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.
Apparent molar volumes (V 2,φ ) and heat capacities (C p2,φ ) of glycine in known concentrations (1.0, 2.0, 4.0, 6.0, and 8.0 mol⋅kg−1) of aqueous formamide (FM), acetamide (AM), and N,N-dimethylacetamide (DMA) solutions at T=298.15 K have been calculated from relative density and specific heat capacity measurements. These measurements were completed using a vibrating-tube flow densimeter and a Picker flow microcalorimeter, respectively. The concentration dependences of the apparent molar data have been used to calculate standard partial molar properties. The latter values have been combined with previously published standard partial molar volumes and heat capacities for glycine in water to calculate volumes and heat capacities associated with the transfer of glycine from water to the investigated aqueous amide solutions, D[`(V)]2,tro\Delta\overline{V}_{\mathrm{2,tr}}^{\mathrm{o}} and D[`(C)]p2,tro\Delta\overline{C}_{p\mathrm{2,tr}}^{\mathrm{o}} respectively. Calculated values for D[`(V)]2,tro\Delta\overline{V}_{\mathrm{2,tr}}^{\mathrm{o}} and D[`(C)]p2,tro\Delta\overline{C}_{p\mathrm{2,tr}}^{\mathrm{o}} are positive for all investigated concentrations of aqueous FM and AM solutions. However, values for D[`(C)]p2,tro\Delta\overline{C}_{p\mathrm{2,tr}}^{\mathrm{o}} associated with aqueous DMA solutions are found to be negative. The reported transfer properties increase with increasing co-solute (amide) concentration. This observation is discussed in terms of solute + co-solute interactions. The transfer properties have also been used to estimate interaction coefficients.  相似文献   

2.
The Pb-Bi-Se system in the PbSe-Bi2Se3-Se-Se composition region was studied by measurement of concentration circuits of the type (−) PbSe(solid) liquid electrolyte, Pb2+(Pb-Bi-Se)(solid) (+) in the temperature range 300–430 K and by X-ray powder diffraction. A solid-phase equilibrium diagram was constructed, and the formation was confirmed for the ternary compounds Pb5Bi6Se14, Pb5Bi12Se23, and Pb5Bi18Se32, which belong to the homologous series [(PbSe)5] m · [(Bi2Se3)3] n . From the emf versus temperature equations, the partial thermodynamic functions [`(DG)]\overline {\Delta G}, [`(DH)]\overline {\Delta H}, [`(DS)]\overline {\Delta S} of PbSe in alloys were calculated. Based on the solid-phase equilibrium diagram from these partial molar quantities using the corresponding data for PbSe and Bi2Se3, the standard thermodynamic functions of formation and standard entropies of the above ternary compounds were calculated.  相似文献   

3.
The thermodynamic parameters D[`(H)], D[`(G)], D[`(S)]298\Delta \bar H, \Delta \bar G, \Delta \bar S_{298} and lg n resp. of the reactions indicated in the title have been computed from polarographic data. The numerical values obtained are nearly independent from the xanthate used. The overall formation constants increase as follows: Zn(II)<>相似文献   

4.
Densities, viscosities and ultrasonic speeds of sound for binary mixtures of 1,2-dimethoxyethane (DME) with benzene, toluene, chlorobenzene, benzyl chloride, benzaldehyde, nitrobenzene, and aniline are reported over the entire composition range at ambient pressure and temperature (i.e., T=298.15 K and p=1.01×105 Pa). These experimental data were utilized to derive the excess molar volumes (VmEV_{\mathrm{m}}^{\mathrm{E}}), excess viscosities (η E), and various acoustic parameters including the deviation in isentropic compressibility (Δκ S ), internal pressure (π I), and excess enthalpy (H E). From the excess molar volumes (VmEV_{\mathrm{m}}^{\mathrm{E}}), the excess partial molar volumes ([`(V)]m,1E\overline{V}_{\mathrm{m},1}^{\mathrm{E}} and [`(V)]m,2E\overline{V}_{\mathrm{m},2}^{\mathrm{E}}) and excess partial molar volumes at infinite dilution ([`(V)]m,10,E\overline{V}_{\mathrm{m},1}^{0,\mathrm{E}} and [`(V)]m,20,E\overline{V}_{\mathrm{m},2}^{0,\mathrm{E}}) were derived and discussed for each liquid component in the mixtures. The excess/deviation properties were found to be either negative or positive, depending on the molecular interactions and the nature of the liquid mixtures.  相似文献   

