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 78 毫秒
1.
The self‐reactions of the linear pentylperoxy (C5H11O2) and decylperoxy (C10H21O2) radicals have been studied at room temperature. The technique of excimer laser flash photolysis was used to generate pentylperoxy radicals, while conventional flash photolysis was used for decylperoxy radicals. For the former, the recombination rate coefficients were estimated for the primary 1‐pentylperoxy isomer (n‐C5H11O2) and for the secondary 2‐ and 3‐pentylperoxy isomers combined (“sec‐C5H11O2”) by creating primary and secondary radicals in different ratios of initial concentrations and simulating experimental decay traces using a simplified chemical mechanism. The values obtained at 298 K were: k(n‐C5H11O2+n‐C5H11O2→Products)=(3.9±0.9)×10−13 cm3 molecule−1 s−1; k(sec‐C5H11O2+sec‐C5H11O2→Products)=(3.3±1.2)×10−14 cm3 molecule−1 s−1. Quoted errors are 1σ, whereas the total relative combined uncertainties correspond to an estimated uncertainty factor around 1.65. For decylperoxy radicals, the kinetics of all the types of secondary peroxy isomers reacting with each other were considered equivalent and grouped as sec‐C10H21O2 (as for sec‐C5H11O2). The UV absorption spectrum of these secondary radicals was measured, and the combined self‐reaction rate coefficients then derived as: k(sec‐C10H21O2+sec‐C10H21O2)=(9.4±1.3)×10−14 cm3 molecule−1 s−1 at 298 K. Again, quoted errors are 1σ and the total uncertainty factor corresponds to a value around 1.75. The sec‐dodecylperoxy radical was also investigated using the same procedure, but only an estimate of the rate coefficient could be obtained, due to aerosol formation in the reaction cell: k(sec‐C12H25O2+sec‐C12H25O2)≡1.4×10−13 cm3 molecule−1 s−1, with an uncertainty factor of about 2. Despite the fairly high uncertainty factors, a relationship has been identified between the room‐temperature rate coefficient for the self‐reaction and the number of carbon atoms, n, in the linear secondary radical, suggesting: log(k(sec‐RO2+sec‐RO2)/cm3 molecule−1 s−1)=−13.0–3.2×exp(−0.64×(n‐2.3)). Concerning primary linear alkylperoxy radicals, no real trend in the self‐reaction rate coefficient can be identified, and an average value of 3.5×10−13 cm3 molecule−1 s−1 is proposed for all radicals. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet: 31: 37–46, 1999  相似文献   

2.
Paraffin gel formation in sub-sea petroleum transportation pipelines is a common problem encountered during operational and emergency production shut-down periods. Restart of a gelled oil pipeline usually requires the application of a large pressure drop across the pipeline length. In severe cases, permanent wax plugs have formed, resulting in the loss of production capacity. In this investigation, the structural breakdown of a quiescently-formed model wax-oil gel is measured at shear rates ranging from approximately 10?5 sec?1 to 1 sec?1. It is demonstrated that gel breakdown can be mathematically modeled using a time-dependent Bingham equation in which the yield stress follows third order degradation kinetics. The Bingham plastic viscosity term becomes significant at shear rates above ~10?1 sec?1 and follows a similar third order decay profile with time. When the imposed shear rate is altered in a step-wise manner during the course of the gel breakage, transient non-linear viscoelastic effects are also observed in the mechanical stress response. Finally, explicit evidence is presented which indicates that after fracture, the gel strength behaves as a point function of the absolute strain, signifying a path-independent gel structure in terms of the single-dimensional flow history.  相似文献   

3.
The kinetics of the metal exchange reaction between the Cu(II)-poly(vinyl alcohol) complex (Cu(II)-PVA) and Ca(II)-ethylenediamine-N,N,N′,N′-tetraacetic acid (Ca(II)-EDTA) were studied by mixing both solutions in a spectrophotometer at pH 9.7–11.0, at μ = 0.10(KNO3) and at 25°C. The reaction is initiated by the formation of unstable Cu(II)-H-PVA by the attack of H+ to Cu(II)-PVA, and while both ligand exchange and metal exchange steps occur, the latter may be rate-determining. The kinetic expression of this reaction was determined as -d[Cu(II)-PVA]/dt = k[Cu(II)-PVA] [H+] [PVA]/[Ca(II)-EDTA], where k = k1 + k′2[H+], k1 = 3.85 × 10−2 sec−1, k2 = k′2 · K−HCu(II)-H-PVA 9.59 × 105 1 mol−1 sec−1.  相似文献   

