The Influence of Porous Features on the Electrochemical Performance of Its Supported Platinum Catalyst in Porous Carbon Nanofibers

Abstract

Porous carbon nanofibers (PCNFs) were prepared through electrospinning, pre-oxidation and carbonization with polyacrylonitrile (PAN) as carbon precursor and polymethyl methacrylate (PMMA), CaCO3 as pore-forming agents. The structure, morphology, specific surface area and electrochemical performance of the carbon nanofibers were characterized by scanning electron microscopy (SEM), transmission electron microscopy (TEM), N2 adsorption/desorption method and electrochemical tests. Compared with PCNFs without CaCO3, PCNFs(CaCO3 1%) had higher specific surface area, better dispersion of Pt nanoparticles, and the particle size become smaller, which was corresponding with the results of electrochemical performance test. It could be seen in cyclic voltammetry (CV) and linear sweep voltammetry (LSV) test, ECSA of Pt/PCNFs (CaCO3 1%) attained 82 m2?g?1, while that of JM20 and Pt/PCNFs without CaCO3 were 77 m2?g?1 and 60 m2?g?1, respectively. These results revealed that CaCO3 as the second pore-forming agent can further increase the mesoporous number and specific surface area of nanofibers, and can improve the electrochemical properties of Pt catalyst as the support.

Share and Cite:

Mao, F. , Wang, Y. , Fu, K. , Jin, J. , Yang, S. and Li, G. (2017) The Influence of Porous Features on the Electrochemical Performance of Its Supported Platinum Catalyst in Porous Carbon Nanofibers. Journal of Materials Science and Chemical Engineering, 5, 10-18. doi: 10.4236/msce.2017.57002.

1. Introduction

Nowadays with the shortage of energy resource and serious pollution of our environment, fuel cells with their advantage of being a clean and efficient energy have been the focus of the recent studies. The electrocatalyst plays a decisive role in the performance of fuel cells. At present, platinum catalyst is mainly used in the fuel cell, which has great catalytic activity, chemical stability, and exchange current density [1].

Carbon support is one of the key components of the catalyst, which not only directly affects the particle size, dispersion and activity of the catalyst, but also affects the conductivity, mass transfer and specific surface area of the catalyst. [2] For carbon support, the size and distribution of the pores on it affects the dispersion and size of Pt nanoparticles, which in turn affects the performance of the catalyst. Commonly used support materials are mesoporous carbon, carbon nanotubes, porous carbon nanofibers and graphene [3]-[9]. There has been research focus on PCNFs as support. Álvarez et al. [10] found that Pt/CNFs catalysts show a better performance when compared to Pt/Vulcan XC-72R catalyst in the same fuel cell environment, may result from the smaller number of micropores in carbon nanofibers increasing the use of the Pt catalyst. This indicated the possibility of improving the electrochemical performance of the catalyst by improving the feature of the pores in the carbon nanofibers as support.

The previous work [11] showed that PCNFs were successfully prepared from polyacrylonitrile/polymethyl methacrylate (PAN/PMMA) blend solution by electrospinning, preoxidation and carbonization steps, and got better electrocatalytic performance on its supported platinum catalyst than Commercial electrocatalyst. In order to further improve the specific surface area and increase the abundance of the pores on PCNFs, this paper introduced CaCO3 as the second pore-forming agent into the PAN/PMMA system, and more pores were formed by CO2 derived from thermal decomposition of CaCO3.

2. Experimental

2.1. Materials

Dimethylacetamide (DMAc), ethylene glycol (EG), Methanol and hexachloroplatinic acid (H2PtCl6∙6H2O) are purchased from Sinopharm Chemical Reagent Co., Ltd (Shanghai, China) in analytical purity. Nano-CaCO3 was purchased from Xintai Nano CO., China. Commercial electrocatalyst, 20 wt% Pt on Vulcan XC-72 was purchased from Johnson Matthey Corp, which was denoted as JM20. 5% Nafion solution were purchased from DuPont.

2.2. Fabrication of PCNFs

The precursor nanofibers were prepared by a common electrospinning method. In a typical procedure, 0 g, 0.05 g, 0.1 g, 0.2 g and 0.3 g of nano-CaCO3 were dispersed in 9 g of DMAc by sonication for 30 minutes, respectively. And then 0.7 g of PAN and 0.3g PMMA were added, and the mixture was heated at 40˚C for 4 h under stirring to form a homogeneous suspension. The milky suspension was loaded into a plastic syringe equipped with a 23 gauge needle made of stainless steel. The needle was connected to a high-voltage supply. A piece of grounded aluminum foil was placed 18 cm below the tip of the needle to collect the nanofibers. The voltage was set at 18 kV, and the solution feeding rate was 0.9 mL∙h−1. The white nonwoven mat composed of composite nanofibers was collected on the aluminum foil collector. The composite nanofibers obtained by electrospinning were preoxidized in air at 280˚C for 1 h and then carbonized at 1400˚C for 2 h under nitrogen protection to obtain porous carbon nanofibers. The PCNFs we obtained were denoted as PCNFs, PCNFs (CaCO3 0.5%), PCNFs (CaCO3 1%), PCNFs (CaCO3 2%) and PCNFs (CaCO3 3%).

