Hydrogen Electrosorption into Pd−Cd Nanostructures - American

Jan 25, 2010 - Department of Chemistry, Lakehead University, 955 Oliver Road, Thunder Bay, ON P7B 5E1, Canada. Received November 20, 2009. Revised ...
0 downloads 0 Views 3MB Size
pubs.acs.org/Langmuir © 2010 American Chemical Society

Hydrogen Electrosorption into Pd-Cd Nanostructures Brian D. Adams, Cassandra K. Ostrom, and Aicheng Chen* Department of Chemistry, Lakehead University, 955 Oliver Road, Thunder Bay, ON P7B 5E1, Canada Received November 20, 2009. Revised Manuscript Received January 2, 2010 Hydrogen-absorbing materials are crucial for both the purification and storage of hydrogen. Pd and Pd-based alloys have been studied extensively for their use as both hydrogen dissociation catalysts and hydrogen selective membrane materials. It is known that incorporating metal atoms of different sizes into the Pd lattice has a major impact on the hydrogen absorption process. In this paper, hydrogen electrosorption into nanostructured Pd-Cd alloys has been studied for different compositions of Cd that varied from 0 to 15 at. %. The low cost of Cd makes it an attractive material to combine with Pd for hydrogen sorption. A combination of chronoamperometry and cyclic voltammetric experiments was used to determine the ratio of the H/(Pd þ Cd) and the kinetics of hydrogen sorption into these Pd-Cd alloys at different potentials. It was found that the maximum H/(Pd þ Cd) value was 0.66 for pure Pd, and this decreased with increasing the amount of Cd. Also, the R (solid solution) to β phase (metal hydride) hydrogen transition was determined to be the slowest step in the absorption process and was practically eliminated when an optimum amount of Cd atoms was doped (i.e., Pd-Cd(15%)). With increasing the amount of Cd, more hydrogen was absorbed into the Pd-Cd nanostructures at the higher potentials (the R phase region). The faster kinetics, along with the decrease in the phase transition of hydrogen sorption into the Pd-Cd nanostructures when compared to pure Pd, makes the Pd-Cd nanostructures attractive for use as a hydrogen dissociation catalytic capping layer for other metal hydrides or as a hydrogen selective membrane.

1. Introduction Given the current global situation involving limited energy resources and rising pollution linked to traditional energy production, there is a need for cleaner, cheaper, and more efficient energy technologies.1-4 A hydrogen economy is thought to be one possible solution to the world’s energy crisis; however, we are a long way from making this a reality for two reasons: (i) pure hydrogen is not a resource and must first be produced and purified; (ii) storing of hydrogen is a difficult task as it is gaseous at atmospheric conditions. The objective of the hydrogen economy is to be able to make use of renewable energy sources such as solar or wind energy to produce hydrogen from the splitting of water and, then, to harness the energy through the recombination of hydrogen and oxygen in a hydrogen fuel cell. Hydrogen-absorbing materials are crucial for both the purification and storage of hydrogen. Many types of hydrogen storage materials have been studied, and it is thought that inexpensive, lightweight, and high capacity metal hydrides, such as MgH2, are the best candidates for practical applications.4-6 There are two major problems with the MgH2: (i) the sorption kinetics involved are quite sluggish, and (ii) MgH2 has a high thermodynamic stability requiring temperatures greater than 300 °C for the *Corresponding author: tel þ1 807 343 8318; fax þ1 807 346 7775; e-mail [email protected]. (1) Borgschulte, A.; Rector, J. H.; Schreuders, H.; Dam, B.; Griessen, R. Appl. Phys. Lett. 2007, 90, 071912. (2) Qu, J.; Wang, Y.; Xie, L.; Zheng, J.; Liu, Y.; Li, X. Int. J. Hydrogen Energy 2009, 34, 1910–1915. (3) Jeng, R.-R.; Lee, S.-L.; Hsu, C.-W.; Wu, Y.-P.; Lin, J.-C. J. Alloys Compd. 2008, 464, 467–47. (4) Pozzo, M.; Alfe, D. Int. J. Hydrogen Energy 2009, 34, 1922–1930. (5) Shao, H.; Liu, T.; Wang, Y.; Xu, H.; Li, X. J. Alloys Compd. 2008, 465, 527– 533. (6) Gross, A. F.; Ahn, C. C.; Van Atta, S. L.; Liu, P.; Vajo, J. J. Nanotechnology 2009, 20, 204005. (7) Friedrichs, O.; Kolodziejczyk, L.; Sanchez-Lopez, J. C.; Fernandez, A.; Lyubenova, L.; Zander, D.; K€oster, U.; Aguey-Zinsou, K. F.; Klassen, T.; Bormann, R. J. Alloys Compd. 2008, 463, 539–545.