4.
The heterogeneous structure of poly(vinyl alcohol)-poly(sodium acrylate) composite hydrogels (PVA-PAA gels), prepared by repeatedly freezing and thawing mixed aqueous solutions of PVA and PAA, has been studied using pulsed NMR. The free induction decay (FID) signals of the transverse magnetization for the PVA-PAA gels were measured by both Carr-Purcell-Meiboom-Gill and solid-echo pulse sequences. The FID signal was decomposed into two or three decaying signals with various spin-spin relaxation times, based upon differences in the molecular mobility that reflected structural heterogeneity. The structural heterogeneity of the PVA-PAA gel is quite different from that of PVA gel free of PAA. We have made observations on “crosslinked” domains in the PVA-PAA gels. They are formed by PVA solid chains, probably phase-separated in the freezing-thawing process. The formation of crosslinked domains is influenced by the number of freezing-thawing cycles. © 1994 John Wiley & Sons, Inc.  相似文献   

5.
《Electroanalysis》2018,30(5):828-833
Ferrocene was linked to polyvinyl alcohol polymer via ethylene dioxi‐bis‐ethylamine (PV‐Fc) spacer and the synthesized redox polymer was characterised by FT‐IR and H1NMR spectroscopy. Aqueous solution of PVA‐Fc was used to prepare its silica composite with tetraethylorthosilicate (TEOS). The immobilized PVA‐Fc/silica film showed E0 about 0.245 V and apparent diffusion coefficient was measured to be 2.76×10−8 cm2 s−1. The ferrocene with flexible spacer in PVA‐Fc/silica film showed an excellent electrocatalytic activity towards the sulfite oxidation at 0.35 V, that was minimum 0.3 V negative shift of the overpotential than the bare electrode. The effect of the interfearing species like acetate, ammonium, carbonate, phosphate and sulfate has been checked for the sulfite detection. Under the optimized conditions the prepared composite showed sulfite oxidation in the linear range of 5×10−7 M to 1×10−4 M and lower detection limits of 1.5×10−7 M.  相似文献   

6.
Together with the known chromium (II)/silica gel catalyst (Phillips catalyst) for the polymerization of ethylene, two new ones have been investigated. It was found that a chromium(II)-“repoly” catalyst (prepared by short reaction of the chromium(II)/silica gel with ethylene at temperatures between 100 and 225°C) and a chromium(III)/silica gel catalyst have up to hundred times higher activity than the chromium(II) one. Activation energies were calculated as 54.6, 49.6 and 43.8 kJ per mol, respectively. The number of active sites was determined by measuring the integrated absorbance of the C? H and C?O stretching vibrations of the polymer. At low chromium concentration (0.056%) roughly 50% of all chromium was catalytically active in the case of chromium(II) and chromium(III) on silica gel. For the chromium(II)-“repoly” catalyst all chromium atoms can be active. The turnover numbers for the polymerization at 20°C were calculated as 0.1 (chromium(II)), 7.5 (chromium(II)-“repoly”) and 20 (sec?1 atm?1) (chromium(III)).  相似文献   

7.
Caralluma retrospiciens (Ehrenb) is a desert plant widely distributed in the hilly semi-desert regions of southern part of Saudi Arabia. The exudate gel (EG) from the stem of plant is occasionally used for wound healing by the people of the southern part of Saudi Arabia. This study investigated the phytochemical composition, FT-IR, GC–MS spectral analysis and in vitro antibacterial activity of the EG from the stem of C. retrospiciens (Ehrenb). The plant C. retrospiciens (Ehrenb.) was collected from the hills of Rijal Almaa, a heritage village of Saudi Arabia. The EG was isolated from the stem of C. retrospiciens (Ehrenb.). Physical parameters such as viscosity and zeta potential (ZP) were determined. Phytochemical analysis, FT-IR and GC–MS spectroscopy analysis were performed to determine the bio active constituents. The antibacterial activity of the isolated gel was performed by in vitro agar well diffusion technique. The study demonstrated that the viscosity and ZP of EG influenced the efficacy of antibacterial spectral properties. The FT-IR spectroscopy of the EG showed various functional groups at 3278.29, 2951.16, 2840.44, 2527.55, 2161.67, 1647.40, 1450.06, 1406.7, 1286, 1108.34, 5536.63 cm−1. Various pharmaceutically important chemical compounds were identified using GC–MS analysis. The bioactive compounds are “Sorbic Acid”, “Rhodopsin”, “1-Heptatriacotanol”, “Oxiraneundecanoic acid, 3-pentyl-, methyl ester, trans”, “Cholestan-3-ol, 2-methylene-, (3á,5à)”, “Benzoic acid”, “3-pentyl-, methyl ester trans”, “Hexanoic acid, 2-ethyl-, oxybis (2,1-ethanediyloxy-2,1-ethanediyl) ester”, etc. The antibacterial effect of the EG showed a wide spectrum of activity against the screened human pathogenic bacteria. The results demonstrate the bioactive principles of EG from C. retrospiciens (Ehrenb.) exerts the antibacterial properties in vitro.  相似文献   

