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 62 毫秒
1.
A new surface pretreatment for the preparation of wall-coated open tubular (WCOT) glass capillary columns has been evaluated. This technique involves the application of a non-extractable layer of Superox?-4, a 4,000,000 MW polyethylene glycol, to the glass surface as a pretreatment and deactivation agent. Unlike other polyethylene glycols, Superox-4 is stable at high temperatures (> 300°C) in the absence of oxygen, coats smoothly onto a bare glass surface and resists droplet formation. WCOT columns (SE-54, Carbowax 20M, and SP-2250) prepared using this technique are compared to columns prepared using a modified Grob [2,3] BaCO3 procedure. The Superox-4 pretreated columns were equal or superior in quality to the BaCO3 pretreated columns, based on the appearance of an activity standard and the effective theoretical plates (Neff) per meter. Chromatograms showing practical application of the WCOT columns prepared using both methods is given.  相似文献   

2.
Microwaves (MWs) are applied to initialize deoxygenation of graphene oxide (GO) in the solid state and at low temperatures (~165 °C). The Fourier‐transform infrared (FTIR) spectra of MW‐reduced graphene oxide (rGO) show a significantly reduced concentration of oxygen‐containing functional groups, such as carboxyl, hydroxyl and carbonyl. X‐ray photoelectron spectra confirm that microwaves can promote deoxygenation of GO at relatively low temperatures. Raman spectra and TGA measurements indicate that the defect level of GO significantly decreases during the isothermal solid‐state MW‐reduction process at low temperatures, corresponding to an efficient recovery of the fine graphene lattice structure. Based on both deoxygenation and defect‐level reduction, the resurgence of interconnected graphene‐like domains contributes to a low sheet resistance (~7.9×104 Ω per square) of the MW‐reduced GO on SiO2‐coated Si substrates with an optical transparency of 92.7 % at ~547 nm after MW reduction, indicating the ultrahigh efficiency of MW in GO reduction. Moreover, the low‐temperature solid‐state MW reduction is also applied in preparing flexible transparent conductive coatings on polydimethylsiloxane (PDMS) substrates. UV/Vis measurements indicate that the transparency of the thus‐prepared MW‐reduced GO coatings on PDMS substrates ranges from 34 to 96 %. Correspondingly, the sheet resistance of the coating ranges from 105 to 109 Ω per square, indicating that MW reduction of GO is promising for the convenient low‐temperature preparation of transparent conductors on flexible polymeric substrates.  相似文献   

3.
4.
The kinetics of decomposition of organocobalt chelates in the pH range of 2.2–7.0 has been studied. It has been shown that the rate constant of decomposition of the octyl chelate complex at 20°C changes from ~3 × 10?3 to ~6 × 10?6 s?1 in the above pH range. The rate constants of decomposition of complexes with ethyl, octyl, and cetyl ligands, as estimated at 20°C and pH 8.3, are 1.69 × 10?4, 1.39 × 10?4, and 2.42 × 10?5 s?1, respectively. As evidenced by emission spectrometry measurements, ~100% of organocobalt chelates with ethyl and isopropyl ligands occur in the aqueous phase, while organocobalt chelates with octyl and cetyl ligands are partitioned between monomer and aqueous phases. The rates of initiation of the emulsion polymerization of styrene have been measured by the inhibited polymerization procedure. It has been demonstrated that among three tested compounds (diphenyl picryl hydrazyl, hydroquinone, and benzoquinone), benzoquinone has been found to be a suitable inhibitor for the polymerization under study. The rates of initiation of styrene polymerization at 30°C for organocobalts with ethyl, octyl, and cyclohexyl ligands are 1.0 × 10?7, 1.04 × 10?7, and 3.7 × 10?6 mol/(l s), respectively. The rate constant of decomposition of the organocobalt complex with the octyl ligand at 30°C is 2.28 × 10?5 s?1, and the efficiency of initiation with this complex is 0.95.  相似文献   

5.
In this study, we experimentally obtained partitioning ratios of radioiodide (125I?) in the three phases at two different temperatures, 4 and 23 °C, in order to observe the effect of biological activity for upland soil samples. Even at 4 °C, 125I emission was found; its partitioning ratio in the gas phase ranged from 0 to 27 %. As expected, the ratio at 23 °C was higher than that at 4 °C. Additionally, in comparison of the data for upland soil samples obtained in this study and our previous data for paddy soil samples, for the latter, I? was not only sorbed in the soil but also more of it was released into the air than for upland soil samples. The land-use difference for I? partitioning in soil might be attributed to the differences of exchangeable K and stable I concentrations in soil from the statistical analysis. On the other hand, there were good correlations of partitioning ratios in solid, liquid, and gas phases between the two temperatures. The results implied that the biological activity can enhance partitioning in not only the gas phase, but also the solid phase. Indeed, the soil–soil solution distribution coefficient at 23 °C was about three times as high as that at 4 °C.  相似文献   