7632 DOI: 10.1021/la9044072

desorption of hydrogen.4,7,8 Electrochemical loading of MgH2 is also problematic due to corrosion and degradation in both alkaline and acidic electrolytes and subsequent loss of hydrogen capacity.7,9,10 Palladium, on the other hand, is a material that can reversibly store a very large volumetric amount of hydrogen at room temperature and is, itself, extremely stable.11 The use of Pd as the sole metal hydride for hydrogen storage is not practical because it is expensive and has poor gravimetric capacity. Since many metals have slow hydrogenation kinetics, often a catalyst is necessary to dissociate the hydrogen molecule. A capping layer of Pd can be used for this purpose as this not only facilitates the hydrogen uptake but also protects the surface from corrosion.8,12-15 A Pd capping of 2-5 nm is usually sufficient, and it is essential to use nanosized particles to minimize costs.12 Pd and Pd-alloys are also used in industry for hydrogen selective membranes.16-20 Hydrogen selective membranes are important for technological processes such as the separation and recovery of hydrogen from process steam reforming, purification of hydrogen for fuel cell applications, and chemical (8) Tang, F.; Parker, T.; Li, H.-F.; Wang, G.-C.; Lu, T.-M. Nanotechnology 2008, 19, 465706. (9) Ma, T.; Hatano, Y.; Abe, T.; Watanabe, K. J. Alloys Compd. 2005, 391, 313– 317. (10) Xiao, X.; Chen, L.; Hang, Z.; Wang, X.; Li, S.; Chen, C.; Lei, Y.; Wang, Q. Electrochem. Commun. 2009, 11, 515–518.  Dekany, I. Langmuir 2003, (11) Berger, F.; Varga, M.; Mulas, G.; Molnar, A.; 19, 3692–3697. (12) Remhof, A.; Borgschulte, B. ChemPhysChem 2008, 9, 2440–2455. (13) Singh, S.; Eijt, S. W. H.; Zandbergen, M. W.; Legerstee, W. J.; Svetchnikov, V. L. J. Alloys Compd. 2007, 441, 344–351. (14) Paillier, J.; Roue, L. J. Alloys Compd. 2005, 404, 473–476. (15) Lebouin, C.; Soldo-Olivier, Y.; Sibert, E.; De Santis, M.; Maillard, F.; Faure, R. Langmuir 2009, 25, 4251–4255. (16) Sun, Y.; Tao, Z.; Chen, J.; Herricks, T.; Xia, Y. J. Am. Chem. Soc. 2004, 126, 5940–5941. (17) Foletto, E. L.; Da Silveira, J. V. W.; Jahn, S. L. Lat. Am. Appl. Res. 2008, 38, 79–84. (18) Lin, W.-H.; Chang, H.-F. Surf. Coat. Technol. 2005, 194, 157–166. (19) Li, H.; Xu, H.; Li, W. J. Membr. Sci. 2008, 324, 44–49. (20) Zeng, G.; Goldbach, A.; Xu, H. J. Membr. Sci. 2009, 326, 681–687.