2.3. Preparation of the Electrocatalysts

The preparation method and electrochemical evaluation of the catalyst are both based on the report of the research group. [11] 40 mg of PCNFs were put into a three-necked flask and mixed with 30 ml of EG and 15 ml deionized water under ultrasonic treatment for 1h. Then 2 ml of H2PtCl6∙6H2O was added into the slu- rry slowly to ensure about 20 wt% Pt content, and continue to ultrasound for 1 h. After completion of the sonication, the mixture was stirred for 1 h and then heated to 140˚C for 4 h under N2. After cooling down to room temperature, the mixture was washed with deionized water and acetone for several times in centrifuge. And then dried at 80˚C in vacuum oven for 12 h.

2.4. Electrochemical evaluation of Electrocatalysts

Three-electrode system was used for electrochemical test at room temperature. In the preparation of working electrode, 4 mg catalyst was dripped into 2 ml of Nafion/methanol solution and ultrasonicated for 1 h to get a homogeneous catalyst ink. Then, 20 μl ink was dripped onto a clean glass carbon working ring disk electrode (RDE: 0.2472 cm2) using a pipette, and dry the ink by infrared lamp to bake for about 30 seconds. The liquid electrolyte was 0.1 M HClO4 aqueous solvent, Pt foil was used as the counter electrode and saturated calomel electrode (SCE) was employed as the reference electrode. All potentials in this work are referred to the reversible hydrogen electrode (RHE). The liquid electrolyte was bubbled with N2 and O2 about 30 minutes for cyclic voltammetry (CV) and linear sweep voltammetry (LSV) test.

3. Results and Discussion

Figure 1 showed the morphological structure of PCNFs. It can be seen that the diameter of PCNFs was about 200 nm, and PCNFs had mesoporous structure, which were present at both the surface and the cross section. Moreover, surface of PCNFs were defective and rough, which was benefit the distribution and deposition of Pt nanoparticles. Comparing PCNFs (CaCO3) with PCNFs without CaCO3, the PMMA-forming pores mainly appeared on the surface of the fiber, showing a gully-like shape. While the CaCO3-forming pores mainly distributed inside the fiber and can be observed in the section. In addition, with the increase of CaCO3, the pores of the fiber cross-section were significantly increased, when the content reached 3%, the cross-section showing blasting pores.

The porosity of the PCNFs samples were determined by the N2 adsorp-

Figure 1. SEM images of PCNFs (a); PCNFs(CaCO3 0.5%) (b); PCNFs(CaCO3 1%) (c); PCNFs(CaCO3 2%)(d); and PCNFs(CaCO3 3%)(e).

tion/desorption method. Figure 2 showed the N2 adsorption/desorption isotherms of PCNFs, PCNFs (CaCO3 0.5%), PCNFs (CaCO3 1%), PCNFs (CaCO3 2%) and PCNFs (CaCO3 3%). The BET surface areas were shown in Table 1. Pore-forming agents in the samples were PMMA and CaCO3. Mesoporous was formed by decomposing of PMMA at about 280˚C. CaCO3 was thermally decomposed into CaO and CO2 during carbonation, and the released CO2 would be an active ingredient, which contacted with the C elements in nanofibers and consumed it in the nanofibers according to the equation CO2 + C = CO and then formed mesoporous [12] [13]. The specific surface area gradually increased with the addition of CaCO3 content. However, when the CaCO3 content was too large, it would be agglomerated together to form macropores, therefore the specific surface area was reduced.

