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.
Electrochemical impedance spectroscopy (EIS) studies were performed to analyze the passive properties of tantalum and niobium oxides films potentiostatically formed in a 0.1 M KOH solution. The quantitative characterization of these passive materials was carried out through a transfer function previously developed by our research group, which is based on the point defect model (PDM) framework considering the formation of molecular hydrogen. According to the PDM prediction criteria, Ta2O5 and Nb2O5 films exhibited an inherent n-type semiconductor behavior, which was confirmed by the parameters obtained from the fit to the transfer function. The diffusion coefficients of the oxygen vacancies were 0.53 ± 0.14 × 10−16 and 2.18 ± 0.14 × 10−16 cm2 s−1, for Ta2O5 and Nb2O5, respectively. And a slight increase of the corresponding hydroxyl vacancies diffusion (2.73 ± 0.02 and 2.23 ± 0.65 × 10−16 cm2 s−1) was obtained, suggesting the favorable diffusion of these defects due to the alkaline conditions.  相似文献   

2.
The electrochemical behaviors of magnolol have been studied at glassy carbon electrode using cyclic voltammetry, linear sweep voltammetry and chronocoulometry. Moreover, its interaction with DNA was investigated in solution by electrochemical methods and ultraviolet–visible spectroscopy. The experiment results indicated that the electrochemical oxidation of magnolol was an irreversible process with one proton and one electron transfer. The electron transfer coefficient (α) was calculated to be 0.441 ± 0.001. At the scan rate from 100 mV/s to 450 mV/s, the electrode process was controlled by the adsorption step and at the range of 600–950 mV/s the electrochemical oxidation was diffusion controlled process. The corresponding electrochemical rate constant (ks) was 0.0760 ± 0.0001 s−1. Through chronocoulometry experiment, the diffusion coefficient (D) and the surface concentration (Γ) were obtained as (3.76 ± 0.01) × 10−7 cm2/s and (2.98 ± 0.01) × 10−10 mol/cm2. In addition, the interaction of magnolol and DNA was ascribed to be electrostatic interaction and the calculated association constant (β) and Hill coefficient (m) were 1.14 × 105 M−1 and 0.973. At last a sensitive and convenient electrochemical method was proposed for the determination of magnolol.  相似文献   

3.
The in situ formation and electrochemical behaviour of a complex between gadolinium(III) and 1,10-phenanthroline (phen) are discussed. The electron withdrawing ability of phen decreases the redox potential of Gd(III). A cathodic peak at ∼−1.7 V vs. Ag/Ag+ 10 mM, corresponds to an irreversible electron transfer reaction involving the exchange of two electrons, where the first charge transfer reaction is the rate-determining step. The diffusion coefficient and the total number of electrons were derived using the Shoup–Zsabo equation. Kinetic parameters such us k0, α and i0 are calculated. Voltammetric simulations are carried out to evaluate the accuracy of the results. A fast chemical reaction consumes the reduced species formed at ∼−1.7 V, which is then oxidised at ∼−0.1 V. The stoichiometry of the complex is determined by the Yoe and Jones’ method. An average value for the stability constant of the complex of (2.3 ± 0.8) 105 M−2 is obtained.  相似文献   

4.
Functionalized poly N,N-dimethylaniline film was prepared by adsorption of ferrocyanide onto the polymer forming at the surface of carbon paste electrode (CPE) in aqueous solution. The electrocatalytic ability of poly N,N-dimethylaniline/ferrocyanide film modified carbon paste electrode (PDMA/FMCPE) was demonstrated by oxidation of l-cysteine. Cyclic voltammetry and chronoamperometry techniques were used to investigate this ability. In the optimum pH (6.00), the electrocatalytic ability about 480 mV and the catalytic reaction rate constant, (kh), can be seen 3.08 × 103 M−1 s−1. The catalytic oxidation peak current determined by cyclic voltammetry method was linearly dependent on the l-cysteine concentration and the linearity range obtained was 8.00 × 10−5 –2.25 × 10−3 M. Detection limit of this method was determined as 6.17 × 10−5 M (2σ). At a fixed potential under hydrodynamic conditions (stirred solution), the calibration plot was linear over the l-cysteine concentration range 7.40 × 10−6 M–1.38 × 10−4 M. The detection limit of the method was 6.38 × 10−6 M (2σ).  相似文献   