8.
Summary: Silica gel, an important inorganic polymer with many applications, was silylated with 3-(phenylaminopropyl)trimethoxysilane (PAPTMOS) by means of a novel “low solvent” method, whereby the silane was dissolved in a small amount of methanol, mixed with silica and reaction carried out in a heated vacuum oven. Polyaniline (PANI) was grafted to the silylated silica by in situ polymerization of aniline, then dedoped with aqueous ammonia. Physically adsorbed PANI was removed from the modified silica by washing with tetrahydrofuran (THF) and N-methyl-2-pyrrolidone (NMP). The silylated, PANI-modified silica had electrical conductivity 1.2 × 10−3 S cm−1 after being re-doped with methanesulfonic acid. FTIR, elemental analysis, X-ray photoelectron spectroscopy, solid-state 13C and 29Si NMR and morphological studies by SEM confirmed successful formation of the SiO2-polyaniline hybrid material.  相似文献   

9.
Herein, we present the results obtained from our studies on supramolecular self‐assembly and molecular mobility of low‐molecular‐weight gelators (LMWGs) in organic solvents using pulsed field gradient (PFG) diffusion ordered spectroscopy (DOSY) NMR. A series of concentration‐dependent DOSY NMR experiments were performed on selected LMWGs to determine the critical gelation concentration (CGC) as well as to understand the behaviour of the gelator molecules in the gel state. In addition, variable‐temperature DOSY NMR experiments were performed to determine the gel‐to‐sol transition. The PFG NMR experiments performed as a function of gradient strength were further analyzed using monoexponential DOSY processing, and the results were compared with the automated Bayesian DOSY transformation to obtain 2D plots. Our results provide useful information on the stepwise self‐assembly of small molecules leading to gelation. We believe that the results obtained from these experiments are applicable in determining the CGC and gel melting temperatures of supramolecular gels. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

10.
《化学:亚洲杂志》2017,12(15):1920-1926
An “in situ sacrifice” process was devised in this work as a room‐temperature, all‐solution processed electrochemical method to synthesize nanostructured NiOx and FeOx directly on current collectors. After electrodepositing NiZn/FeZn bimetallic textures on a copper net, the zinc component is etched and the remnant nickel/iron are evolved into NiOx and FeOx by the “in situ sacrifice” activation we propose. As‐prepared electrodes exhibit high areal capacities of 0.47 mA h cm−2 and 0.32 mA h cm−2, respectively. By integrating NiOx as the cathode, FeOx as the anode, and poly(vinyl alcohol) (PVA)‐KOH gel as the separator/solid‐state electrolyte, the assembled quasi‐solid‐state flexible battery delivers a volumetric capacity of 6.91 mA h cm−3 at 5 mA cm−2, along with a maximum energy density of 7.40 mWh cm−3 under a power density of 0.27 W cm−3 and a maximum tested power density of 3.13 W cm−3 with a 2.17 mW h cm−3 energy density retention. Our room‐temperature synthesis, which only consumes minute electricity, makes it a promising approach for large‐scale production. We also emphasize the in situ sacrifice zinc etching process used in this work as a general strategy for metal‐based nanostructure growth for high‐performance battery materials.  相似文献   

11.
The dynamics of three glass-forming polymers, PVC, PB and PI, has been investigated by time of flight (TOF) neutron scattering in a time scale from 10−13s to 1011s looking for the crossover from microscopic dynamics to segmental dynamics (α relaxation). A new analysis procedure has been applied to TOF data in order to separate harmonic vibrational and relaxational contributions. Due to the involved assumptions, this procedure can be considered only as a first approximation adequate for the case of “fragile” systems (in the Angell's meaning) like the polymers here investigated. The behaviour obtained was the same for the three polymers studied. The intermediate scattering function corresponding to the relaxational dynamics, Io(Q,t), shows two different dynamical regimes separated by a crossover time tc (≈ 2 ps), which hardly depends on Q (momentum transfer) and temperature. At t < t<c, Io(Q,t) displays a Debye-like behaviour (exponential decay). The activation energy found for the relaxation time corresponding to this regime was in the range of 2–5 Kcal/mol, i.e., in the range of the activation energy for local conformational transitions in isolated macromolecular chains. At t > tc and, at least at high temperature, Io(Q,t) shows a Kohlrausch-Williams-Watts (KWW) behaviour similar to the obtained one by means of backscattering neutron techniques in the mesoscopic time scale (10−11s to 10−7s) and dielectric measurements in the macroscopic time scale (10−7s to 10°s). This KWW regime can be associated to the segmental dynamics involved in the α relaxation. A phenomenological interpretation is outlined. In this framework, the Debye-like regime is interpreted to be the segmental dynamics free from intermolecular hindrances. Therefore, tc should be the time at which intermolecular interactions start to play a significant role concerning to the segmental dynamics. This interpretation recalls some of the basic ideas of the so called “Coupling-Model” proposed a long time ago by Ngai.  相似文献   

