Processing math: 100%
Yue Liu, Shaobo Huang, Shanlong Peng, Heng Zhang, Lifan Wang, and Xindong Wang, Novel Au nanoparticles-inlaid titanium paper for PEM water electrolysis with enhanced interfacial electrical conductivity, Int. J. Miner. Metall. Mater., 29(2022), No. 5, pp.1090-1098. https://dx.doi.org/10.1007/s12613-022-2452-1
Cite this article as: Yue Liu, Shaobo Huang, Shanlong Peng, Heng Zhang, Lifan Wang, and Xindong Wang, Novel Au nanoparticles-inlaid titanium paper for PEM water electrolysis with enhanced interfacial electrical conductivity, Int. J. Miner. Metall. Mater., 29(2022), No. 5, pp.1090-1098. https://dx.doi.org/10.1007/s12613-022-2452-1
Research Article

Novel Au nanoparticles-inlaid titanium paper for PEM water electrolysis with enhanced interfacial electrical conductivity

Author Affilications
  • Corresponding author:

    Xindong Wang      E-mail: echem@ustb.edu.cn

    Lifan Wang      E-mail: wanglifanustb@163.com

  • Proton-exchange membrane water electrolysis (PEM WE) is a particularly promising technology for renewable hydrogen production. However, the excessive passivation of the gas diffusion layer (GDL) will seriously affect the high surface-contact resistance and result in energy losses. Thus, a mechanism for improving the conductivity and interface stability of the GDL is an urgent issue. In this work, we have prepared a hydrophilic and corrosion resistant conductive composite protective coating. The polydopamine (PDA) film on the Ti surface, which was obtained via the solution oxidation method, ensured that neither micropores nor pinholes existed in the final hybrid coatings. In-situ reduced gold nanoparticles (AuNPs) improved the conductivity to achieve the desired interfacial contact resistance and further enhanced the corrosion resistance. The surface composition of the treated samples was investigated using scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray diffraction (XRD), and Fourier transform infrared spectroscopy (FTIR). The results indicated that the optimized reaction conditions included a pH value of 3 of HAuCl4 solution with PDA deposition (48 h) on papers and revealed the lowest contact resistance (0.5 mΩ·cm2) and corrosion resistance (0.001 µA·cm−2) in a 0.5 M H2SO4 + 2 ppm F solution (1.7 V vs. RHE) among all the modified specimens, where RHE represents reversible hydrogen electrode. These findings indicated that the Au–PDA coating is very appropriate for the modification of Ti GDLs in PEM WE systems.
  • Proton-exchange membrane water electrolysis (PEM WE) is a very promising technology for sustainable hydrogen production [12] due to its wide range of current density, flexibility, and high conversion efficiency [35]. As one of the key stack components in PEM systems, the gas diffusion layer (GDL) acts as a functional layer to provide optimized water distribution over the electrode and efficiently remove the produced gas during cycles. To achieve stability, scholars often use various structures based on titanium substrates, such as felts [67] and meshes [8], as GDLs on the anode side of PEM electrolyzers. Typically, the mesh structure performs less well than paper or sintered structures. However, Ti-based components on the anode side show a high potential and are in contact with an acidic (usually Nafion®-type) membrane over time [2], which will cause an increase in the thickness of the passivation layer and further result in a large surface resistance accompanied by a low energy efficiency [9].

    Considering the coating layers, metal coatings are the most direct and efficient coating protection. Based on the type of coating material, three main categories are considered: (i) noble metal, such as Au [1012], Pt [1113], and IrO2 [14]; (ii) conductive metallic oxide (Ti4O7) [1516]; (iii) metal nitrides (TiN) [1721]. The precious metal modification or coating layer has better electrical conductivity and stability than nitrides or oxides. Therefore, a mechanism for enhancing the effectiveness and reducing the consumption of noble metals has become a key technology. As an example, magnetron sputtering [22] has been reported for noble metal ion implantation. In addition, alloying Ti with noble metals can greatly improve the corrosion resistance of Ti [23]. Although the coating preparation by plasma immersion technique has good adhesion to the substrate, the complicated equipment and relatively high cost are not universal. Meanwhile, the easily and commonly generated oxide layers on the surface during the process will greatly increase the resistance after a long period of water electrolysis. Comparatively, conducting polymer coatings not only provide a physical barrier to ion intrusion but also form a passivated film at the coating/metal interface, providing anodic protection [2427]. Zhang and Sharma [28] explored the two-step electrochemical deposition technique for the development of Au nanoparticles (NPs) and a polyaniline hybrid coating on SS316L, which showed a low load of precious metal conductive coating. The adhesion of polymers on the substrate surface is also a key factor affecting the stability of the corrosion interface [2930]. Mussel-inspired polydopamine (PDA) coating is applied on metal surfaces to provide a barrier effect to prevent attacks in a corrosive environment [31]. Catechol is a unique adhesion molecule in mussel proteins, and it reduces metal ions during oxidation and polymerization with amines to form PDA [3238]. The commonly used mussel protein mimic, that is, dopamine (DA) [3940], is an effective reductant in the synthesis of AuNPs [4144]. Besides its reducing capability, DA and the generated PDA can be used as binding agents, which help in stabilizing the produced NPs [45]. However, PDA is rarely used as a reactive template for the growth of metal particles, especially GDLs in PEM WE.

    In this work, AuNPs and PDA hybrid coatings on Ti papers were investigated using a two-step chemical deposition technique for application in GDLs. The PDA coating provided a strong corrosion resistance and shielding function, and the AuNPs improved conductivity to achieve the desired interfacial contact resistance (ICR) and further improve the corrosion resistance. The morphology and content of AuNPs were efficiently regulated by adjusting the aggregation time of PDA and HAuCl4 of different pH. Then, the corrosion behavior was evaluated using electrochemical impedance spectroscopy (EIS), potentiodynamics, and potentiostatic polarization in a simulated PEM WE environment.

    The commercially available titanium alloy paper (Ti–Al, 99.7%, 1.5 cm × 2.5 cm) was used as the base material of GDLs, chloroauric acid (HAuCl4, 99.9%), DA hydrochloride, and Tris(hydroxymethyl)aminomethane were purchased from Shanghai Aladdin Biochemical Technology Co., Ltd., China. DA hydrochloride used in this study were of analytical grade and used without any purification. All other chemicals were of analytical grade and used as received.