6.
The transport properties of separating membranes MF-4SK are studied during electrolysis of H2O in solutions of KOH. The effective diffusion coefficients of molecules of KOH and H2O and the transfer coefficients of ions K+ and OH? and molecules of H2O are measured at KOH concentrations reaching 11 M, currents reaching 0.31 A cm?2, at ambient temperature and at 80°C. In contact with a KOH solution in the concentration interval 0.1 to 11 M, the membranes that initially swelled in H2O lose a considerable fraction of water that was present in them and the overall volume of clusters and solution-filled channels in them noticeably decreases. The coefficients of transfer by current of ions K+ out of anodic compartment into cathodic and the OH? ions in the reverse direction, respectively, happen to be equal to about 0.6 and 0.4 at ambient temperature and 0.8 and 0.2 at 80°C. The coefficients of transfer of water molecules out of the anodic volume into the cathodic volume in the process of electrolysis happen to be in the limits 1.6–1.9 at ambient temperature and in the limits 2.2–2.8 at 80°C. The effective diffusion coefficients of molecules of KOH and H2O at moderate concentrations of KOH (5.6 M) amount to ~2.6 × 10?7 and 30 × 10?7 cm2s?1 at ambient temperature and ~4 × 10?7 and 61 × 10?7 cm2s?1 at 80°C, respectively. At a high concentration of KOH (~10 M) these quantities substantially diminish.  相似文献   

7.
Kinetics of the oxidation of methyl ethyl ketone (MEK) and diethyl ketone (DEK) by N-bromoacetamide (NBA) have been studied in perchloric acid media in the presence of mercuric acetate. A zero order dependence to NBA and a first-order dependence to both ketones and H+ have been observed. Acetamide, mercuric acetate and sodium perchlorate additions have negligible effect while addition of acetic acid has a positive effect on the reaction rate. A solvent isotope effect (K0D2O/k0H2O = 2.–.4 and 2.2-2.5 for MEK and DEK, respectively) has been observed at 40°. Kinetic investigations have revealed that the order of reactivity is MEK > DEK. The rates were determined at four different temperatures and the activation parameters were evaluated. The main product of the oxidation is the corresponding 1,2-diketone. A suitable mechanism consistent with the above observations has been proposed.  相似文献   

8.
The investigation of the vacuum thermal decomposition of polydiphenylenesulfophthalide at 100–530 °C showed that there are at least four main types of paramagnetic species (PMSs). The ESR spectrum of type I PMSs (120–250 °C) has a signal (g = 2.0028; ΔH ~1 mT) with a poorly resolved hyperfine structure (HFS) and an even number of lines. The electronic spectrum of these particles shows an absorption band at ~410 nm. These particles were assigned to “low-temperature” triarylmethyl-type radicals (TAMTR), which are apparently generated from dioxothioxanthene defect structures of the polymer. Type II PMSs (250–360 °C) give a smooth symmetrical ESR singlet (g = 2.0028; ΔH ~1 mT) and two absorption bands in the electronic spectrum at ~410 nm (strong band) and ~710 nm (weak band). Based on the results of calculations of electronic spectra for a series of model structures at the TD-DFT B3LYP/6-311G(d,p) level of theory, these PMSs were assigned to “high-temperature” TAMTR, which have a fluorenyl structure and are formed through the opening of the sulfophthalide ring. The maximum concentration of TAMTR II (~1020 spin g?1) is achieved at 320 °C. At T > 320 °C, type II radicals decay and type III radicals are generated. The latter are condensed aromatic species presumably having a phenalene structure. In the temperature range of 350–450 °C, the ESR line width and shape remain mainly unchanged, which attests to the retention of the dominant structure of the radicals. An increase in the thermal decomposition temperature to ~450 °C or above leads to a decrease in the ESR line width and a change in its shape from the Gaussian to Lorentzian type. This fact is an evidence of type IV paramagnetic species corresponding to even higher condensed aromatic structures.  相似文献   

