isolated POT-PAAm composites shows a cocontinuous phase morphology without any trend of gross phase separation. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 3243–3256, 1999  相似文献   

11.
The electrochemical and chemical polymerization of acrylamide (AA) has been studied. The electrolysis of the monomer in N,N-dimethylformamide (DMF) containing (C4H9)4NClO4 as the supporting electrolyte leads to polymer formation in both anode and cathode compartments. The cathodic polymer dissolves in the reaction mixture and the anodic polymer precipitates during the course of polymerization. A plausible mechanism for the anodic and cathodic initiation reaction has been given. The chemical polymerization of acrylamide that has been initiated by HClO4 is analogous to its anodic polymerization. The polymer yield increases with an increase in concentration of the monomer and HClO4. Raising the reaction temperature also enhances the polymerization rate. The overall apparent activation energy of the polymerization was determined to be ca. 19 kcal/mole. The copolymerization of acrylamide was carried out with methyl methacrylate (MMA) in a solution of HClO4 in DMF. The reactivity ratios are r1 (AA) = 0.25 and r2 = 2.50. The polymerization with HClO4 appears to be by a free radical mechanism. When the polymerization of acrylamide is carried out with HClO4 in H2O, a crosslinked water-insoluble gel formation takes place.  相似文献   

12.
IR spectra of soluble poly-o-toluidine (POT) and poly-m-toluidine (PMT) have been studied. Preliminary assignment of their IR spectra is given by comparing their spectra in HCl and I2, doped states and in subsequently NH3 dedoped state with that in intrinsic state.  相似文献   

13.
Silica gels doped with several protonic acids such as HClO4, H2SO4 and H3PO4 have been prepared by the sol-gel method and totally solid electric double-layer capacitors have been successfully fabricated using the highly proton-conductive silica gels as an electrolyte and activated carbon powder (ACP) hybridized with the silica gels as a polarizable electrode. It was found that the addition of HClO4, which had the highest value of acid dissociation constant among these three acids, most effectively increased the proton conductivity of the resultant acid-doped silica gels. Tablets of the HClO4-doped silica gels exhibited conductivities as high as 10–5–10–2 S cm–1 at room temperature in dry N2 atmosphere. One of the capacitors fabricated using the protonic acid-doped silica gels had a capacitance of 44 F/(gram of total ACP in the capacitor), which was comparable to those of conventional capacitors using liquid electrolytes.  相似文献   

14.
Poly(o‐methoxyaniline) (POMA) and poly(o‐toluidine) (POT) salts doped with different acids (methanesulphonic acid (MeSA), trifluoroacetic acid (TFA), and hydrochloric acid (HCl)) were synthesized by using solid‐state polymerization method. The polymers were characterized by Fourier transform infrared (FTIR) spectra, ultraviolet–visible (UV–Vis) spectrometry, X‐ray diffraction (XRD), cyclic voltammetry (CV), and conductivity measurements. Transmission electron microscopy (TEM) was done to study the morphologies of POMA and POT salts. The FTIR and UV‐Vis absorption spectra revealed that the reduced phase was predominant in POMA salts, and the pernigraniline phase was predominant in POT salts. It was found that POMA salts displayed higher doping level and conductivity. In contrast, POT salts were lower at doping levels and conductivity. In accordance with these results, the electrochemical activity was also found to be lower in POT salts. The XRD patterns showed that the POMA salts displayed higher crystallinity than POT salts. The results from TEM revealed that the morphologies of POMA salts were different from those of POT salts. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

15.
The thermal decomposition rate constant of AIBN (??d) in N,N-dimethylformamide (DMF)/methyl methacrylate (MM) mixtures of various compositions at 60°C is studied. The ??d value is 6.45 × 10?4min?1 for pure DMF and 7.20 × 10?1 min?1 for pure methyl methacrylate. The ??d values of DMF/MM mixtures were found to be dependent on the mixture composition. This dependence is not a linear function of the monomer mole fraction, but has a minimum at ca. 20 30 mol% of MM. The relationship between the AIBN decomposition rate constant and the monomer mole fraction was interpreted on the basis of solvation of the initiator molecules. © 1995 John Wiley & Sons, Inc.  相似文献   

16.
Attempts to prepare previously unknown simple and very Lewis acidic [RZn]+[Al(ORF)4]? salts from ZnR2, AlR3, and HO?RF delivered the ion‐like RZn(Al(ORF)4) (R=Me, Et; RF=C(CF3)3) with a coordinated counterion, but never the ionic compound. Increasing the steric bulk in RZn+ to R=CH2CMe3, CH2SiMe3, or Cp*, thus attempting to induce ionization, failed and led only to reaction mixtures including anion decomposition. However, ionization of the ion‐like EtZn(Al(ORF)4) compound with arenes yielded the [EtZn(arene)2]+[Al(ORF)4]? salts with arene=toluene, mesitylene, or o‐difluorobenzene (o‐DFB)/toluene. In contrast to the ion‐like EtZn(η3‐C6H6)(CHB11Cl11), which co‐crystallizes with one benzene molecule, the less coordinating nature of the [Al(ORF)4]? anion allowed the ionization and preparation of the purely organometallic [EtZn(arene)2]+ cation. These stable materials have further applications as, for example, initiators of isobutene polymerization. DFT calculations to compare the Lewis acidities of the zinc cations to those of a large number of organometallic cations were performed on the basis of fluoride ion affinity. The complexation energetics of EtZn+ with arenes and THF was assessed and related to the experiments.  相似文献   