    The Ti paper was cut into disks with the size of 1 cm × 1.2 cm by scissors. In a typical procedure, the Ti paper disks were cleaned using consecutive ultrasonication in acetone, ethanol, and deionized water for 15 min and finally dried in an oven. The Ti surface was chemically pretreated by dipping the Ti pieces in a solution containing 0.5 M NaOH and 1.0 M H2O2 for 10 min at room temperature.

    Subsequently, the Ti papers were immersed in a DA solution (2 mg·mL−1 DA in 10 mM Tris, pH = 8.5) for 12, 24, 48, and 72 h, washed with ultrapure water, and dried. The self-polymerization of PDA was visually confirmed by the color change of the surface to light yellow. For the Au coating, the PDA-coated Ti papers were immersed in 50 mL HAuCl4 solutions with pH = 2, 3, and 4 for 30 min and rinsed with deionized water. All experiments were performed at room temperature.

    Scanning electron microscopy (SEM) and energy dispersive spectroscopy (EDS) were conducted on a field-emission scanning electron microscope (FEI Quanta 650FEG). High-resolution transmission electron microscope (HRTEM, JEOL JEM-2100) were used to evaluate morphology and microstructure of as-prepared samples. X-ray diffraction (XRD) patterns were taken on a Burker D8 Advance (Cu Kα radiation, λ = 0.154 nm). X-ray photoelectron spectroscopy (XPS) was performed on an ESCALAB 250Xi, Al Kα. Atomic force microscopy (AFM) measurements were achieved using Cypher VRS (OXFORD) and Au–PDA-coated AFM tips (AC160TS-R3) were used in tapping mode.

    The electrochemical tests (including potentiodynamic polarization, EIS, and potentiostatic polarization test (long-term stability test) were performed using a conventional three-electrode system controlled by an electrochemical workstation VMP2 (Bio-logic, France). A Ti paper electrode was used as the working electrode, and a platinum sheet and saturated calomel electrode were employed as the counter and reference electrodes, respectively. Electrochemical tests were performed in 0.5 M H2SO4 + 2 ppm F solution to simulate the PEM WE working environment. The polarization voltage was 1.70 V vs. reversible hydrogen electrode (RHE). The EIS test was conducted in the frequency range of 100 kHz –10 mHz with a perturbation amplitude of 10 mV under the open circuit potential (OCP) conditions.

    ICR analysis was performed under different cell compressions to investigate the improved electrical and interfacial contact behavior of the coatings. To calculate the through-plane resistance, which is the sum of the GDL resistance, we measured the plate-to-plate voltage drop along with the two contact resistances present between the samples and gold-coated plates. In the through-plane direction, the contact resistance was the major factor affecting the conductivity of GDL.

    PDA has been extensively studied due to its simple preparation, biocompatibility, and capability to self-polymerize with almost any inorganic solid material surface to form strong adhesive films. The amino and catechol groups in the DA structure are the key structures of the mussel foot-filament protein (Fig. 1(a)), and the results showed that DA also has good adhesion properties after polymerization. The autopolymerization of DA under alkaline aerobic conditions (pH = 8.5) is one of the most widely studied methods for the preparation of PDA to date. This method requires only the addition of DA monomer to an alkaline solution, and the DA will spontaneously begin to polymerize using oxygen as the oxidizing agent [4648]. This type of oxidative autopolymerization is the most widely used because it is mild and requires no complex apparatus or demanding reaction conditions. The thickness of films prepared by this method can be adjusted by controlling the concentration of DA and the polymerization time [49]. DA and the generated PDA, besides their reducing ability, can act as binding agents to stabilize the generated AuNPs [3940,45]. In these reports, the authors considered that the initial reducing species is the phenolic or phenolics of DA and that the reduction of Au ions occurs through the two-electron oxidation of DA to generate quinone [43,50]. However, DA also undergoes single-electron oxidation, in which semiquinone radicals (SMQs) are the major intermediates [5152]. Further, the researchers proposed that the synthesis of DA-based AuNPs requires the initial oxidization of DA to form the quinone group and its catalysis to generate SMQs, providing new insights into the formation mechanism of AuNPs using DA as [5355] the reductant. In this context, we have successfully prepared a Au–PDA bilayer composite coating on the surface of a titanium substrate by a two-step method, in which a thin PDA layer was deposited by chemical oxidation on the surface of a chemically treated titanium paper substrate, followed by the in-situ generation of AuNPs on the PDA layer by an in-situ exchange method (Fig. 1(b)).

    Fig. 1.  (a) Mechanism of polymerization of polydopamine (PDA). (b) Schematic of the procedure for the preparation of Au–PDA bilayer composite coating on Ti papers.

    Considering the deposition conditions of PDA layers for the preparation of Au–PDA bilayers with good adhesion, we first investigated the relationship between Au–PDA bilayers and substrate adhesion at different deposition times. The experiments were set at different deposition times of 12, 24, 48, and 72 h, and the Ti papers had a dense PDA coating on the surface (Fig. S1(a)–(e)). With the increase in deposition time, the formation of PDA microspheres on the Ti paper surface gradually increased, forming a thick film (Fig. S1(f)). Fig. S2 shows the mapping at the deposition time of 48 h and homogeneous PDA microspheres on the surface of the titanium mat with deposition. Subsequently, Ti papers with different deposition times were immersed in chloroauric acid solution (pH = 2) (Fig. 2(a)–(d)). Based on the morphology of AuNPs, at the deposition time of 12 h (Fig. 2(a)), the AuNPs were nearly 20 nm in size. At the deposition time of 24 h, given the uneven local response, some gold particles exhibited uneven sizes, and their diameters were about 50–200 nm (Fig. 2(b)). At the deposition time of 48 h, the gold particles were evenly distributed, and the particle size was almost consistent (100–200 nm) (Fig. 2(c)). At the deposition time of 72 h, the gold particles (20–200 nm) exhibited local agglomeration and sparse dispersion (Fig. 2(d)). To evaluate the adhesion of PDA with AuNPs at different deposition times in 0.5 M H2SO4 + 2 ppm F−1 solution, we obtained the morphological results acquired through potentiostatic polarization tests (Fig. 2(e)–(h)). After a 20 h polarization test, the AuNPs on the surface of the Ti paper with a PDA deposition time of 12 h were exfoliated, and a small number of gold particles remained (Fig. 2(e)). At the deposition time of 24 h, pitting occurred on the surface, indicating that the PDA film had been partially peeled off, which was detrimental to the protection of the Ti papers. However, for the sample soaked in Ti paper for 48 h, small-sized Au particles were still evenly distributed on the surface. For the same sample at 72 h, the agglomerated AuNPs were dislodged, and only nucleated gold particles existed stably on the surface. After the comparison of morphological and electrochemical test results, the optimal deposition time was 48 h. Meanwhile, the pristine PDA samples were evaluated for corrosion performance, yielding an optimum performance of 48 h. The corrosion current density (Icorr) of 48 h Au–PDA@Ti was 0.015 µA·cm−2 (Fig. S3). Fig. 2(i)–(l) show the EDS mapping of Fig. 2(i) inset for the Au–PDA@Ti (48 h) paper; the AuNPs were uniformly distributed on the skeleton of the Ti papers, which can reduce the ICR of the Ti paper.