9.
Recently, new additions to nonconjugated conductive polymers have been made that include: poly(β-pinene), poly(ethylenepyrrolediyl) derivative and polynorbornene. These polymers have different double-bond number fractions per repeat compared to previously reported nonconjugated conductive polymers. Upon doping with electron acceptor (e.g. iodine) the conductivities increase more than ten orders of magnitude to levels consistent with their double-bond number fractions. Radical cations that are formed upon doping have dimensions less than a nanometer leading to unexpectedly large nonlinear optical effects. The measured Kerr coefficient of doped polyisoprene is ~1.6 × 10-10 m/V2 at 633 nm. For doped poly(β-pinene) the Kerr coefficient is ~1.2 × 10-10 m/V2 and for poly(ethylenepyrrolediyl) derivative it is ~1.2 × 10-9m/V2. Exceptionally large two-photon absorption coefficient (~2.6 cm/MW at 810 nm) has been measured for doped poly(β-pinene). The mechanism of the nonlinear optical effect has been explained using a model of subnanometer size metallic quantum dots. These novel materials with wide transparency are expected to have significant applications in photonics.  相似文献   

10.
Tensile stress–strain and stress relaxation properties of 1100 equivalent weight Nafion have been measured from 23 to 120 °C at 0–100% relative humidity. At room temperature, the elastic modulus of Nafion decreases with water activity. At 90 °C, the elastic modulus goes through a maximum at a water activity of ~ 0.3. At temperatures ≥90 °C, hydrated membranes are stiffer than dry membranes. Stress‐relaxation was found to have two very different rates depending on strain, temperature, and water content. At high temperature, low water activity, and small strain, the stress relaxation displays a maximum relaxation time with stress approaching zero after 103–104 s. Water absorption slows down stress‐relaxation rates. At high water activity, the maximum stress relaxation time was >105 s at all temperatures. No maximum relaxation time was seen at T ≤ 50 °C. Increasing the applied strain also resulted in no observed upper limit to the stress relaxation time. The results suggest that temperature, absorbed water, and imposed strain alter the microstructure of Nafion inducing ordering transitions; ordered microstructure increases the elastic modulus and results in a stress relaxation time of >105 s. Loss of microphase order reduces the elastic modulus and results in a maximum stress relaxation time of 103–104 s. © 2008 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 47: 11–24, 2009.  相似文献   

11.
An indirect simple and rapid cloud point extraction is proposed for separation and preconcentration of sulfadiazine and its determination by flow injection‐flame atomic absorption spectroscopy (FI‐FAAS). The sulfadiazine from 35 mL of solution was readily converted to silver sulfadiazine upon addition of silver nitrate (9.7 × 10‐5 mol/L). Then, Triton X‐114 a non ionic surfactant was added and the solution was heated to 60 °C. At this stage, two separate phases was formed and silver sulfadiazine enters the surfactant rich phase of non‐ionic micelles of Triton X‐114. The surfactant‐rich phase (~50 μL) was then separated and diluted to 300 μL with acidic methanol. The concentration of silver in the surfactant‐rich phase which is proportional to the concentration of sulfadiazine in sample solution was determined by FI‐FAAS. The parameters affecting extraction and separation were optimized. Under the optimum conditions (i.e. pH 2‐10, silver concentration (9.7 × 10 ‐5 mol/L), Triton X‐114 (0.075% v/v) and temperature 60 °C) a preconcentration factor of 117 and a relative standard deviation of 4.9% at 37 μg L‐1 of sulfadiazine was obtained. The method was successfully applied to analysis of milk, urine and tablet samples and accuracy was determined by recovery experiments.  相似文献   

12.
It is known that by lowering the impact energy the sputter rate and surface transient width in SIMS will be reduced. However, few studies have been done at ultralow energies over a wide range of impact angles. This study examines the dependence of sputter rate and transient width as a function of O2+ primary ion energy (Ep = 250 eV, 500 eV and 1 keV) and incidence angles of 0–70°. The instrument used is the Atomika 4500 SIMS depth profiler and the sample was Si with 10 delta‐layers of Si0.7Ge0.3. We observed that the lowest transient width of 0.7 nm is obtainable at normal and near‐normal incidence with Ep ~ 250 eV and Ep ~ 500 eV. There is no significant improvement in transient width going down in energy from Ep ~ 500 to ~250 eV. The onset of roughening is also not obvious at Ep ~ 250 eV over the whole angular range studied. Although the sputter rate during the surface transient is normally different from that at steady state, only at Ep ~ 250 eV was it observed that the sputter rate remained fairly independent of depth. We conclude that the best working ranges to achieve a narrow transient width and accurate depth calibration are at Ep ~ 250 eV/0° < θ < 20°and 500 eV/0°< θ < 10°. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