5.
Zusammenfassung Das Ziel der vorliegenden Untersuchung ist vor allem die Ermittlung des kristallinen und amorphen Anteils der untersuchten Proben aus dem mittleren Schwankungsquadrat der Elektronendichte, die aus der Absolutintensität der Röntgenkleinwinkelstreuung hergeleitet werden kann. Es sollen ferner die mittleren Dimensionen der kristallinen Anteile aus der Intensität des Auslaufes der Streukurven berechnet werden.
The mean square electron density fluctuations of three Nylon-6 samples (untreated, stretched, stretched and tempered) were obtained from the scattering curves. The degree of crystallinity of each sample was calculated from and from the densities of the crystalline and amorphous phases,d c =1.23 andd a =1.10, which were available from the literature. Crystallinities thus obtained were found to be as much as 50% less than those calculated from measured sample densitiesd. The discrepancy was ascribed to a small uncertainty ind a , the amorphous density: the calculated crystallinities are very sensitive to small changes in this parameter. If one sets the crystalline volume fractions obtained from equal to those obtained fromd, an equation ford a in terms of andd results, and is solved to yieldd a -values only about 1% higher than the literature value and of course consistent with both X-ray and density determinations. The volume fractions of the crystalline phases range from 20 to 30%. The possibility that the samples contained small amounts of air-filled holes may be excluded, since thed a -values (based on andd) calculated on the assumption of a volume fraction of as little as 0.3% for such holes are already much too high.The specific inner surface and an average dimension of the regions of the two phases were also determined. The crystalline regions are roughly 30 Å in size, the amorphous ones three to four times as large, and largest for the tempered sample.


Mit 3 Abbildungen  相似文献   

6.
Summary The relative partial molal thermodynamic functions 1 and have been derived from reported experimental light scattering and calorimetric measurements on micellar solutions ofn-dodecyltrimethylammonium bromide. It is shown that the magnitudes of the thermodynamic functions are considerably larger than can be accounted for by the interaction of the electrical double layers surrounding the micelles. Possible enthalpy changes arising from a change in micelle shape, size and degree of dissociation on dilution are discussed.
Zusammenfassung Die relativen molaren thermodynamischen Partialfunktionen , andS 1wurden abgeleitet aus Literaturwerten von Lichtstreuungsmessungen und kalorimetrischen Untersuchungen an mizellaren Lösungen vonn-Dodecyltrimethylammoniumbromid. Es wird gezeigt, daß die Zahlenwerte der thermodynamischen Funktionen erheblich größer sind als von der Wechselwirkung der elektrischen Doppelschichten zu erwarten ist, die die Mizellen umgeben. Mögliche Enthalpieänderungen, die sich aus Änderungen von Größe, Gestalt und Dissoziationsgrad bei Verdünnung herleiten, werden diskutiert.
  相似文献   

7.
We have combined the calcite force field of Hwang et al. (J. Phys. Chem. B 105:4,122–4,127, 2001) with the F3C water model and a hybrid Lennard-Jones/van der Waals 3-site potential for CO2 to investigate the (10[`1]4){\left({10\overline14}\right)} and ( 10[`1]0){\left( {10\overline10}\right)} cleaving surfaces of calcite under dry and wetted conditions. The wetting fluid included both pure water and water–carbon dioxide mixture. Excess surface energies and structural features of the calcite–fluid interface were analyzed, with the simulation results for the relaxed surfaces confirming the experimentally observed morphology and supporting our conclusion that the relative stability order of calcite cleaving surfaces under investigation will remain unchanged in the presence of water–carbon dioxide mixture as well.  相似文献   