    Fig. 2.  (a–d) SEM of different morphologies of AuNPs obtained by regulating the PDA aggregation time. (e–h) SEM images of the Ti papers prepared with different PDA aggregation times after potentiostatic polarization for 20 h: (e) 12 h, (f) 24h, (g) 48 h, and (h) 72 h. (i–l) EDS mapping spectra of Ti, Al, C, and Au of the scanning area (insert in (i)) in Au–PDA@Ti (48 h).

    A more detailed exploration of reaction concentration was conducted to explore the morphological evolution of AuNPs at the PDA polymerization time of 48 h for the Ti papers. The reduction reaction occurred very rapidly, and the high concentration of HAuCl4 was promptly reduced on the PDA@Ti paper. For effective comparison, the lower reaction concentration must be controlled to obtain a suitable concentration (Fig. 3). The PDA@Ti papers were immersed in different solution concentrations of HAuCl4 (pH = 2, 3, and 4), and the reaction time was 0.5 h. At a high concentration (pH = 2), the AuNPs agglomerated (Fig. 3(a)); the AuNPs can be evenly and densely tiled on the surface of the PDA@Ti at pH value of 3 (Fig. 3(b)); a small amount of AuNPs can be generated and were sparsely dispersed on the surface at pH value of 4 (Fig. 3(c)), and the optimum reaction concentration was pH = 3 (Fig. 3(b)). Fig. 3(d)–(i) show the low-resolution SEM of different concentrations in Fig. 3(a)–(c). The optimized reaction condition can be observed at pH value of 3 (loaded with 4.5 mg gold nanoparticles).

    Fig. 3.  Different resolution SEM images of the AuNPs by regulating pH values: (a, d, g) pH = 2; (b, e, h) pH = 3; (c, f, i) pH = 4.

    Fig. 4 shows the contact angles of the Au–PDA@Ti (48 h, pH = 3) to show its hydrophilicity. Once the paper surface is infiltrated, the capillary action “sucks” water droplets into the paper. The instantaneous morphology of the water droplets during the measurement of contact angle within 1 s can be observed when the water drops leaving the capillary tube formed half balls on the surface of the paper. The contact surface between the water droplets and the paper was rapidly reduced as soon as the surface came into contact with the water droplets flowing from the end of the capillary (Fig. 4(a)–(b)), permeating into the paper (Fig. 4(c)–(f)). In addition, pure Ti was tested for transient contact angle (Fig. S4), and the transient change in contact angle was small (47.8°), indicating that the modified Ti (29.5°) had excellent hydrophilicity. A low water contact angle of the GDLs indicates a hydrophilic character, and it is required for the effective circulation of water for transportation during PEM WE operation. The water contact angles of the Ti paper decreased remarkably when the GDLs were coated with Au–PDA.

    Fig. 4.  Instantaneous morphologies of the water and the instant contact angles during the water dropped to the Au–PDA@Ti (48 h, pH = 3) paper within 1 s.

    Thus, the microscopic structure of the Au–PDA nanocomposites, Au–PDA@Ti (48 h, pH = 3), analyzed by TEM confirmed the incorporation of AuNPs inlaid in the PDA films over the Ti papers (Fig. 5(a)−(c)). The images revealed particle clusters with an amorphous coating of PDA (Fig. 5(a)), whereas the AuNPs were sphere-shaped with sizes ranging from 80 to 150 nm. We observed the reduction of HAuCl4 to Au in the PDA particles, tightly integrating with the PDA shown in Fig. 5(b). The TEM image of the AuNPs (Fig. 5(c)) showed that the interfringe distance of AuNPs measured 0.209 and 0.083 nm, which can be assigned to the (200) and (422) crystal planes of fcc Au, respectively. In addition, C, N, and Au were observed on the surface of Au–PDA. Combining EDS (Fig. 5(d)–(f)) with XRD (Fig. 5(g)), XPS (Fig. 5(h)), and FTIR (Fig. 5(i)) curves, AuNPs have been successfully inlaid on the surface of PDA owing to the in-situ reduction effect of PDA. Fig. 5(g) shows the XRD spectrum of Ti paper and Au–PDA. The characteristic peaks of Au are at = 38.12°, 44.39°, and 64.58°, which correspond to the Au (111), (200), and (220) planes, respectively. The chemical structures of PDA and Au–PDA were characterized by FTIR (Fig. 5(i)). Then, the AuNPs species were analyzed by XPS (Fig. 5(h)). In the Au 4f XPS spectrum, two pairs of characteristic peaks corresponding to Au 4f7/2 and 4p5/2 of Au0 were observed at the binding energies of 84.1 and 87.88 eV, respectively. The chemical structures of PDA and Au–PDA were characterized by FTIR. From Fig. 5(i), the PDA spectrum and Au–PDA showed several characteristic peaks at 3228, 2947, and 1499 cm−1, which were attributed to the −OH, −CH2, and –NH groups, respectively [47]. The former can be attributed to the absorption of the aromatic ring on PDA and the latter to the bending vibration of the N–H bond on PDA.

    Fig. 5.  TEM images of (a) Au–PDA and (b) AuNPs, i.e., the red rectangular area in (a); (c) high-resolution TEM (HRTEM) image of enclosed AuNPs. (d–f) Corresponding EDS mapping images of C, N, and Au in partial view of (a). (g) XRD patterns of Ti and Au–PDA. (h) XPS spectra (Au 4f spectra) of Au–PDA@Ti (48 h, pH = 3). (i) FTIR spectra of PDA and Au–PDA.