13.
The precise molecular structure of the title compound has been determined by single crystal X-ray diffractometry. It consists of a cyclohexadiene ring fused at the 5 and 6 positions to a cyclobutane ring which is in turn fused to a cyclohexane ring. The two six-membered rings are trans to each other with respect to the shared four-membered ring. The Fe(CO)3 moiety is bound in the usual way to the conjugated diene portion of the cyclohexadiene ring. The feature of greatest interest is the mutual influence of the conformations of the two fused cycloalkane rings, whose intrinsically preferred conformations are mutually incompatible. Under the influence of the fused cyclohexadiene ring the C4 ring would tend to be planar, while the cyclohexane ring would tend, of itself, to have a chair conformation. The actual result is a compromise, with the C4 ring being folded by 15° along its diagonal and the C6 ring having a conformation intermediate between planarity and a chair. Crystallographic data: space group, P21, Z = 2. Unit cell dimensions at 3°C are a = 6.176(1), b = 11.307(2), c = 9.781(2) Å and β = 92.89(2)°. A set of 1733 reflections having 2θ(Mo-Kα) < 63.7° and I > σ(I) was refined to convergence (R1 = 0.055; R2 = 0.034) with hydrogen atoms refined isotropically and all others anisotropically.  相似文献   

14.
As a model system, thin films of trisilanolphenyl‐POSS (TPP) and two different number average molar mass (5 and 23 kg mol?1) poly(t‐butyl acrylate) (PtBA) were prepared as blends by Langmuir–Blodgett film deposition. Films were characterized by ellipsometry. For comparison, bulk blends are prepared by solution casting and the samples are characterized via differential scanning calorimetry. The increase in Tg as a function of TPP content for bulk high and low molar mass samples are in the order of ~10 °C. Whereas bulk Tg shows comparable increases for both molar masses (~10 °C), the increase in surface Tg for higher molar mass PtBA is greater than for low molar mass (~22 °C vs. ~10 °C). Nonetheless, the total enhancement of Tg is complete by the time 20 wt % TPP is added without further benefit at higher nanofiller loads. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2015 , 53, 175–182  相似文献   

15.
Water sorption, volumetric expansion, and proton conductivity of 1100 EW Nafion and 555 EW sulfonated polyetheretherketone (SPEEK) were compared as functions of water activity at 60 and 80 °C. Water sorption in Nafion occurs with a small positive volume of mixing, ~0.005 cm3/cm3. In contrast, water sorption in SPEEK has a large negative volume of mixing ~?0.05 cm3/cm3. The percolation thresholds for proton conduction occur at hydrophilic volume fractions of 0.10 in Nafion and 0.30 in SPEEK. Proton conductivity increases quadratically with hydrophilic volume fraction above the percolation threshold. The different percolation thresholds suggest the hydrophilic domains in Nafion grow from lamella, whereas the hydrophilic domains in SPEEK grow from spheres. © 2011 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 49: 1437–1445, 2011  相似文献   

16.
The effects of molecular characteristics and processing conditions on melt‐drawing behavior of ultrahigh molecular weight polyethylene (UHMW‐PE) are discussed, based on a combination of in situ X‐ray measurement and stress–strain behavior. The sample films of metallocene‐ and Ziegler‐catalyzed UHMW‐PEs with a similar viscosity average MW of ~107 were prepared by compression molding at 180 °C. Stress profiles recorded at 160 °C above the melting temperature of 135 °C exhibited a plateau stress region for both films. The relative change in the intensities of the amorphous scattering recorded on the equator and on the meridian indicated the orientation of amorphous chains along the draw axis with increasing strain. However, there was a substantial difference in the subsequent crystallization into the hexagonal phase, reflecting the molecular characteristics, that is, MW distribution of each sample film. Rapid crystallization into the hexagonal phase occurred at the beginning point of the plateau stress region in melt‐drawing for metallocene‐catalyzed UHMW‐PE film. In contrast, gradual crystallization into the hexagonal phase occurred at the middle point of the plateau stress region for the Ziegler‐catalyzed film, suggesting an ease of chain slippage during drawing. These results demonstrate that the difference in the MW distribution due to the polymerization catalyst system dominates the phase development mechanism during melt‐drawing. The effect of the processing conditions, that is, the including strain rate and drawing temperature, on the melt‐drawing behavior is also discussed. The obtained results indicate that the traditional temperature–strain rate relationship is effective for transient crystallization in to the hexagonal phase during melt‐drawing, as well as for typically oriented crystallization during ultradrawing in the solid state. © 2006 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 44: 2455–2467, 2006  相似文献   