Figure 3 showed the TEM images of as-prepared PCNFs without CaCO3 (Fig. 3 (a)) and PCNFs (CaCO3) (Figures 3(b)-(e)). The Pt nanoparticles simultaneously distributed in the surface, cross section and pore channels as shown in the figure. With the addition of CaCO3 content changes, the uniformity of the

Figure 2. Nitrogen adsorption/desorption isotherms of PCNFs, PCNFs (CaCO3 0.5%), PCNFs (CaCO3 1%), PCNFs (CaCO3 2%), PCNFs (CaCO3 3%)

Figure 3. TEM images of PCNFs (a), PCNFs(CaCO3 0.5%) (b), PCNFs(CaCO3 1%)(c), PCNFs(CaCO3 2%)(d), and PCNFs(CaCO3 3%)(e)

Table 1. BET surface area of PCNFs.

dispersion of Pt nanoparticles were different. The nanoparticles uniformity of PCNFs, PCNFs (CaCO3 0.5%) and PCNFs (CaCO3 1%) are relatively good compared with PCNFs (CaCO3 2%) and PCNFs (CaCO3 3%). As can be clearly seen, partly aggregation of Pt particles occurs in PCNFs (CaCO3 2%) and PCNFs (CaCO3 3%), which resulted from a larger pore size on the fiber than that of Pt nanoparticles. Average size of Pt nanoparticles was approximately ranged from 2 to 4 nm, when Pt showed smaller size distribution on PCNFs (CaCO3 1%) than that on PCNFs without CaCO3. The reason for the difference of dispersion of Pt nanoparticles in these samples may result from the change of size and quantity of mesoporous in the nanofibers.

Figure 4 showed the cyclic voltammograms (CVs) of the samples. As ob served, the CVs had obvious hydrogen adsorption/desorption peaks from 0~0.4 V and Pt oxidation/reduction peaks from 0.6 - 0.9 V. Obviously, the samples showed large area of hydrogen absorption. The electrochemical active surface (ECSA) which is an significant parameter used to represent the intrinsic activity of the Pt catalysts can be calculated based on the equation: ECSA = Q/(0.21 × M), where Q (mC) is the electrical charge associated with monolayer adsorption of hydrogen on Pt with an assuming value of 210 μC/cm2, M is the Pt loading on the working electrode [14]. The results of ECSAs were about 77 m2∙g−1 and 82 m2∙g−1 for JM20 and Pt/PCNFs (CaCO3 1%), respectively (Table 2). ECSA values explained that the Pt/PCNFs (CaCO3 1%) exhibited the highest electrocatalytic activity among these six catalysts. Besides the uniform dispersion of Pt nanoparticles, more mesoporous might provide the area for more Pt nanoparticles to load, and also supply extra and more active sites which could enhance electrochemistry activity.

The ORR activity of samples was obtained by rotating disk measurement (RDE) at 1600 rpm and was shown in Figure 5. As seen in the figure, the Pt/ PCNFs (CaCO3 1%) exhibited the superior ORR activity compared with other

Figure 4. CV curves of Pt/PCNFs, Pt/PCNFs (CaCO3 0.5%), Pt/PCNFs (CaCO3 1%), Pt/PCNFs (CaCO3 2%) and Pt/PCNFs (CaCO3 3%) in 0.1 M HClO4 at ambient solution temperature with a potential scanning rate of 50 mV∙s−1.

Figure 5. LSV curves of Pt/PCNFs, Pt/ PCNFs (CaCO3 0.5%), Pt/PCNFs (CaCO3 1%), Pt/PCNFs (CaCO3 2%) and Pt/PCNFs (CaCO3 3%) in 0.1 M HClO4 at ambient temperature with a potential scanning rate of 50 mV∙s−1 and 1600 rpm electrode rotation speed.

Table 2. Totally results of the CV and LSV measurement.

catalysts. The onset potential of Pt/PCNFs (CaCO3 1%) was 0.93 V, 30 mV higher than Pt/PCNFs and even 20 mV higher than commercial JM20. In addition, the half-wave potentials for Pt/PCNFs (CaCO3 1%) were 0.799 V, 26 mV higher than JM20 and 37 mV higher than Pt/PCNFs, respectively. The results displayed that the increase of mesoporous greatly facilitated the catalytic activity.

4. Conclusion

In summary, PCNFs have been successfully prepared by using nano-CaCO3. PCNFs (CaCO3 1%) had superior specific surface area compared with PCNFs without CaCO3, indicating that the introduction of CaCO3 could improve the structure of fibers and quantity of the mesoporous. Furthermore, PCNFs (CaCO3 1%) also showed better electrocatalytic activity than PCNFs without CaCO3 and even JM20, demonstrating the improvement of the structure and the increase in number of mesoporous which can make the dispersion of Pt nanoparticles more uniform and the size of it more smaller, which is beneficial to improve the electrochemical performance of Pt/PCNFs.

Conflicts of Interest

The authors declare no conflicts of interest.