5.
In the present paper, the use of a gold electrode modified by 2-(2,3-dihydroxy phenyl)-1,3-dithiane self-assembled monolayer (DPDSAM) for the determination of epinephrine (EP) and uric acid (UA) was described. Initially, cyclic voltammetry was used to investigate the redox properties of this modified electrode at various scan rates. The apparent charge transfer rate constant, ks, and transfer coefficient, α, were calculated. Next, the mediated oxidation of EP at the modified electrode was described. At the optimum pH of 8.0, the oxidation of EP occurs at a potential about 155 mV less positive than that of an unmodified gold electrode. The values of electron transfer coefficients (α = 0.356), catalytic rate constant (k = 1.624 × 104 M−1 s−1) and diffusion coefficient (D = 1.04 × 10−6 cm2 s−1) were calculated for EP, using electrochemical approaches. Based on differential pulse voltammetry, the oxidation of EP exhibited a dynamic range between 0.7 and 500.0 μM and a detection limit (3σ) of 0.51 μM. Furthermore, simultaneous determination of EP and UA at the modified electrode was described. Finally, this method was used for the determination of EP in EP ampoule.  相似文献   

6.

Objectives

This work is concerned with the study of the sorption and desorption process of water, ethanol or ethanol/water solution 50% (v/v) or 75% (v/v) by the dental resins prepared by light curing of Bis-GMA, Bis-EMA, UDMA, TEGDMA or D3MA.

Methods

A thin resin disc is placed in a bath of time to obtain the sorption curve mt = f(t). Then the liquid is desorbed until a constant mass for the disc is reached and the desorption curve is recorded. These experimental curves help in the determination of the sorbed/desorbed liquid amount at equilibrium, the percentage of the extracted mass of unreacted monomer known as “solubility”, and the sorption/desorption diffusion coefficient which expresses correspondingly the rate of the liquid sorption/desorption.

Results

The highest liquid uptake by dental resins was 13.3 wt% ethanol for Bis-GMA-resin, 12.0 wt% ethanol for UDMA-resin, 10.10 wt% ethanol/water solution for TEGDMA-resin, 7.34 wt% ethanol for D3MA-resin and 6.61 wt% ethanol for Bis-EMA-resin. The diffusion coefficient for all resins was higher in water than in ethanol/water solution or ethanol. Bis-GMA-resin showed the highest diffusion coefficient (11.01 × 10−8 cm2 s−1) followed by Bis-EMA-resin (7.43 × 10−8 cm2 s−1), UDMA-resin (6.88 × 10−8 cm2 s−1), D3MA-resin (6.22 × 10−8 cm2 s−1) and finally by TEGDMA-resin (1.52 × 10−8 cm2 s−1).

Significance

All studied dental resins, except TEGDMA-resin, absorbed higher amount of pure ethanol than water or ethanol water solution. TEGDMA-resin absorbed higher amount of ethanol/water solution (50/50 or 75/25 (v/v)) than water or ethanol. For all studied dental resins the diffusion coefficient was higher in water than in ethanol/water solution or ethanol.  相似文献   

7.
Voltammetric and electrochemical impedance spectroscopic (EIS) studies of generation one poly(propylene imine) (G1 PPI) dendrimer as an electroactive and catalytic nanomaterials both in solution and as an electrode modifier based on a simple one step electrodeposition method is presented. The G1 PPI exhibited a reversible one electron redox behaviour at E0′ ca 210 mV in phosphate buffer pH 7.2 with diffusion coefficient and Warburg coefficient of 7.5 × 10−10 cm2 s−1 and 8.87 × 10−4 Ω s−1/2 respectively. Cyclic voltammetric electrodeposition of a monolayer of G1 PPI on glassy carbon electrode was carried out between −100 mV and 1100 mV for 10 cycles. The nanoelectrode was electroactive in PBS at E0′ ca 220 mV. Kinetic profiles such as time constant (4.64 × 10−5 s rad−1), exchange current (1.55 × 10−4 A) and heterogeneous rate constant (4.52 × 10−3 cm s−1) obtained from EIS showed that the dendrimer layer catalysed the redox reaction of Fe2+/3+ in [Fe(CN)6]3−/4− redox probe.  相似文献   

8.