17.
The o‐substituted hybrid phenylphosphines, PPh2(o‐C6H4NH2) and PPh2(o‐C6H4OH), could be deprotonated with LDA or n‐BuLi to yield PPh2(o‐C6H4NHLi) and PPh2(o‐C6H4OLi), respectively. When added to a solution of (η5‐C5H5)Fe(CO)2I at room temperature, these two lithiated reagents produce a chelated neutral complex 1 (η5‐C5H5)Fe(CO)[C(O)NH(o‐C6H4)PPh2C,P‐η2] for the former and mainly a zwitterionic complex 2 , (η5‐C5H5)Fe+(CO)2[PPh2(o‐C6H4O?)] for the latter. Complex 1 could easily be protonated and then decarbonylated to give 4 [(η5‐C5H5)Fe(CO){NH2(o‐C6H4)PPh2N,P‐η2}+]. Complexes 1 and 4‐I have been crystallographically characterized with X‐ray diffraction.  相似文献   

18.
Two Dy(III) complexes with benzoate derivative and 2,2′‐bipyridine ligands, [Dy(2,4‐DClBA)3bipy]2 and [Dy(o‐MOBA)3bipy]2·4H2O (2,4‐DClBA=2,4‐dichlorobenzoate; o‐MOBA=o‐methoxybenzoate; bipy=2,2′‐bipyridine), were prepared and characterized by elemental analysis, infrared spectra, ultraviolet spectra and thermogravimetry and differential thermogravimetry techniques. The thermal decomposition behavior of the two complexes under a static air atmosphere was discussed by thermogravimetry, differential thermogravimetry and infrared spectral techniques. The non‐isothermal kinetics were investigated by using a double equal‐double step method, a non‐linear isoconversional integral method and a Starink method. The mechanism functions of the first decomposition step for [Dy(2,4‐DClBA)3bipy]2 and the second decomposition step for [Dy(o‐MOBA)3bipy]2·4H2O were determined. Meanwhile, the thermodynamic parameters (ΔHne;, ΔGne; and ΔSne;) and kinetic parameters (activation energy E and the pre‐exponential factor A) for the two complexes were also calculated.  相似文献   

19.
To find the optimal way of the sample preparation of sponges for the determination of their elemental composition, three techniques of digestion were tested: acidic (mixtures of nitric, chloric, and hydrofluoric acids and hydrogen peroxide) upon heating in open vessels, autoclave-assisted at higher temperature and pressure, and decomposition in microwave ovens (MW). Using inductively coupled plasma mass spectrometry (ICP MS), we determined the concentrations of 23 elements in sponges Baikalospongia bacillifera Dubowski, 1880. The accuracy checks have demonstrated a good agreement between the measured and assured values. The attained limits of detection ranged from 32 mg/kg (for Al after autoclave digestion) to 0.1–0.2 μg/kg (for Tl, open acidic with HClO4 and MW decomposition). The relative standard deviation varied from rather modest values (for Cu, Cs, Ba, autoclave decomposition; Se, open acidic HClO4-free; Zn, open acidic with HClO4) to 30% (Be, Tl, Pb, autoclave decomposition; Be, Cr, Ga, Pb, open HClO4-free acidic digestion), 40% (Cs, MW), and 60% (Se, open acidic with HClO4). Different ways of digestion showed good compatibility of the results for the elements under study, except for Li, Sr, and Se. For Cr and Co (open acidic with HClO4), Cr, Ni, and As (MW), the method appeared insufficiently sensitive.  相似文献   

20.
With the intent to demonstrate that the charge of Z‐type ligands can be used to modulate the electrophilic character and catalytic properties of coordinated transition metals, we are now targeting complexes bearing polycationic antimony‐based Z‐type ligands. Toward this end, the dangling phosphine arm of ((o‐(Ph2P)C6H4)3)SbCl2AuCl ( 1 ) was oxidized with hydrogen peroxide to afford [((o‐(Ph2P)C6H4)2(o‐Ph2PO)C6H4)SbAuCl2]+ ([ 2 a ]+) which was readily converted into the dicationic complex [((o‐(Ph2P)C6H4)2(o‐Ph2PO)C6H4)SbAuCl]2+ ([ 3 ]2+) by treatment with 2 equiv AgNTf2. Both experimental and computational results show that [ 3 ]2+ possess a strong Au→Sb interaction reinforced by the dicationic character of the antimony center. The gold‐bound chloride anion of [ 3 ]2+ is rather inert and necessitates the addition of excess AgNTf2 to undergo activation. The activated complex, referred to as [ 4 ]2+ [((o‐(Ph2P)C6H4)2(o‐Ph2PO)C6H4)SbAuNTf2]2+ readily catalyzes both the polymerization and the hydroamination of styrene. This atypical reactivity underscores the strong σ‐accepting properties of the dicationic antimony ligand and its activating impact on the gold center.  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