8.
For any analytical system the population mean (μ) number of entities (e.g., cells or molecules) per tested volume, surface area, or mass also defines the population standard deviation $ (\sigma = \sqrt {\mu } ) For any analytical system the population mean (μ) number of entities (e.g., cells or molecules) per tested volume, surface area, or mass also defines the population standard deviation (s = ?{m} ) (\sigma = \sqrt {\mu } ) . For a preponderance of analytical methods, σ is very small relative to μ due to their large limit of detection (>102 per volume). However, in theory at least, DNA-based detection methods (real-time, quantitative or qPCR) can detect ≈ 1 DNA molecule per tested volume (i.e., μ ≈ 1) whereupon errors of random sampling can cause sample means ([`(x)] \overline x ) to substantially deviate from μ if the number of samplings (n), or “technical replicates”, per observation is too small. In this work the behaviors of two measures of sampling error (each replicated fivefold) are examined under the influence of n. For all data (μ = 1.25, 2.5, 5, 7.5, 10, and 20) a large sample of individual analytical counts (x) were created and randomly assigned into N integral-valued sub-samples each containing between 2 and 50 repeats (n) whereupon N × n = 322 to 361. From these data the average μ-normalized deviation of σ from each sub-sample’s standard deviation estimate ( sj ;  j = 1  to  N;  N = 7  [ n = 50 ]  to  180  [ n = 2 ] )\left( {s_j ;\;j = 1\;{\hbox{to}}\;N;\;N = 7\;\left[ {n = 50} \right]\;{\hbox{to}}\;180\;\left[ {n = 2} \right]} \right) was calculated (Δ). Alternatively, the average μ-normalized deviation of μ from each sub-sample’s mean estimate ([`(x)]j {\overline x_{\rm{j}}} ) was also evaluated (Δ′). It was found that both of these empirical measures of sampling error were proportional to { - 2}?{n ·m} \sqrt[{ - 2}]{{n \cdot \mu }} . Derivative (∂/∂n · Δ or Δ′) analyses of our results indicate that a large number of samplings (n ? 33±3.1) (n \approx {33}\pm {3}.{1}) are requisite to achieve a nominal sampling error for samples with a μ ≈ 1. This result argues that pathogen detection is most economically performed, even using highly sensitive techniques such as qPCR, when some form of organism cultural enrichment is utilized and which results in a binomial response. Thus, using a specific gene PCR-based (+ or −) most probable number (MPN) assay one could detect anywhere from 0.2 to 105 CFU mL−1 using 6 to 48 reactions (i.e., 8 dilutions × 6 replicates per dilution) depending on the initial concentration of the pathogen and volume sampled.  相似文献   

9.
Dilatometric measurements of excess molar volumes, VE and excess partial molar volumes, [`(V)] \texti\textE\overline V _{\text{i}}^{\text{E}} have been made for binary mixtures of acetonitrile with 1,2-ethanediol, 1,2-propanediol, 1,2-butanediol, 1,2-pentanediol, and 1,2-hexanediol at 20°C over the entire composition range. VE for acetonitrile + 1,2-ethanediol and 1,2-propanediol mixtures are negative over the entire range of mole fractions and positive values are obtained for all remaining mixtures. The results are explained in terms of dissociation of the self-associated 1,2-alkanediol molecules and the formation of aggregates between unlike molecules through O—H...N=C hydrogen bonding. From the experimental results, VE were calculated and correlated by Redlich–Kister type function in terms of mole fractions. The excess partial molar volumes were extrapolated to zero concentration to obtain the limiting values at infinite dilution, [`(V)] \texti\textE,o\overline V _{\text{i}}^{{\text{E,o}}} .  相似文献   

10.
Atom transfer radical polymerization of 2-(dimethylamino) ethylmethacrylate and styrene was carried out using tertiary bromine-terminated telechelic polyurethane as a macroinitiator. The resulting ABA-type amphiphilic, poly (2-(dimethylamino) ethylmethacrylate)-b-polyurethane-b-poly (2-(dimethylamino) ethylmethacrylate) and hydrophobic, polystyrene-b-polyurethane-b-polystyrene tri-block copolymers were characterized by spectral, thermal, and chromatographic techniques. As the conversion increases, [`(M)]\textn {\overline M_{_{\text{n}}}} also increases linearly. Theoretical M n values of the tri-block copolymers were comparable with the experimental [`(M)]\textn {\overline M_{_{\text{n}}}} values. These results show that the polymerization of styrene and 2-(dimethylamino) ethylmethacrylate occurred through controlled radical polymerization mechanism. Mole percentage of polystyrene and poly (2-(dimethylamino) ethylmethacrylate) blocks in the tri-block copolymers was calculated using proton nuclear magnetic resonance spectroscopy, and the results were comparable with the gel permeation chromatography results. The glass transition temperatures of polystyrene and poly (2-(dimethylamino) ethylmethacrylate) blocks in the tri-block copolymers appeared at 72 °C and 110 °C, respectively. These results confirm the presence of two phases in the tri-block copolymers.  相似文献   