12.
The galvanostatic intermittent titration technique (GITT) has been used to electrochemically determine the chemical and component diffusion coefficients, the electrical and general lithium mobilities, the partial lithium ionic conductivity, the parabolic tarnishing rate constant, and the thermodynamic enhancement factor in “Li3Sb” and “Li3Bi” as a function of stoichiometry in the temperature range from 360 to 600°C. LiCl, KCl eutectic mixtures were used as molten salt electrolytes and Al, “LiAl” two-phase mixtures as solid reference and counterelectrodes. The stoichiometric range of the antimony compound is rather small, 7 × 10?3 at 360°C, whereas the bismuth compound has a range of 0.22 (380°C), mostly on the lithium deficit side of the ideal composition. The thermodynamic enhancement factor in “Li3Sb” depends strongly on the stoichiometry, and has a peak value of nearly 70 000; for “Li3Bi” it rises more smoothly to a maximum of 360. The chemical diffusion coefficient for “Li3Sb” is 2 × 10?5 cm2 sec?1 at negative deviations from the ideal stoichiometry and increases by about an order of magnitude in the presence of excess lithium at 360°C. The corresponding value for “Li3Bi” is 10?4 cm2 sec?1 with high lithium deficit, and increases markedly when approaching ideal stoichiometry. The activation energies are small, 0.1–0.3 eV, depending on the stoichiometry, in both phases. The mobility of lithium in “Li3Bi” is about 500 times greater than in “Li3Sb” with a lithium deficit. The ionic conductivity in “Li3Sb” increases from about 10?4 Ω?1 cm?1 in the vacancy transport region to about 2 × 10?3 where transport is probably by interstial motion at 360°C. For “Li3Bi” a practically constant value of nearly 10?1 Ω?1 cm?1 is found at 380°C. The parabolic tarnishing rate constant shows a sharp increase at higher lithium activities in “Li3Sb” whereas in “Li3Bi” it has a roughly linear dependence upon the logarithm of the lithium activity. The tarnishing process is about 2 orders of magnitude slower for “Li3Sb” than for “Li3Bi.” Because of the fast ionic transport in these mixed conducting materials, “Li3Sb” and “Li3Bi” may be called “fast electrodes.”  相似文献   

13.
The decay of a N2(v = 1) molecule within a N+2n cluster, causing subsequent loss of a “magic” number of molecules from the cluster, is modeled statistically. Up to clusters of size n = 19, the losses are explicable with a constant average binding energy of 660 ± 30cm−1.  相似文献   

14.
Summary: “True” initial polymerization rates can be calculated from the adiabatic temperature rise under isoperibolic conditions. They are much higher than initial polymerization rates measured by standard mass flow meter methods under quasi‐steady‐state conditions. These high initial rates are followed by a fast apparent deactivation (“attenuation”) until a constant (“plateau”) activity is reached at low polymerization yields of 2–3 g PP · (g catalyst)−1 caused by an “increasing degree of encapsulation” of active sites. Mass transfer limitations are not observed. Cross sectional SEM images of the polymer samples support these kinetic findings.

Rate of polymerization (g of PP · (g catalyst)−1 · h−1) and yield of polymerization (g of PP · (g catalyst)−1) as function of time.  相似文献   


15.
The self‐assembly of eight PdII cations and sixteen phenanthrene‐derived bridging ligands with 60° bite angles yielded a novel M8L16 metallosupramolecular architecture composed of two interlocked D4h‐symmetric barrel‐shaped containers. Mass spectrometry, NMR spectroscopy, and X‐ray analysis revealed this self‐assembled structure to be a very large “Hopf link” catenane featuring channel‐like cavities, which are occupied by NO3 anions. The importance of the anions as catenation templates became imminent when we observed the nitrate‐triggered structural rearrangement of a mixture of M3L6 and M4L8 assemblies formed in the presence of BF4 anions into the same interlocked molecule. Furthermore, the densely packed structure of the M8L16 catenane was exploited in the preparation of a hexyloxy‐functionalized analogue, which further self‐assembled into vesicle‐like aggregates in a reversible manner.  相似文献   