Published on Web 01/25/2010

Langmuir 2010, 26(10), 7632–7637

Adams et al.

reactions in membrane reactors.17 A major issue for hydrogen filters made from pure Pd is hydrogen embrittlement.19 The large expansion of the lattice constant due to the R (solid solution) to β phase (metal hydride) hydrogen transition can cause cracking of the membrane. By alloying palladium with other metals, the lattice sizes and the number of vacant d-band holes change remarkably. The hydrogen solubility in different phases, the plateau pressure, and the extent of the miscibility gap vary from alloy to alloy. This also occurs with varying the amount of alloying elements.21 The extent of the lattice expansion and contraction and brittleness causing can be controlled by alloying Pd with metals which have larger atoms than Pd itself. In PdAg, for example, which is the most common hydrogen extraction metal used by industry,16 the lattice has already been expanded by silver atoms, and the PdAg lattice is less influenced by hydrogen and thus less brittle than a pure Pd lattice.17 Pd and Pd-alloys can be reversibly loaded with hydrogen and studied in both the gas phase and under electrochemical conditions.22-30 Extremely large currents generated from the electrochemical absorption and desorption of hydrogen in bulk Pd (wires or foils) can be avoided by using so-called “limited volume electrodes” (LVEs).31-36 These LVEs consist of a thin layer of Pd or Pd-alloy deposited onto a matrix, such as gold, which does not absorb hydrogen. LVEs allow for the control of the amount of absorbed hydrogen in the thin layer. Pd25-27,32,37 and various Pdalloys have been studied this way, including PdPt,33 PdRh,35,36 PdNi,31 PdP,33 and PdAu.38 All of the Pd-based materials studied using this technique were prepared using electrodeposition. In the past, it has been found that nanostructured Pd can store more hydrogen than bulk Pd metal due to trapping of hydrogen in grain boundaries.16,26,39 This also gives better resistance to hydrogen embrittlement.19 Cadmium is a relatively abundant element. It is used in batteries, cadmium pigments, coatings, and plating alloys as stabilizers for plastics although it is a potential environmental hazard. Our recent studies reveal that PdCd nanostructures are promising for hydrogen storage.40 By incorporating the larger Cd atoms into the Pd crystal lattice, the lattice expansion due to hydrogen absorption can be effectively (21) Fazle, A. K. M.; Sakamoto, Y. Int. J. Hydrogen Energy 2000, 25, 853–859. (22) Kishore, S.; Nelson, J. A.; Adair, J. H.; Eklund, P. C. J. Alloys Compd. 2005, 389, 234–242. (23) Yamauchi, M.; Ikeda, R.; Kitagawa, H.; Takata, M. J. Phys. Chem. C 2008, 112, 3294–3299. (24) Kobayashi, H.; Yamauchi, M.; Kitagawa, H.; Kubota, Y.; Kato, K.; Takata, M. J. Am. Chem. Soc. 2008, 130, 1828–1829. (25) Gabrielli, C.; Grand, P. P.; Lasia, A.; Perrot, H. J. Electrochem. Soc. 2004, 151, A1925–A1936. (26) Gabrielli, C.; Grand, P. P.; Lasia, A.; Perrot, H. J. Electrochem. Soc. 2004, 151, A1937–A1942. (27) Martin, M. H.; Lasia, A. Electrochim. Acta 2008, 53, 6317–6322. (28) yosiewicz, B.; Birry, L.; Lasia, A. J. Electroanal. Chem. 2007, 611, 26–34. (29) Lasia, A. J. Electroanal. Chem. 2006, 593, 159–166. (30) Duncan, H.; Lasia, A. Electrochim. Acta 2008, 53, 6845–6850. (31) Grden, M.; Czerwinski, A.; Golimowski, J.; Bulska, E.; Krasnode- bskaOstre-ga, B.; Marassi, R.; Zamponi, S. J. Electroanal. Chem. 1999, 460, 30–37. (32) Czerwinski, A.; Kiersztyn, I.; Grden, M.; Czapla, J. J. Electroanal. Chem. 1999, 471, 190–195. (33) Grden, M.; Piascik, A.; Koczorowski, Z.; Czerwinski, A. J. Electroanal. Chem. 2002, 532, 35–42. (34) Szojewski, M.; Kowalska, J.; Jurczakowski, R. J. Phys. Chem. C 2009, 113, 3707–3712. _ (35) Zurowski, A.; yukaszewski, M.; Czerwinski, A. Electrochim. Acta 2008, 53, 7812–7816. _ (36) Zurowski, A.; yukaszewski, M.; Czerwinski, A. Electrochim. Acta 2006, 51, 3112–3117. (37) Duncan, H.; Lasia, A. Electrochim. Acta 2008, 53, 6845–6850. (38) yukaszewski, M.; Czerwinski, A. J. Solid State Electrochem. 2008, 12, 1589–1598. (39) Stuhr, U.; Striffler, T.; Wipf, H.; Natter, H.; Wettmann, B.; Janssen, S.; Hempelmann, R.; Hahn, H. J. Alloys Compd. 1997, 253, 393–396. (40) Adams, B. D.; Wu, G.; Nigro, S.; Chen, A. J. Am. Chem. Soc. 2009, 131, 6930–6931.