References

[1] Xin, Y., Liu, J.G., Zhou, Y., Liu, W., Gao, J., Xie, Y., Yin, Y. and Zou, Z. (2011) Preparation and Characterization of Pt Supported on Graphene with Enhanced Electrocatalytic Activity in Fuel Cell. Journal of Power Sources, 196, 1012-1018. https://doi.org/10.1016/j.jpowsour.2010.08.051
[2] Liu, H.S., Song, C.J., Zhang, L., Zhang, J.J., Wang, H.J. and Wilkinson, D.P. (2006) A Review of Anode Catalysis in the Direct Methanol Fuel Cell. Journal of Power Sources, 155, 95-110. https://doi.org/10.1016/j.jpowsour.2006.01.030
[3] Hulicova, D., Kodama, M. and Hattori, H. (2006) Electrochemical Performance of Nitrogen-Enriched Carbons in Aqueous and Non-Aqueous Supercapacitors. Chemistry of Materials, 18, 2318-2326. https://doi.org/10.1021/cm060146i
[4] Wei, D., Liu, Y., Wang, Y., Zhang, H., Huang, L. and Yu, G. (2009) Synthesis of N-Doped Graphene by Chemical Vapor Deposition and Its Electrical Properties, Nano Letters, 9, 1752-1758. https://doi.org/10.1021/nl803279t
[5] Zhou, C.W., Kong, J., Yenilmez, E. and Dai, H.J. (2000) Modulated Chemical Doping of Individual Carbon Nanotubes. Science, 290, 1552-1555. https://doi.org/10.1126/science.290.5496.1552
[6] Lv, R., Cui, T., Jun, M.S., Zhang, Q., Cao, A., Su, D.S., Zhang, Z., Yoon, S.H., Miyawaki, J., Mochida, I. and Kang, F. (2011) Open-Ended, N-Doped Carbon Nanotube—Graphene Hybrid Nanostructures as High-Performance Catalyst Support, Advanced Functional Materials, 21, 999-1006. https://doi.org/10.1002/adfm.201001602
[7] Zhang, L.S., Liang, X.Q., Song, W.G. and Wu, Z.Y. (2010) Identification of the Ni-trogen Species on N-Doped Graphene Layers and Pt/NG Composite Catalyst for Direct Methanol Fuel Cell, Physical Chemistry Chemical Physics, 12, 12055-12059. https://doi.org/10.1039/c0cp00789g
[8] Jafri, R.I., Rajalakshmi, N. and Ramaprabhu, S. (2010) Nitrogen Doped Graphene Nanop-latelets as Catalyst Support for Oxygen Reduction Reaction in Proton Exchange Membrane Fuel Cell. Journal of Materials Chemistry, 20, 7114-7117. https://doi.org/10.1039/c0jm00467g
[9] Maldonado, S. and Stevenson, K.J. (2005) Influence of Nitrogen Doping on Oxygen Re-duction Electro Catalysis at Carbon Nanofiber Electrodes. Journal of Physical Chemistry, 109, 4707-4716. https://doi.org/10.1021/jp044442z
[10] álvarez, G., Alcaide, F., Cabot, P.L., Lázaro, M.J., Pastor, E. and Solla-Gullón, J. (2012) Electrochemical Performance of Low Temperature PEMFC with Surface Tailored Carbon Nanofibers as Catalyst Support. International Journal of Hydrogen Energy, 37, 393-404. https://doi.org/10.1016/j.ijhydene.2011.09.055
[11] Wang, Y., Jin, J., Yang, S.L., Li, G. and Qiao, J.L. (2015) Highly Active and Stable Platinum Catalyst Supported on Porous Carbon Nanofibers for Improved Performance of PEMFC. Electrochimica Acta, 177, 181-189. https://doi.org/10.1016/j.electacta.2015.01.134
[12] Zhao, C., Wang, W., Yu, Z., Zhang, H., Wang, A. and Yang, Y. (2010) Na-no-CaCO3 as Template for Preparation of Disordered Large Mesoporous Carbon with Hierarchical Porosities. J. Mater. Chem, 20, 976-980. https://doi.org/10.1039/B911913B
[13] Blackman, J.M., Patrick, J.W., Arenillas, A., Shi, W. and Snape, C.E. (2006) Activation of Carbon Nanofibers for Hydrogen Storage. Carbon, 44, 1376-1385. https://doi.org/10.1016/j.carbon.2005.11.015
[14] Wang, Y., Jin, J., Yang, S.L., Li, G. and Jiang, J.M. (2016) Nitrogen-Doped Porous Carbon Nanofiber Based Oxygen Reduction Reaction Electrocatalysts with High Activity and Durability. International Journal of Hydrogen Energy, 41, 11174- 11184. https://doi.org/10.1016/j.ijhydene.2016.04.235

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.