16.
This paper presents results from a series of pulsed field gradient (PFG) NMR studies on lipophilic guanosine nucleosides that undergo cation‐templated assembly in organic solvents. The use of PFG‐NMR to measure diffusion coefficients for the different aggregates allowed us to observe the influences of cation, solvent and anion on the self‐assembly process. Three case studies are presented. In the first study, diffusion NMR confirmed formation of a hexadecameric G‐quadruplex [G 1 ]16 ? 4 K+ ? 4 pic? in CD3CN. Furthermore, hexadecamer formation from 5′‐TBDMS‐2′,3′‐isopropylidene G 1 and K+ picrate was shown to be a cooperative process in CD3CN. In the second study, diffusion NMR studies on 5′‐(3,5‐bis(methoxy)benzoyl)‐2′,3′‐isopropylidene G 4 showed that hierarchical self‐association of G8‐octamers is controlled by the K+ cation. Evidence for formation of both discrete G8‐octamers and G16‐hexadecamers in CD2Cl2 was obtained. The position of this octamer–hexadecamer equilibrium was shown to depend on the K+ concentration. In the third case, diffusion NMR was used to determine the size of a guanosine self‐assembly where NMR signal integration was ambiguous. Thus, both diffusion NMR and ESI‐MS show that 5′‐O‐acetyl‐2′,3′‐O‐isopropylidene G 7 and Na+ picrate form a doubly charged octamer [G 7 ]8 ? 2 Na+ ? 2 pic? 9 in CD2Cl2. The anion's role in stabilizing this particular complex is discussed. In all three cases the information gained from the diffusion NMR technique enabled us to better understand the self‐assembly processes, especially regarding the roles of cation, anion and solvent.  相似文献   

17.
High-resolution electron microscopy examination of mixed (Cs,Ba)-hollandites reveals evidence for significant short-range order in the distribution of cation vacancies in the incompletely occupied tunnels. The structure was observed to be quite unstable under 400-keV electron irradiation after only a few minutes' study at a dose rate of ca. 104 electrons Å−2 sec−1. Transformations to a microtwinned phase having interaxial angles of 80°/110° rather than 90°, and finally to an amorphous substance, were directly observed and recorded using video techniques. It is evident that hollandites are structurally, chemically, and mechanically unstable under intense electron irradiation. The significance of these observations to the possible effects of self-irradiation in SYNROC preparations, where radioactive 137Cs+1 transmites to 137Ba2+ by electron decay, is discussed.  相似文献   

18.
《Microporous Materials》1997,8(3-4):193-200
Self-diffusion and diffusive transport of propane in large (≈ 180 μm) crystals of AlPO4-5 have been studied by ZLC and tracer ZLC methods. In contrast to the results of recent studies of the self-diffusion of CH4 and CF4 in the same zeolite by the PFG NMR method the results of the present study show a rather fast one-dimensional Fickian diffusion process (D ≈ 10−5–10−6 cm2 s−1 at 300 K) with no evidence of single-file behavior. A possible explanation for this apparent discrepancy is suggested.  相似文献   

19.
A stable nonlinear optical (NLO) film containing “T” type alkoxysilane dye was prepared by sol–gel technology. This crosslinked “T” type alkoxysilane dye was synthesized and fully characterized by FTIR, UV–Vis spectra, and 1H‐NMR. Followed by hydrolysis and copolymerization processes of the alkoxysilane with γ‐glycidoxypropyl trimethoxysilane (KH560) and tetraethoxysilane (TEOS), high quality inorganic–organic hybrid second‐order NLO films were obtained by spin coating. The “T” type structure of the alkoxysilane was found to be effective for improving the temporal stability of the optical nonlinearity due to the reduction in the relaxation of the chromophore in the film materials. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

20.
The mobilities of mass-identified Ne+ and Ne+2 ions in neon have been measured by the “four-gauze” electrical shutter method at 300°K. The reduced zero-field mobilities of Ne+ and Ne+2 ions, corrected to 273°K, are 4.13± 0.04 and 6.20± 0.07 cm2V?1sec?1 respectively. The reaction rate coefficient for the termolecular ion-neutral association reaction is (4.6 ± 0.35) × 10?32 cm6 sec?1 and in the range from 2 to 10 V cm?1 torr?1 it does not depend on the electric field strength.  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