    The corrosion resistance of pristine Ti paper and Au–PDA@Ti (48 h) papers at different pH values were evaluated in the simulated PEM WE in Fig. 6(a)–(f). Fig. 6(a) shows the potentiodynamic polarization curves of the pristine Ti and different concentrations of Au–PDA@Ti (48 h) specimens in 0.5 M H2SO4 + 2 ppm F solution (1.7 V vs. RHE). In the potentiodynamic polarization curves, the Au–PDA@Ti (48 h) at pH = 3 exhibited an active dissolution at about 1.2 V. With the increase in the anodic potential, a dimensional passivation zone appeared, indicating the formation of a stable passivation film. At the potential of 1.4–1.5 V, an over-passivation zone appeared, which may indicate the partial shedding of AuNPs. The potential reached 1.7 V, and the dimensional passivation zone appeared again. Thus, the PDA film played a key role in inhibiting the corrosion dissolution of Ti. The other samples showed only passivated areas, indicating the stability of the passivated layer. Table 1 presents the Tafel parameters, such as the corrosion potential (Ecorr) and corrosion current density (Icorr), compared with the pure Ti paper. The Au–PDA@Ti (48 h) at pH = 3 exhibited a higher Ecorr (801.02 mV vs. RHE) and the Icorr (0.001 µA·cm−2), which is about 1% of the department of energy (DOE) target (below 1 µA·cm−2) and close to the value of TiN coating (0.009 µA·cm−2) prepared by Bi et al. [18]. The Au–PDA@Ti (48 h) with pH = 3 had a low corrosion current, indicating a high corrosion resistance. The EIS test was carried out to evaluate the corrosion resistance to determine the stability of the passive film on the surface of the Ti papers. Fig. 6(b) and (c) display the Nyquist and Bode plots of pure Ti at different concentrations (pH = 2, 3, and 4) of Au–PDA@Ti (48 h), respectively. Fig. 6(b) shows the equivalent circuits adopted to fit the EIS data of pure Ti and Au–PDA@Ti (48 h) samples. The insert of Fig. 6(b) depicts the equivalent circuits used to fit EIS data. Rs is for solution resistance (the resistance between the working and reference electrodes), CPE stands for double-layer capacitance, and Rct is the charge transfer resistance between the coating interface and the substrate.The capacitive arc diameter of the sample treated by Au–PDA enlarged, indicating that the corrosion resistance of the Au–PDA@Ti sample increased. This finding was due to the less number of defects on the surface, which can be due to the existence of the stable film that resisted the inward penetration of corrosive ions, such as F and SO24 [56]. In the phase-angle plot (Fig. 6(b)), the high-frequency plots of untreated and Au–PDA@Ti together resulted from the local surface defects on the surface in low frequency, and the phase-angle was relatively higher for Au–PDA@Ti compared with the untreated one. Thus, corrosion occurred on the surface of pure Ti. The Au–PDA@Ti (48 h) at pH = 2, 3, and 4 exhibited identical resistances throughout the entire compression range, whereas the pristine Ti paper showed a significantly higher resistance in the overall compression ranges in Fig. 6(d). Fig. 6(e) compares the ICRs of the different concentrations Au–PDA@Ti before and after potentiostatic polarizations for 10 h at 140 N·cm−2. The pristine Ti paper had a resistance of 3.20 mΩ·cm2 at 140 N·cm−2, and the lowest contact resistance was 0.5 mΩ·cm2 at pH = 3. After potentiostatic polarization for 10 h, the different concentrations of Au–PDA@Ti (48 h) showed no significant increase in their contact resistance. However, the pure Ti paper exhibited a dramatic increase in the contact resistance of 5.50 mΩ·cm2 at 140 N·cm−2. The potentiostatic test was conducted to examine the current density of the Ti paper for a long-term operation in the anode environment. Fig. 6(f) shows the potentiostatic curves for pure Ti paper and Au–PDA@Ti (48 h) papers. The reduction in the current density reflected the formation of the passive film on the Ti paper. The Au–PDA@Ti (48 h) had a lower corrosion current than the pure one in the anode PEM WE environment. The corrosion resistance of Ti paper was improved significantly by Au–PDA@Ti (48 h), especially the pH = 3 sample. Moreover, the pure Ti paper exhibited a more turbulent current fluctuation than the Au–PDA@Ti (48 h). In Fig. S5, after polarization for 10 h (1.7 V vs. RHE), the PDA on the surface of Au–PDA@Ti (48 h) with pH = 2 fractured, and the AuNPs were unevenly distributed, resulting in a local increase in the current density and cracking of the Ti paper. In addition, the Au–PDA@Ti (48 h) at pH = 3 revealed a uniform dispersion of Au particles on the surface. Furthermore, this finding was verified by the AFM results (Fig. S6), indicating that the conductivity was significantly higher in Au–PDA@Ti (48 h) polarized for 10 h than in pure Ti. The Au–PDA@Ti (48 h) at pH = 3 revealed high conductivity and corrosion resistance in stabilizing the cycling stability of PEM WE.

    Fig. 6.  (a) Potentiodynamic polarization curves of pure Ti and Au–PDA@Ti (48 h) papers in the PEM WE environment. (b) Nyquist and (c) Bode phase-angle (θ) plots of pure Ti and Au–PDA@Ti (48 h) papers in PEM WE environment; insert in (b) is the equivalent circuits used to fit EIS data. (d) Contact resistance of pure Ti and Au–PDA@Ti (48 h) papers using a conventional press setup, in which all the papers were pressed between two gold-coated copper plates. (e) ICR of pure Ti and Au–PDA@Ti (48 h) papers at 140 N·cm−2 before and after potentiostatic polarization for 10 h. (f) Potentiostatic polarization curves of pure Ti and Au–PDA@Ti (48 h) papers in the PEM WE environment.
    Table  1.  Corrosion resistance and conductivity of pure Ti paper and Au–PDA@Ti papers
    SpecimenCorrosion resistanceConductivity (ICR at 140 N·cm−2) / (mΩ·cm2)
    Ecorr / mV vs. RHEIcorr / (μA·cm−2)Rct / (Ω·cm2)
    Pure Ti534.670.7212.330 × 1033.2
    48 h, pH = 2734.540.0466.604 × 1030.9
    48 h, pH = 3801.020.0011.596 × 1040.5
    48 h, pH = 4714.180.2732.835 × 1031.4
     | Show Table
    DownLoad: CSV