Objectives

Sn2+ has promising erosion-inhibiting properties in solutions, but little is known about respective effects in toothpastes. In addition, biopolymers might have protecting potential. Aim of this study was to investigate the effects of Sn2+ in toothpastes and of a biopolymer (chitosan) added to a Sn2+ formulation on erosion/abrasion.

Methods

Enamel samples were subjected to cyclic erosion procedures (10 days; 0.50% citric acid, pH 2.5; 6× 2 min/day), and brushing (2× 15 s/day, load 200 g) during immersion in slurries (2 min). The toothpastes were NaF formulations (NaF/1, NaF/2, NaF/3) and Sn2+ formulations (NaF/SnCl2, AmF/SnF2, AmF/NaF/SnCl2) and AmF/NaF/SnCl2 + 0.5% chitosan.Declared concentrations of active ingredients in toothpastes were 1400–1450 μg/g F and 3280–3500 μg/g Sn2+. Negative controls were erosion only and placebo, positive control was a SnF2 gel. Tissue loss was quantified profilometrically, Sn on enamel surfaces was measured by energy dispersive X-ray spectroscopy.

Results

Loss values (μm) for erosion only and placebo were 14.4 ± 4.5 and 20.2 ± 3.8, respectively, and 4.6 ± 1.9 for the positive control (p ≤ 0.001 each compared to erosion only). The other loss values were: NaF/1 16.5 ± 3.0, NaF/2 14.0 ± 2.7, NaF/3 12.6 ± 3.9, NaF/SnCl2 14.7 ± 5.1, AmF/SnF2 13.5 ± 4.8, AmF/NaF/SnCl2 12.4 ± 4.2, AmF/NaF/SnCl2 + chitosan 6.6 ± 3.5 (except NaF/1 all p ≤ 0.01 compared to placebo). AmF/NaF/SnCl2/chitosan was more effective than all other toothpastes (p ≤ 0.01 each). Sn on the enamel surface ranged between 1.3 ± 0.3 and 2.8 ± 0.04 wt.% with no obvious relationship with efficacy.

Conclusions

The NaF and Sn2+ toothpastes without chitosan exhibited similar anti-erosion and abrasion-prevention effects. The experimental Sn2+ formulation with chitosan revealed promising results similar to those of the positive control.

Clinical significance

NaF toothpastes offer a degree of protection against erosion/abrasion, which is likely sufficient for most subjects with average acid exposures. For patients with initial erosive lesions, however, more effective toothpaste is desirable. The combination of Sn2+ and a biopolymer appears promising in this context.  相似文献   

9.
Saliva plays a critical role in the protection of oral hard and soft tissues and contains a multitude of constituents with well-characterized biological activities in vitro. Among these are histatins and acidic proline-rich proteins (PRPs). Nevertheless, few functional studies have recognized the structural instability of these proteins in the proteolytic environment of whole saliva. The aim of this investigation was to determine histatin and acidic PRP levels in parotid secretion (PS) and in whole saliva (WS) as well as to establish their susceptibility to proteolysis in these salivary fluids. Using cationic polyacrylamide gel electrophoresis and densitometric analysis the average total histatin concentration (histatin 1 + 3 + 5) in WS was determined to be 33.3 ± 16.7 μg/ml (n = 22) and the average total acidic PRP concentration (PRP1/PIF-s + PRP3/PIF-f) was 427.9 ± 123.3 μg/ml (n = 22). Histatin and acidic PRP concentrations in PS were 6 and 1.5 times higher than in WS (n = 7), respectively. WS histatin and acidic PRP levels each correlated significantly with WS total protein concentrations (P < 0.01 and P < 0.05, respectively), as well as with each other (P < 0.01). Stability studies of histatin 3 and PRP1/Pif-s in PS revealed t1/2 times of 7.2 ± 5.5 and 50.3 ± 24.8 h, respectively (n = 7). Histatin 3 (40 μg/ml) and PRP1 (400 μg/ml), added to WS in concentrations equivalent to their concentrations in PS, disappeared at a much faster rate, with t1/2 values of 1.7 ± 1.6 min and 29.3 ± 15.3 min, respectively (n = 7). The data indicate that proteolysis in WS is an important factor in explaining the substantially lower concentrations of histatins and acidic PRPs in WS as compared to in glandular secretions.  相似文献   