11.
A quantitative method including peak-fitting for determination of the content of short chain branching (SCB) in ethylene/??-olefin copolymers based on differential scanning calorimetry is described. After stepwise isothermal crystallization, the fractions with similar SCB and lamellar thickness are sorted into groups. The content of each group is determined using the peak-fitting area. The statistical terms, the arithmetic mean SCB content $ \overline{C}_{\text{n}} $ , the weighted mean SCB content $ \overline{C}_{\text{w}} $ and the branching broadness index $ I = \overline{C}_{\text{w}} /\overline{C}_{\text{n}} $ are calculated. Through comparing with the SCB contents measured by 13CNMR analysis, the results show that this method can quantitatively characterize the content of SCB in ethylene/??-olefin copolymers with a high degree of accuracy.  相似文献   

12.
A new approach for determining the activation energy of amorphous alloys is developed. Setting the second order differential coefficient of heterogeneous reaction rate equation of non-isothermal heating as zero at extreme points of DSC curve, we obtain the new correlation taking form:
g1 = Lambertw( g3 e - g2 ) + g2 \gamma_{1} = Lambertw\left( {\gamma_{3} e^{{ - \gamma_{2} }} } \right) + \gamma_{2}  相似文献   

13.
The partial molar excessGibbs energies \(\Delta \overline G _{AgCl}^E \) of AgCl in the binary system AgCl?LiCl have been measured over the entire composition range at temperatures between 923.15K and 1175.15K in steps of 50K, using the reversible formation cell $${{Ag\left( s \right)} \mathord{\left/ {\vphantom {{Ag\left( s \right)} {AgCl\left( l \right)}}} \right. \kern-\nulldelimiterspace} {AgCl\left( l \right)}}---LiCl\left( l \right)/C,Cl_2 $$ The measured \(\Delta \overline G _{AgCl}^E \) values were fitted by the use of theRedlich-Kister-Ansatz for thermodynamic excess functions. The evaluatedRedlich-Kister parameters have been used to calculate the molar excessGibbs energies ΔG E and the partial molar excessGibbs energies \(\Delta \overline G _{LiCl}^E \) of LiCl. From the temperature dependence of theRedlich-Kister parameters for ΔG E the partial and integral molar heats of mixing and excess entropies were calculated. For 1073 K and the mole fractionx=0.5 the following values were obtained: $$\Delta G^E = 2130\left[ {J mol^{ - 1} } \right], \Delta H^E = 1994\left[ {J mol^{ - 1} } \right], \Delta S^E = 0.127 \left[ {J mol^{ - 1} K^{ - 1} } \right]$$   相似文献   

14.
Experimental values of the mean polarizability of molecules, [`(g)]\bar \gamma , and components of the Lorentz tensor, L j , in the nematic and smectic A phases are obtained for a homologous series of n-alkyl-p-(4-ethoxybenzylideneamino)-α-methylcinnamates. Dependences of the [`(g)]\bar \gamma and L j values on the mesophase temperature, the orientational order parameter S of molecules, and the number n in the homologous series are revealed. The quadratic dependence of [`(g)]\bar \gamma (S) in the nematic and smectic phases is established that is invariant with respect to the nematic-smectic A transition. Polarizability densities of the molecular core and the alkyl chain are found from the monotone decreasing dependence [`(g)]\bar \gamma (n)/v (where v is volume per one molecule) in the smectic phase. The presence (or absence) of the odd-even alternation of L j (n) in the nematic (smectic) phase is shown. A monotone decrease in the Lorentz tensor anisotropy L with an increase in n is revealed in the smectic phase, and limiting values L j (n → ∞) are determined.  相似文献   