    This study presented unique approaches to improving the corrosion resistance and conductivity of the GDL used in water-splitting devices. The Au–PDA coating was prepared efficiently by depositing trace amounts of AuNPs on the surface of PDA-coated Ti paper by employing a two-step chemical deposition process in a HAuCl4 solution with an extremely low concentration. The optimized reaction conditions included treatment with pH 3 HAuCl4 for 0.5 h, and the PDA aggregation time was 48 h. The Au–PDA@Ti paper showed a very low ICR (0.5 mΩ·cm2 at 140 N·cm−2) and excellent corrosion resistance in simulated PEM WE environments (1.7 V vs. RHE) with an Ecorr of 0.801 V vs. RHE. The potentiodynamic polarization Icorr of Au–PDA@Ti (48 h, pH = 3) was reduced from 0.015 and 0.721 μA·cm−2 for pure Ti paper and Au–PDA samples to 0.001 μA·cm−2, respectively. The SEM–EDS and HRTEM study indicated that the average particle size was about 80 nm, with agglomerations reaching up to 200 nm. This finding demonstrated that Au–PDA@Ti is a promising GDL material in the PEM WE environment.

    This work was financially supported by the National Key Research and Development Program of China (No. 2018YFB1502403).

    The authors declare no potential conflict of interest.

    The online version contains supplementary material available at https://doi.org/10.1007/s12613-022-2452-1