10.
In this study, the electrochemical behavior of thianthrene (TH) and its application toward the electrocatalytic oxidation of guanosine (Gs) and DNA in a non-aqueous solution are investigated using different voltammetric techniques. Guanosine and DNA are adsorbed on the glassy carbon electrode (GCE) by applying a positive potential to the GCE. The rate constant of catalytic reaction between DNA and TH and also between Gs and TH were evaluated using chronoamperometry which gave rate constants of 2.41 × 106 cm3 mol−1 s−1 and 2.68 (±0.19) × 106 cm3 mol−1 s−1, respectively. Also the diffusion coefficient of TH was obtained using hydrodynamic voltammetry (3.17 × 10−5 cm2 s−1). Furthermore, using hydrodynamic voltammetry, a one-electron mechanism for oxidation of Gs is suggested.  相似文献   

11.
A stable modified glassy carbon electrode based on the poly 3-(5-chloro-2-hydroxyphenylazo)-4,5-dihydroxynaphthalene-2,7-disulfonic acid (CDDA) film was prepared by electrochemical polymerization technique to investigate its electrochemical behavior by cyclic voltammetry. The properties of the electrodeposited films, during preparation under different conditions, and their stability were examined. The homogeneous rate constant, ks, for the electron transfer between CDDA and glassy carbon electrode was calculated as 5.25(±0.20) × 102 cm s−1. The modified electrode showed electrocatalytic activity toward ascorbic acid (AA), dopamine (DA), and uric acid (UA) oxidation in a buffer solution (pH 4.0) with a diminution of their overpotential of about 0.12, 0.35, and 0.50 V for AA, DA, and UA, respectively. An increase could also be observed in their peak currents. The modified glassy carbon electrode was applied to the electrocatalytic oxidation of DA, AA, and UA, which resolved the overlapping of the anodic peaks of DA, AA, and UA into three well-defined voltammetric peaks in differential pulse voltammetry (DPV). This modified electrode was quite effective not only for detecting DA, AA, and UA, but also for simultaneous determination of these species in a mixture. The separation of the oxidation peak potentials for ascorbic acid–dopamine and dopamine–uric acid were about 0.16 V and 0.17 V, respectively. The final DPV peaks potential of AA, DA and UA were 0.28, 0.44, and 0.61 V, respectively. The calibration curves for DA, AA, and UA were linear for a wide range of concentrations of each species including 5.0–240 μmol L−1 AA, 5.0–280 μmol L−1 DA, and 0.1–18.0 μmol L−1 UA. Detection limits of 1.43 μmol L−1 AA, 0.29 μmol L−1 DA and 0.016 μmol L−1 UA were observed at pH 4. Interference studies showed that the modified electrode exhibits excellent selectivity toward AA, DA, and UA.  相似文献   

12.

Objectives

To evaluate the efficacy of simplified dehydration protocols, in the absence of tubular occlusion, on bond strength and interfacial nanoleakage of a hydrophobic experimental adhesive blend to acid-etched, ethanol-dehydrated dentine immediately and after 6 months.

Methods

Molars were randomly assigned to 6 treatment groups (n = 5). Under pulpal pressure simulation, dentine crowns were acid-etched with 35% H3PO4 and rinsed with water. Adper Scotchbond Multi-Purpose was used for the control group. The remaining groups had their dentine surface dehydrated with ethanol solutions: group 1 = 50%, 70%, 80%, 95% and 3 × 100%, 30 s for each application; group 2 the same ethanol sequence with 15 s for each solution; groups 3, 4 and 5 used 100% ethanol only, applied in seven, three or one 30 s step, respectively. After dehydration, a primer (50% BisGMA + TEGDMA, 50% ethanol) was used, followed by the neat comonomer adhesive application. Resin composite build-ups were then prepared using an incremental technique. Specimens were stored for 24 h, sectioned into beams and stressed to failure after 24 h or after 6 months of artificial ageing. Interfacial silver leakage evaluation was performed for both storage periods (n = 5 per subgroup).