15.
The molar enthalpies of solution of 2-aminopyridine at various molalities were measured at T=298.15 K in double-distilled water by means of an isoperibol solution-reaction calorimeter. According to Pitzer’s theory, the molar enthalpy of solution of the title compound at infinite dilution was calculated to be DsolHm = 14.34 kJ·mol-1\Delta_{\mathrm{sol}}H_{\mathrm{m}}^{\infty} = 14.34~\mbox{kJ}\cdot\mbox{mol}^{-1}, and Pitzer’s ion interaction parameters bMX(0)L, bMX(1)L\beta_{\mathrm{MX}}^{(0)L}, \beta_{\mathrm{MX}}^{(1)L}, and CMXfLC_{\mathrm{MX}}^{\phi L} were obtained. Values of the relative apparent molar enthalpies ( φ L) and relative partial molar enthalpies of the compound ([`(L)]2)\bar{L}_{2}) were derived from the experimental enthalpies of solution of the compound. The standard molar enthalpy of formation of the cation C5H7N2 +\mathrm{C}_{5}\mathrm{H}_{7}\mathrm{N}_{2}^{ +} in aqueous solution was calculated to be DfHmo(C5H7N2+,aq)=-(2.096±0.801) kJ·mol-1\Delta_{\mathrm{f}}H_{\mathrm{m}}^{\mathrm{o}}(\mathrm{C}_{5}\mathrm{H}_{7}\mathrm{N}_{2}^{+},\mbox{aq})=-(2.096\pm 0.801)~\mbox{kJ}\cdot\mbox{mol}^{-1}.  相似文献   

16.
The work presents the results of studies of the 2PbTe–AgSbTe2 system using the EMF technique with the Ag4RbI5 solid electrolyte in the temperature range of 300–430 K. Formation of a wide (0–80 mol % AgSbTe2) region of PbTe-based solid solutions in the system is shown. Partial thermodynamic functions \(\left( {\overline {\Delta G} ,\overline {\Delta H} ,\overline {\Delta S} } \right)\) of alloyed silver are calculated on the basis of the equations of the temperature dependences of EMF. Potential-forming reactions are determined on the basis of the data on solid-phase equilibriums in the Ag2Te–PbTe–Sb2Te3–Te system that are used to calculate standard thermodynamic functions of formation and standard entropies of (2PbTe) x (AgSbTe2)1–x (x = 0.2; 0.4; 0.6; 08; 0.9) solid solutions.  相似文献   

17.
The concentration curve of mean activity coefficients to the required power was approximated by a product function. The product function parameters were optimized by experimental data for the mean activity coefficients using a nonlinear regression model. Assuming that the product function parameters can be determined, the factor functions are clearly known. The mathematical complexity and a concept solution are presented. Clear, reliable results were obtained with the help of asymptotic theory when corresponding approximations were used. The method described makes it possible to split the experimentally determinable concentration curve of the mean activity coefficients to the required power in individual factor functions of complementary ion species, [`(g)]\textC(m) {\bar{\gamma }_{\text{C}}}(m) and [`(g)]\textA(m) {\bar{\gamma }_{\text{A}}}(m) . The results are verified by comparing them with experimentally determined quotients of single-ion activity coefficients of ternary systems. The calculated individual parts for single-ion species are plausible and show a characteristic, typical concentration curve for cations as well as for anions. They correlate with the ion parameters.  相似文献   

18.
Densities, ??, and viscosities, ??, of binary mixtures of 2-methyl-2-propanol with acetone (AC), ethyl methyl ketone (EMK) and acetophenone (AP), including those of the pure liquids, were measured over the entire composition range at 298.15, 303.15 and 308.15?K. From these experimental data, the excess molar volume $V_{\mathrm{m}}^{\mathrm{E}}$ , deviation in viscosity ????, partial and apparent molar volumes ( $\overline{V}_{\mathrm{m},1}^{\,\circ }$ , $\overline{V}_{\mathrm{m},2}^{\,\circ }$ , $\overline{V}_{\phi ,1}^{\,\circ}$ and $\overline{V}_{\phi,2}^{\,\circ} $ ), and their excess values ( $\overline{V}_{\mathrm{m},1}^{\,\circ \mathrm{E}}$ , $\overline{V}_{\mathrm{m,2}}^{\,\circ \mathrm{ E}}$ , $\overline {V}_{\phi \mathrm{,1}}^{\,\circ \mathrm{ E}}$ and $\overline{V}_{\phi \mathrm{,2}}^{\,\circ \mathrm{ E}}$ ) of the components at infinite dilution were calculated. The interaction between the component molecules follows the order of AP > AC > EMK.  相似文献   