  • [1]
    C. Liu, M. Shviro, A.S. Gago, et al., Exploring the interface of skin-layered titanium fibers for electrochemical water splitting, Adv. Energy Mater., 11(2021), No. 8, art. No. 2002926. DOI: 10.1002/aenm.202002926
    [2]
    M. Carmo, D.L. Fritz, J. Mergel, and D. Stolten, A comprehensive review on PEM water electrolysis, Int. J. Hydrogen Energy, 38(2013), No. 12, p. 4901. DOI: 10.1016/j.ijhydene.2013.01.151
    [3]
    C. Niether, S. Faure, A. Bordet, et al., Improved water electrolysis using magnetic heating of FeC–Ni core-shell nanoparticles, Nat. Energy, 3(2018), No. 6, p. 476. DOI: 10.1038/s41560-018-0132-1
    [4]
    A. Landman, H. Dotan, G.E. Shter, et al., Photoelectrochemical water splitting in separate oxygen and hydrogen cells, Nat. Mater., 16(2017), No. 6, p. 646. DOI: 10.1038/nmat4876
    [5]
    F. Barbir, PEM electrolysis for production of hydrogen from renewable energy sources, Sol. Energy, 78(2005), No. 5, p. 661. DOI: 10.1016/j.solener.2004.09.003
    [6]
    H. Ito, T. Maeda, A. Nakano, A. Kato, and T. Yoshida, Influence of pore structural properties of current collectors on the performance of proton exchange membrane electrolyzer, Electrochim. Acta, 100(2013), p. 242. DOI: 10.1016/j.electacta.2012.05.068
    [7]
    F. Arbabi, A. Kalantarian, R. Abouatallah, R. Wang, J.S. Wallace, and A. Bazylak, Feasibility study of using microfluidic platforms for visualizing bubble flows in electrolyzer gas diffusion layers, J. Power Sources, 258(2014), p. 142. DOI: 10.1016/j.jpowsour.2014.02.042
    [8]
    S. Siracusano, A. Di Blasi, V. Baglio, et al., Optimization of components and assembling in a PEM electrolyzer stack, Int. J. Hydrogen Energy, 36(2011), No. 5, p. 3333. DOI: 10.1016/j.ijhydene.2010.12.044
    [9]
    N. Sato, An overview on the passivity of metals, Corros. Sci., 31(1990), p. 1. DOI: 10.1016/0010-938X(90)90086-K
    [10]
    H.Y. Jung, S.Y. Huang, P. Ganesan, and B.N. Popov, Performance of gold-coated titanium bipolar plates in unitized regenerative fuel cell operation, J. Power Sources, 194(2009), No. 2, p. 972. DOI: 10.1016/j.jpowsour.2009.06.030
    [11]
    S.H. Wang, J. Peng, and W.B. Lui, Surface modification and development of titanium bipolar plates for PEM fuel cells, J. Power Sources, 160(2006), No. 1, p. 485. DOI: 10.1016/j.jpowsour.2006.01.020
    [12]
    M.J. Hwang, E.J. Park, W.J. Moon, H.J. Song, and Y.J. Park, Characterization of passive layers formed on Ti–10wt% (Ag, Au, Pd, or Pt) binary alloys and their effects on galvanic corrosion, Corros. Sci., 96(2015), p. 152. DOI: 10.1016/j.corsci.2015.04.007
    [13]
    Z.X. He, Y.R. Lv, T.A. Zhang, et al., Electrode materials for vanadium redox flow batteries: Intrinsic treatment and introducing catalyst, Chem. Eng. J., 427(2022), art. No. 131680. DOI: 10.1016/j.cej.2021.131680
    [14]
    S.H. Wang, W.B. Lui, J. Peng, and J.S. Zhang, Performance of the iridium oxide (IrO2)-modified titanium bipolar plates for the light weight proton exchange membrane fuel cells, J. Fuel Cell Sci. Technol., 10(2013), No. 4, art. No. 041002. DOI: 10.1115/1.4024565
    [15]
    H. Wakayama and K. Yamazaki, Low-cost bipolar plates of Ti4O7-coated Ti for water electrolysis with polymer electrolyte membranes, ACS Omega, 6(2021), No. 6, p. 4161. DOI: 10.1021/acsomega.0c04786
    [16]
    Y.Z. Chen, D.J. Jiang, Z.Q. Gong, J.Y. Li, and L.N. Wang, Anodized metal oxide nanostructures for photoelectrochemical water splitting, Int. J. Miner. Metall. Mater., 27(2020), No. 5, p. 584. DOI: 10.1007/s12613-020-1983-6
    [17]
    T.J. Toops, M.P. Brady, F.Y. Zhang, et al., Evaluation of nitrided titanium separator plates for proton exchange membrane electrolyzer cells, J. Power Sources, 272(2014), p. 954. DOI: 10.1016/j.jpowsour.2014.09.016
    [18]
    J. Bi, J.M. Yang, X.X. Liu, et al., Development and evaluation of nitride coated titanium bipolar plates for PEM fuel cells, Int. J. Hydrogen Energy, 46(2021), No. 1, p. 1144. DOI: 10.1016/j.ijhydene.2020.09.217
    [19]
    K. Feng, D.T.K. Kwok, D.A. Liu, Z.G. Li, X. Cai, and P.K. Chu, Nitrogen plasma-implanted titanium as bipolar plates in polymer electrolyte membrane fuel cells, J. Power Sources, 195(2010), No. 19, p. 6798. DOI: 10.1016/j.jpowsour.2010.04.053
    [20]
    A. Shenhar, I. Gotman, E.Y. Gutmanas, and P. Ducheyne, Surface modification of titanium alloy orthopaedic implants via novel powder immersion reaction assisted coating nitriding method, Mater. Sci. Eng. A, 268(1999), No. 1-2, p. 40. DOI: 10.1016/S0921-5093(99)00111-2
    [21]
    X. Zhang, W.W. Yang, M.Y. Gao, H. Liu, K.F. Li, and Y.S. Yu, Room-temperature solid phase surface engineering of BiOI sheets stacking g-C3N4 boosts photocatalytic reduction of Cr(VI), Green Energy Environ., 7(2022), No. 1, p. 66. DOI: 10.1016/j.gee.2020.07.024
    [22]
    E. Petkucheva, G. Borisov, E. Lefterova, J. Heiss, U. Schnakenberg, and E. Slavcheva, Gold-supported magnetron sputtered Ir thin films as OER catalysts for cost-efficient water electrolysis, Int. J. Hydrogen Energy, 43(2018), No. 35, p. 16905. DOI: 10.1016/j.ijhydene.2018.01.188
    [23]
    M. Stern and H. Wissenberg, The influence of noble metal alloy additions on the electrochemical and corrosion behavior of titanium, J. Electrochem. Soc., 106(1959), No. 9, art. No. 759. DOI: 10.1149/1.2427493
    [24]
    L. Jiang, J.A. Syed, Y.Z. Gao, H.B. Lu, and X.K. Meng, Electrodeposition of Ni(OH)2 reinforced polyaniline coating for corrosion protection of 304 stainless steel, Appl. Surf. Sci., 440(2018), p. 1011. DOI: 10.1016/j.apsusc.2018.01.145
    [25]
    Y. Wang and D.O. Northwood, An investigation into the effects of a nano-thick gold interlayer on polypyrrole coatings on 316L stainless steel for the bipolar plates of PEM fuel cells, J. Power Sources, 175(2008), No. 1, p. 40. DOI: 10.1016/j.jpowsour.2007.09.089
    [26]
    C. Xia, Y. Li, Y. Tian, et al., Intermediate temperature fuel cell with a doped ceria-carbonate composite electrolyte, J. Power Sources, 195(2010), No. 10, p. 3149. DOI: 10.1016/j.jpowsour.2009.11.104
    [27]
    L. Ai, Y. Liu, X.Y. Zhang, X.H. Ouyang, and Z.Y. Ge, A facile and template-free method for preparation of polythiophene microspheres and their dispersion for waterborne corrosion protection coatings, Synth. Met., 191(2014), p. 41. DOI: 10.1016/j.synthmet.2014.02.004
    [28]
    K. Zhang and S. Sharma, Site-selective, low-loading, Au nanoparticle-polyaniline hybrid coatings with enhanced corrosion resistance and conductivity for fuel cells, ACS Sustain. Chem. Eng., 5(2017), No. 1, p. 277. DOI: 10.1021/acssuschemeng.6b01504
    [29]
    A. Jacques, B. Barthélémy, J. Delhalle, and Z. Mekhalif, 1-Pyrrolyl-10-decylammoniumphosphonate monolayer: A molecular nanolink between electropolymerized pyrrole films and nickel or titanium surfaces, Electrochim. Acta, 170(2015), p. 218. DOI: 10.1016/j.electacta.2015.04.123
    [30]
    M. Rohwerder and A. Michalik, Conducting polymers for corrosion protection: What makes the difference between failure and success? Electrochim. Acta, 53(2007), No. 3, p. 1300. DOI: 10.1016/j.electacta.2007.05.026
    [31]
    J.L. Tan, Z. Zhang, and D.T. Ge, Electrodeposition of adherent polypyrrole film on titanium surface with enhanced anti-corrosion performance, MATEC Web Conf., 130(2017), art. No. 08007. DOI: 10.1051/matecconf/201713008007
    [32]
    V. Ball, Polydopamine films and particles with catalytic activity, Catal. Today, 301(2018), p. 196. DOI: 10.1016/j.cattod.2017.01.031
    [33]
    Y. Liang, J. Wei, Y.X. Hu, et al., Metal-polydopamine frameworks and their transformation to hollow metal/N-doped carbon particles, Nanoscale, 9(2017), No. 16, p. 5323. DOI: 10.1039/C7NR00978J
    [34]
    T.L. Chang, X.J. Yu, and J.F. Liang, Polydopamine-enabled surface coating with nano-metals, Surf. Coat. Technol., 337(2018), p. 389. DOI: 10.1016/j.surfcoat.2018.01.009
    [35]
    W. Tamakloe, D.A. Agyeman, M. Park, J. Yang, and Y.M. Kang, Polydopamine-induced surface functionalization of carbon nanofibers for Pd deposition enabling enhanced catalytic activity for the oxygen reduction and evolution reactions, J. Mater. Chem. A, 7(2019), No. 13, p. 7396. DOI: 10.1039/C9TA00025A
    [36]
    X.H. Guo, M. Zhang, J. Zheng, et al., Fabrication of Co@SiO2@C/Ni submicrorattles as highly efficient catalysts for 4-nitrophenol reduction, Dalton Trans., 46(2017), No. 35, p. 11598. DOI: 10.1039/C7DT02095C
    [37]
    C.H. Liu, Y.Y. Qiu, Y.J. Xia, et al., Noble-metal-free tungsten oxide/carbon (WOx/C) hybrid manowires for highly efficient hydrogen evolution, Nanotechnology, 28(2017), No. 44, art. No. 445403. DOI: 10.1088/1361-6528/aa8613
    [38]
    K.M. Im, T.W. Kim, and J.R. Jeon, Metal-chelation-assisted deposition of polydopamine on human hair: A ready-to-use eumelanin-based hair dyeing methodology, ACS Biomater. Sci. Eng., 3(2017), No. 4, p. 628. DOI: 10.1021/acsbiomaterials.7b00031
    [39]
    H. Lee, S.M. Dellatore, W.M. Miller, and P.B. Messersmith, Mussel-inspired surface chemistry for multifunctional coatings, Science, 318(2007), No. 5849, p. 426. DOI: 10.1126/science.1147241
    [40]
    H.Q. Li, Y.V. Aulin, L. Frazer, et al., Structure evolution and thermoelectric properties of carbonized polydopamine thin films, ACS Appl. Mater. Interfaces, 9(2017), No. 8, p. 6655. DOI: 10.1021/acsami.6b15601
    [41]
    J.A.A. Ho, H.C. Chang, and W.T. Su, DOPA-mediated reduction allows the facile synthesis of fluorescent gold nanoclusters for use as sensing probes for ferric ions, Anal. Chem., 84(2012), No. 7, p. 3246. DOI: 10.1021/ac203362g
    [42]
    C.C. Lu, M. Zhang, A.J. Li, X.W. He, and X.B. Yin, 3, 4-dihydroxy-L-phenylalanine for preparation of gold nanoparticles and as electron transfer promoter in H2O2 biosensor, Electroanalysis, 23(2011), No. 10, p. 2421. DOI: 10.1002/elan.201100291
    [43]
    P.C. Huang, W.J. Ma, P. Yu, and L.Q. Mao, Dopamine-directed in situ and one-step synthesis of Au@Ag core-shell nanoparticles immobilized to a metal-organic framework for synergistic catalysis, Chem. Asian J., 11(2016), No. 19, p. 2705. DOI: 10.1002/asia.201600469
    [44]
    Y.Z. Ni, G.S. Tong, J. Wang, et al., One-pot preparation of pomegranate-like polydopamine stabilized small gold nanoparticles with superior stability for recyclable nanocatalysts, RSC Adv., 6(2016), No. 47, p. 40698. DOI: 10.1039/C6RA05902C
    [45]
    G.X. Su, C. Yang, and J.J. Zhu, Fabrication of gold nanorods with tunable longitudinal surface plasmon resonance peaks by reductive dopamine, Langmuir, 31(2015), No. 2, p. 817. DOI: 10.1021/la504041f
    [46]
    V. Ball, D.D. Frari, V. Toniazzo, and D. Ruch, Kinetics of polydopamine film deposition as a function of pH and dopamine concentration: Insights in the polydopamine deposition mechanism, J. Colloid Interface Sci., 386(2012), No. 1, p. 366. DOI: 10.1016/j.jcis.2012.07.030
    [47]
    J.H. Jiang, L.P. Zhu, L.J. Zhu, B.K. Zhu, and Y.Y. Xu, Surface characteristics of a self-polymerized dopamine coating deposited on hydrophobic polymer films, Langmuir, 27(2011), No. 23, p. 14180. DOI: 10.1021/la202877k
    [48]
    W. Zhang, F.K. Yang, Y.G. Han, R. Gaikwad, Z. Leonenko, and B.X. Zhao, Surface and tribological behaviors of the bioinspired polydopamine thin films under dry and wet conditions, Biomacromolecules, 14(2013), No. 2, p. 394. DOI: 10.1021/bm3015768
    [49]
    F. Bernsmann, V. Ball, F. Addiego, et al., Dopamine–melanin film deposition depends on the used oxidant and buffer solution, Langmuir, 27(2011), No. 6, p. 2819. DOI: 10.1021/la104981s
    [50]
    Y.H. Lee and T.G. Park, Facile fabrication of branched gold nanoparticles by reductive hydroxyphenol derivatives, Langmuir, 27(2011), No. 6, p. 2965. DOI: 10.1021/la1044078
    [51]
    M. Bisaglia, S. Mammi, and L. Bubacco, Kinetic and structural analysis of the early oxidation products of dopamine: Analysis of the interactions with α-synuclein, J. Biol. Chem., 282(2007), No. 21, p. 15597. DOI: 10.1074/jbc.M610893200
    [52]
    I. Iftikhar, K.M.A. El-Nour, and A. Brajter-Toth, Detection of transient dopamine antioxidant radicals using electrochemistry in electrospray ionization mass spectrometry, Electrochim. Acta, 249(2017), p. 145. DOI: 10.1016/j.electacta.2017.07.087
    [53]
    S.N. Du, Y. Luo, Z.F. Liao, et al., New insights into the formation mechanism of gold nanoparticles using dopamine as a reducing agent, J. Colloid Interface Sci., 523(2018), p. 27. DOI: 10.1016/j.jcis.2018.03.077
    [54]
    O. Terland, T. Flatmark, A. Tangerås, and M. Grønberg, Dopamine oxidation generates an oxidative stress mediated by dopamine semiquinone and unrelated to reactive oxygen species, J. Mol. Cell. Cardiol., 29(1997), No. 6, p. 1731. DOI: 10.1006/jmcc.1997.0412
    [55]
    A. Klegeris, L.G. Korkina, and S.A. Greenfield, Autoxidation of dopamine: A comparison of luminescent and spectrophotometric detection in basic solutions, Free. Radic. Biol. Med., 18(1995), No. 2, p. 215. DOI: 10.1016/0891-5849(94)00141-6
    [56]
    S.P. Mani, C. Anandan, and N. Rajendran, Formation of a protective nitride layer by electrochemical nitridation on 316L SS bipolar plates for a proton exchange membrane fuel cell (PEMFC), RSC Adv., 5(2015), No. 79, p. 64466. DOI: 10.1039/C5RA05412E
  • Related Articles