Results

Group 1 showed higher bond strengths at 24 h or after 6 months of ageing (45.6 ± 5.9a/43.1 ± 3.2a MPa) and lower silver impregnation. Bond strength results were statistically similar to control group (41.2 ± 3.3ab/38.3 ± 4.0ab MPa), group 2 (40.0 ± 3.1ab/38.6 ± 3.2ab MPa), and group 3 at 24 h (35.5 ± 4.3ab MPa). Groups 4 (34.6 ± 5.7bc/25.9 ± 4.1c MPa) and 5 (24.7 ± 4.9c/18.2 ± 4.2c MPa) resulted in lower bond strengths, extensive interfacial nanoleakage and more prominent reductions (up to 25%) in bond strengths after 6 months of ageing.

Conclusions

Simplified dehydration protocols using one or three 100% ethanol applications should be avoided for the ethanol-wet bonding technique in the absence of tubular occlusion, as they showed decreased bond strength, more severe nanoleakage and reduced bond stability over time.  相似文献   

13.
Surface oxides on a Pt-Au alloy in 0.05 M KOH are formed potentiostatically at potentials, varying between 2.00 V and 2.50 V for polarization times, tp, up to 1000 s at temperature values in the range from 293 K to 308 K. This procedure results in the deposition of β-oxide films whose charge densities, qox, do not exceed 5.848 mC cm−2. Cyclic voltammetry reveals three oxide-reduction peaks whose potential values depend on the polarization potential and time but not on the temperature of the electrolyte. The temperature increase leads only to augmentation of the oxide thickness. The linear plots of 1/qox versus log tp obtained for different polarization potential and temperature values show that the oxide growth follows inverse-logarithmic kinetics and hence is limited by the rate of escape of Pt2+ and Au2+ cations from the metal into the oxide at the inner metal/oxide interface. The experimental data treatment provides to evaluate the potential drop across the oxide and the intensity of the electric field controlling the oxide layer growth. The values obtained are compared to those found for β-oxide films formation on Pt and Au under identical conditions. The electric field values for the oxide growth on the alloy electrode at 2.5 V are within the range from 0.165 × 109 V m−1 to 0.180 × 109 V m−1 which refer to those for the process on Au and Pt, respectively.  相似文献   

14.
Potentiometric electrodes based on the incorporation of surfactant-modified zeolite Y (SMZ) particles into poly vinyl chloride (PVC) membranes were described. The electrode characteristics were evaluated regarding the response towards perchlorate ions. PVC membranes plasticized with dioctyl phthalate and without lipophilic additives (co-exchanger) are used throughout this study. The influence of membrane composition on the electrode response was studied. The electrode exhibited a Nernstian response towards perchlorate in the concentration range of 7.9 × 10−6–8.0 × 10−2 M with a slope of 59.7 ± 0.9 mV per decade of perchlorate concentration with a working pH range of 1.7–9.5 with a fast response time of ≤10 s. The lower and upper detection limits were 4.07 × 10−7 and 0.13 M, respectively. The electrode response to perchlorate remains constant in the temperature range of 20–40 °C and in the presence of 2.5 × 10−6–1 × 10−2 M NaNO3. The selectivity coefficients for perchlorate anion as test species with respect to other anions were determined. The proposed modified zeolite-PVC electrode can be used for at least 30 days without any considerable divergence in potential. It was applied as indicator electrode in water samples with satisfactory results. The results of this study and our previous work show HDTMA plays different roles according to the zeolite type and matrix, as HDTMA-zeolite Y in a carbon paste matrix showed a good Nernstian behavior towards phosphate anion.  相似文献   

15.
In this study we investigated the electrocatalytic oxidation of cysteine, cystine, N-acetyl cysteine, and methionine on cobalt hydroxide nanoparticles modified glassy carbon electrode in alkaline solution. Different electrochemical techniques such as cyclic voltammetry, chronoamperometry and steady-state polarization were used to track the oxidation process and its kinetics. From voltammetric studies we concluded that in the presence of amino acids the anodic peak current of Co(IV) species increased, followed by a decrease in the corresponding cathodic current peak. This indicates that amino acids were oxidized on the redox mediator which was immobilized on the electrode surface via an electrocatalytic mechanism. Using Laviron’s equation, the values of αs and ks for the immobilized redox species were determined as αs,a = 0.63, αs,c = 0.38 and ks = 0.28 s−1, respectively. The catalytic rate constants, the electron transfer coefficients and the diffusion coefficients involved in the electrocatalytic oxidation of amino acids were determined.  相似文献   