17.
In this paper, activated carbon materials were synthesized from pomegranate rind through carbonization and alkaline activation processes. The effects of pyrolytic temperature on the textual properties and electrochemical performance were investigated. The surface area of the activated carbon can reach at least 2200 m2 g?1 at different pyrolytic temperatures. It was found that, at the range of 600–900 °C, decreasing the carbonization temperature leads to the increase of t-plot micropore area, t-plot micropore volume, and capacitance. Further decreasing the carbonization temperature to 500 °C also leads to the increase of t-plot micropore area and t-plot micropore volume, but the capacitance is slightly poorer. The activated carbon carbonized at 600 °C and activated at 800 °C possesses very high specific area (2931 m2 g?1) and exhibits very high capacitance (~268 F g?1 at 0.1 A g?1 and ~242 F g?1 at 1 A g?1). There is no capacitance fading after 2000th cycle.  相似文献   

18.
A gene coding for the extracellular esterase (EstK) was cloned from the psychrotrophic bacterium Pseudomonas mandelii based on its partial amino acid sequence as determined by mass spectrometry. The entire open reading frame consisting of 1,011 bp was expressed in Escherichia coli as a soluble protein and purified by nickel-chelated affinity chromatography and Capto Q column chromatography. Here, we show that the 33-kDa recombinant EstK protein (rEstKsp) had a substrate preference for esters of short-chain fatty acids, especially, p-nitrophenyl acetate. Optimum activity of rEstKsp was at pH 8.5 and 40 °C. The esterase activity remained similar from a range of 4~20 °C, but the maximum activity varied depending upon pH. With p-nitrophenyl acetate as the substrate, K M was 210 μM and k cat was 3.4 s?1. Circular dichroism and fluorescence spectroscopy results revealed that rEstKsp had a predominantly α-helical structure and maintained its folded state at 4~40 °C. Interestingly, the tertiary structure of rEstKsp was predicted based on the structures of other hyperthermophilic esterases. Our results demonstrated that both native and rEstKsp are active at low temperatures and have a unique substrate preference for p-nitrophenyl acetate.  相似文献   

19.
Abstract

Calcium phosphate glass-ceramics were prepared by reheating the glass rods between 675° and 750°C under a temperature gradient of ~100°C/cm. Floating zone method and silicon carbide furnace was made use to crystallizing the glass. The crystalline phases occurred as long fibers aligned parallel to the growth direction of the specimen. The spacings between the fibrous phases were smaller than 1 μm and very uniform.

Microstructure of the crystallized phases was examined by scanning electron microscope. Crystalline phases and crystallography were checked by X-ray diffraction.  相似文献   

20.

β-Xylosidases are critical for complete degradation of xylan, the second main constituent of plant cell walls. A minor β-xylosidase (BXYL II) from Penicillium janczewskii was purified by ammonium sulfate precipitation (30% saturation) followed by DEAE-Sephadex chromatography in pH 6.5 and elution with KCl. The enzyme presented molecular weight (MW) of 301 kDa estimated by size exclusion chromatography. Optimal activity was observed in pH 3.0 and 70–75 °C, with higher stability in pH 3.0–4.5 and half-lives of 11, 5, and 2 min at 65, 70, and 75 °C, respectively. Inhibition was moderate with Pb+2 and citrate and total with Cu+2, Hg+2, and Co+2. Partially purified BXYL II and BXYL I (the main β-xylosidase from this fungus) were individually immobilized and stabilized in glyoxyl agarose gels. At 65 °C, immobilized BXYL I and BXYL II presented half-lives of 4.9 and 23.1 h, respectively, therefore being 12.3-fold and 33-fold more stable than their unipuntual CNBr derivatives (reference mimicking soluble enzyme behaviors). During long-term incubation in pH 5.0 at 50 °C, BXYL I and BXYL II glyoxyl derivatives preserved 85 and 35% activity after 25 and 7 days, respectively. Immobilized BXYL I retained 70% activity after 10 reuse cycles of p-nitrophenyl-β-D-xylopyranoside hydrolysis.

  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