    [1]Adnan I. Khdair, A. Ibrahim. Effect of graphene addition on the physicomechanical and tribological properties of Cu nanocomposites [J]. International Journal of Minerals, Metallurgy and Materials, 2022, 29(1): 161-167. DOI: 10.1007/s12613-020-2183-0
    [2]Behrouz Bagheri, Mahmoud Abbasi, Amin Abdollahzadeh, Amir Hossein Kokabi. A comparative study between friction stir processing and friction stir vibration processing to develop magnesium surface nanocomposites [J]. International Journal of Minerals, Metallurgy and Materials, 2020, 27(8): 1133-1146. DOI: 10.1007/s12613-020-1993-4
    [3]Gholam Hossein Majzoobi, Kaveh Rahmani. Mechanical characterization of Mg−B4C nanocomposite fabricated at different strain rates [J]. International Journal of Minerals, Metallurgy and Materials, 2020, 27(2): 252-263. DOI: 10.1007/s12613-019-1902-x
    [4]Masoud Sabzi, Saeid Mersagh Dezfuli, Zohre Balak. Crystalline texture evolution, control of the tribocorrosion behavior, and significant enhancement of the abrasion properties of a Ni-P nanocomposite coating enhanced by zirconia nanoparticles [J]. International Journal of Minerals, Metallurgy and Materials, 2019, 26(8): 1020-1030. DOI: 10.1007/s12613-019-1805-x
    [5]Sajjad Sattari, Amir Atrian. Effects of the deep rolling process on the surface roughness and properties of an Al-3vol%SiC nanoparticle nanocomposite fabricated by mechanical milling and hot extrusion [J]. International Journal of Minerals, Metallurgy and Materials, 2017, 24(7): 814-825. DOI: 10.1007/s12613-017-1465-7
    [6]Ekarat Detsri, Sirilak Rujipornsakul, Tanapong Treetasayoot, Pawarit Siriwattanamethanon. Nanostructured multilayer thin films of multiwalled carbon nanotubes/gold nanoparticles/glutathione for the electrochemical detection of dopamine [J]. International Journal of Minerals, Metallurgy and Materials, 2016, 23(10): 1204-1214. DOI: 10.1007/s12613-016-1340-y
    [7]Le Zhang, Hui-ping Shao, Hang Zheng, Tao Lin, Zhi-meng Guo. Synthesis and characterization of Fe3O4@SiO2 magnetic composite nanoparticles by a one-pot process [J]. International Journal of Minerals, Metallurgy and Materials, 2016, 23(9): 1112-1118. DOI: 10.1007/s12613-016-1329-6
    [8]Li-ping Wang, Xu-dong Xu, Jia-xing Wang. Preparation and characterization of oil-soluble In2O3 nanoparticles and In2O3-SnO2 nanocomposites and their calcined thin films [J]. International Journal of Minerals, Metallurgy and Materials, 2015, 22(5): 543-548. DOI: 10.1007/s12613-015-1105-z
    [9]Xiu-yu Liu, Cong-ying Cui, Ying-wen Cheng, Hou-yi Ma, Duo Liu. Shape control technology during electrochemical synthesis of gold nanoparticles [J]. International Journal of Minerals, Metallurgy and Materials, 2013, 20(5): 486-492. DOI: 10.1007/s12613-013-0755-y
    [10]Qing Wu, Hong Qiu, Chun-ying Fang, Kun Fang. Effect of vacuum annealing on the characteristics of composite pellets containing DBSA-doped polyaniline and Fe nanoparticles [J]. International Journal of Minerals, Metallurgy and Materials, 2009, 16(2): 236-241. DOI: 10.1016/S1674-4799(09)60040-6
  • Cited by