16.
An ionic liquid (IL) and double-stranded DNA (dsDNA) composite material was used to investigate the direct electron transfer of myglobin (Mb) on carbon ionic liquid electrode (CILE). The presence of the IL–dsDNA biocomposite film on the electrode surface provided great improvement to the direct electron transfer rate of Mb with the CILE, which was due to the synergistic contributions of specific characteristics of dsDNA, IL and their interaction. The electrochemical parameters of Mb in the IL–dsDNA composite film modified electrode were carefully investigated with the charge transfer coefficient (α) and the electron transfer rate constant (ks) calculated as 0.42 and 0.84 s−1, respectively. The fabricated Mb modified electrode exhibited good electrocatalytic ability to the reduction of trichloroacetic acid and H2O2, which showed the potential applications in the third-generation electrochemical biosensor.  相似文献   

17.
Nafion® 211 differs from previous versions of Nafion in that the membrane is cast from a dispersion rather than being melt-extruded. As such, the water sorption properties are different, as is the rate of increase in water content with temperature. Kinetic and mass-transport parameters for dispersion-cast Nafion® 211 were determined using solid-state electrochemistry in the temperature range 30–70 °C, 100% relative humidity, and 30 psi oxygen pressure. Exchange current densities, Tafel slopes, and transfer coefficients for ORR in Nafion® 211, are similar to those observed in Nafion® 117; mass-transport parameters are not. At 30 °C and 100% RH oxygen solubility and the diffusion coefficient is determined to be 1.16 × 10−5 mol cm−3 and 1.13 × 10−6 cm2 s−1, respectively. Oxygen permeability at 30 °C (1.28 × 10−11 mol cm−1 s−1) is lower than in Nafion® 117 (5.31 × 10−11 mol cm−1 s−1) by factor of 4, while at T > 60 °C the permeability of Nafion® 211 increases significantly to values higher than Nafion® 117, and is correlated with the increase in water content and hydration number (λ) with temperature.  相似文献   

18.
The process of Ag+ ions extraction from nitric solutions by bulk liquid membranes containing di(2-ethylhexyl)phosphoric acid and tri-n-octylamine accompanied with silver(I) electrodeposition from perchloric acid solutions is studied at galvanostatic electrodialysis. The effects of the current density as well as of composition of the liquid membrane and aqueous solutions on the rate of the silver(I) transport are determined. More than 90% of silver ions was extracted from the feed solution containing 0.01 mol l−1 AgNO3 and about 40% of metal was electrodeposited during 1 h of electrodialysis. Sound adherent silver coatings have been deposited on platinum and copper cathodes.  相似文献   

19.
A simple, rapid and effective method of microwave assisted chemical bath deposition (MACBD) has been used to deposit CdS quantum dots on the surface of TiO2 film as photoanode of quantum dot-sensitized solar cells. The photovoltaic performance of the as-prepared cell is investigated. The results show that the cell based on MACBD deposited TiO2/CdS electrode achieves an improved short circuit current density of 6.69 mA cm−2 and power conversion efficiency of 1.03% at one sun (AM 1.5G, 100 mW cm−2) as compared with the one employing conventional sequential chemical bath deposition method.  相似文献   

20.
This research in finding a cheap and efficient catalyst for electrooxidation of formaldehyde give us an attempt to make and examine the behavior of poly(N-methylaniline)/nickel modified carbon paste electrode (Ni/P(NMA)/MCPE) in absence and presence of formaldehyde. This involves in situ electropolymerization of N-methylaniline at carbon paste electrode, which is following to the incorporation of Ni(II) to polymeric layer by immersion of modified electrode in 1.0 M nickel sulphate solution. The electrocatalytic oxidation of formaldehyde was studied by cyclic voltammetry and chronoamperometry methods. The effects of scan rate and formaldehyde concentration on the electrocatalytic oxidation of formaldehyde were also investigated at the surface of Ni/P(NMA)/MCPE. The diffusion coefficient (D = 14.1 × 10−5 cm2 s−1), and some kinetic parameters such as the transfer coefficient (α = 0.45) and also second-order rate constant (k = 8.96 × 10−4 cm3 mol−1 s−1) of formaldehyde were calculated.  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