19.
The irreducible spherical and Cartesian tensors built of the products of two interaction tensors: the second order tensor resulting from the product of two second order tensors \sf Ta l \sf Tl b{{\sf T}_{\alpha\,\lambda}\,{\sf T}_{\lambda\,\beta}} contracted once with the index λ, third order tensor \sf Ta b l \sf Tl g{{\sf T}_{\alpha\,\beta\,\lambda} {\sf T}_{\lambda\,\gamma}} appearing as a product of the third order interaction tensor \sf Ta b l{{\sf T}_{\alpha\,\beta\,\lambda}} and the second order one \sf Tl g{{\sf T}_{\lambda\,\gamma}} contracted once with the index λ and the fourth order product of two second order tensors \sf Ta b \sf Tgd{{\sf T}_{\alpha\,\beta}\,{\sf T}_{\gamma\delta}}, have been considered. This type of products is encountered, e.g., within the London’s dispersive energy formula, inside the second-order virial coefficients of many physical parameters such as the dielectric constant, the Kerr constant, the induced polarizability and hyperpolarizability of a pair of molecules and in other induced quantities. Our results are applied explicitly to the excess induced first and second pair hyperpolarizability.  相似文献   

20.
The molar conductivities (Λ) of solutions of bis(2,2′-bipyridine)bis(thiocyanate)chromium(III) triiodide [CrIII(bipy)2(SCN)2]I3 (where bipy denotes 2,2′-bipyridine, C10H8N2), [ _3^-\mathrm{A}^{+}\mathrm{I}_{3}^{-} ], were measured in acetonitrile (ACN) at the temperatures 294.15, 299.15, and 305.15 K. In addition, cyclic voltammograms (CVs) of [ A+I3-\mathrm{A}^{+}\mathrm{I}_{3}^{-} ] were recorded on platinum, gold, and glassy carbon working electrodes in ACN, using n-tetrabutylammonium hexafluorophosphate (NBu4PF6) as the supporting electrolyte, at scan rates (v) ranging from 0.05 to 0.12 V⋅s−1. Furthermore, electrochemical impedance spectroscopic (EIS) measurements were carried out in the frequency range 50 Hz<f<50 kHz using these three working electrodes. The measured molar conductivities (Λ) demonstrate that [ A+I3-\mathrm{A}^{+}\mathrm{I}_{3}^{-} ] behaves as uni-univalent electrolyte in ACN over the investigated temperature range. The Λ values were analyzed by means of the Lee-Wheaton conductivity equation in order to estimate the limiting molar conductivities (Λ o), as well as the thermodynamic association constants (K A), at each experimental temperature for formation of [A+ I3-\mathrm{I}_{3}^{-} ] ion-pairs. The limiting ionic conductivities ( l±o\lambda_{\pm}^{\mathrm{o}} ), the diffusion coefficients at infinite dilution (D ±), as well as the Stokes’ radii (r St) were determined for both A+ and I3-\mathrm{I}_{3}^{-} ions. The thermodynamic parameters for the ionic association process, i.e. the Gibbs energy ( DGAo\Delta G_{\mathrm{A}}^{\mathrm{o}} ), enthalpy ( DHAo\Delta H_{\mathrm{A}}^{\mathrm{o}} ), and entropy ( DSAo\Delta S_{\mathrm{A}}^{\mathrm{o}} ), were also determined. The mobility and diffusivity of the A+ ion increase linearly with increasing temperature because the solvent medium becomes less viscous as the temperature increases. The K A values indicate that significant ion association occurs that is not influenced by temperature changes. The ion-pair formation process is exothermic ( DHAo < 0\Delta H_{\mathrm{A}}^{\mathrm{o}}<0 ), leading to the generation of additional entropy ( $\Delta S_{\mathrm{A}}^{\mathrm{o}}>0$\Delta S_{\mathrm{A}}^{\mathrm{o}}>0 ). As a result, the Gibbs energy DGAo\Delta G_{\mathrm{A}}^{\mathrm{o}} is negative ( DGAo < 0\Delta G_{\mathrm{A}}^{\mathrm{o}}<0 ) and the formation of [A+I3-][\mathrm{A}^{+}\mathrm{I}_{3}^{-}] becomes favorable. CV studies on [A+I3-][\mathrm{A}^{+}\mathrm{I}_{3}^{-}] solutions indicated that the redox pair Cr3+/2+ appears to be quasi-reversible on a glassy carbon electrode but is completely irreversible on platinum and gold electrodes. EIS experiments confirm that, among these three electrodes, the glassy carbon working electrode has the smallest resistance to electron transfer.  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