    Periodical cited type(12)

    1. Tiantian Wang, Song Peng, Li Xu, et al. TiCN-Coated Ti Felt for Enhanced Polymer Electrolyte Membrane Water Electrolysis. Industrial & Engineering Chemistry Research, 2025, 64(1): 421. DOI:10.1021/acs.iecr.4c03205
    2. Ritesh K. Upadhyay, Ashis K. Chakraborty, Shyam S. Majhi, et al. Influences of redox electrolyte containing AuNPs on microscopic surface structure, material removal and surface roughness of 20MnCr5 steel alloy during electrochemical machining. Surface Science and Technology, 2025, 3(1) DOI:10.1007/s44251-025-00074-9
    3. Tianjing Li, Hainan Sun, Zhenhua Dan, et al. Recent progress on transition metal-based amorphous ribbons as electrocatalysts for water splitting. International Journal of Minerals, Metallurgy and Materials, 2025, 32(4): 757. DOI:10.1007/s12613-024-3015-4
    4. Hainan Sun, Xiaomin Xu, Gao Chen, et al. Perovskite oxides as electrocatalysts for water electrolysis: From crystalline to amorphous. Carbon Energy, 2024, 6(11) DOI:10.1002/cey2.595
    5. Jiuhong Zhang, Xiejing Luo, Yingyu Ding, et al. Effect of bipolar-plates design on corrosion, mass and heat transfer in proton-exchange membrane fuel cells and water electrolyzers: A review. International Journal of Minerals, Metallurgy and Materials, 2024, 31(7): 1599. DOI:10.1007/s12613-023-2803-6
    6. Xia Li, Yuchen Yao, Yunrui Tian, et al. Recent advances in key components of proton exchange membrane water electrolysers. Materials Chemistry Frontiers, 2024, 8(13): 2493. DOI:10.1039/D4QM00086B
    7. Gaoyang Liu, Faguo Hou, Xindong Wang, et al. Conductive Polymer and Nanoparticle-Promoted Polymer Hybrid Coatings for Metallic Bipolar Plates in Proton Membrane Exchange Water Electrolysis. Applied Sciences, 2023, 13(3): 1244. DOI:10.3390/app13031244
    8. O. K. Alekseeva, N. A. Ivanova, V. V. Tishkin, et al. Nanostructured Coatings to Extend the Component Lifetime in Electrochemical Devices Based on Proton Exchange Membrane. Nanobiotechnology Reports, 2023, 18(S2): S225. DOI:10.1134/S2635167624600147
    9. Gaoyang Liu, Dongfang Shan, Baizeng Fang, et al. Novel hybrid coating of TiN and carbon with improved corrosion resistance for bipolar plates of PEM water electrolysis. International Journal of Hydrogen Energy, 2023, 48(50): 18996. DOI:10.1016/j.ijhydene.2023.02.015
    10. Xindong Wang, Ranbo Yu, Chun Zhan, et al. Editorial for special issue on advanced energy storage and materials for the 70th Anniversary of USTB. International Journal of Minerals, Metallurgy and Materials, 2022, 29(5): 905. DOI:10.1007/s12613-022-2490-8
    11. Pasquale Cavaliere. Water Electrolysis for Hydrogen Production. DOI:10.1007/978-3-031-37780-8_6
    12. Huang Jian, Xu Guizhi, Song Jie, et al. PEM Water Electrolysis. DOI:10.1016/B978-0-443-24062-1.00004-8

    Other cited types(0)

Catalog

    Figures(6)  /  Tables(1)

    Share Article

    Article Metrics

    Article views (2009) PDF downloads (92) Cited by(12)

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return