Langmuir 2010, 26(10), 7632–7637

Article

controlled and tuned. Further knowledge of the hydrogenabsorbing properties of promising PdCd nanostructures is essential for their potential application as hydrogen dissociation catalyst capping layers or hydrogen selective membrane materials. In this study, we further investigate the electrochemical sorption of hydrogen at PdCd nanostructured alloys with the amount of Cd varied from 0 to 15 at. %. The nanostructured PdCd alloys were directly grown onto Ti substrates using a hydrothermal reduction method. For the first time, the effects of sweep rate, potential, and composition of Cd on hydrogen adsorption have been systemically studied. Our study has shown that the R to β phase transition is the slowest step in the hydrogen absorption process, which can be practically eliminated when optimum amount of Cd atoms is doped.

2. Experimental Methods 2.1. Fabrication and Surface Analysis of the PdCd Nanostructures. The PdCd nanostructures were directly grown onto

Ti plates using a facile hydrothermal reduction method.41,42 The Ti substrate plates (1.25 cm  0.80 cm  0.5 mm) were first degreased in an ultrasonic bath of acetone for 10 min followed by 10 min in pure water (18 MΩ) obtained from a NANOpure Diamond UV ultrapure water purification system. The substrates were then etched in 18% HCl at ∼85 °C for 30 min. The etched Ti substrates were transferred into Teflon-lined autoclaves, and 10 mL of an aqueous mixture of inorganic metal precursors and the reducing agent was added. Ammonium formate (Aldrich, 99.995%) was used as the reducing agent, and the metal precursors were PdCl2 (Alfa Aesar, 99.9%) and Cd(NO3)2 3 4H2O (Alfa Aesar, 99.9%). In all cases, the amounts of the PdCl2 precursor and reducing agent added were kept constant at 5 mM and 1 M, respectively. Varying amounts of the Cd(NO3)2 3 4H2O precursor were added to obtain the desired ratio of Pd:Cd. The autoclaves were sealed and heated at 180 °C for 2 h in an oven which had been preheated. After 2 h, the autoclaves were immediately removed from the oven and allowed to cool to room temperature for ∼4 h. The Ti plates coated with the Pd-Cd nanostructures were then removed from the autoclaves, dried under argon, and then rinsed with pure water. The coated Ti plates were annealed by heating them to 250 °C at a ramp rate of 8 °C/min and then held at 250 °C for 2 h in a tube furnace under argon to ensure good surface adhesion of the Pd-Cd particles. The surface morphology of the fabricated samples was characterized using JEOL JSM 590LV scanning electron microscopy (SEM). X-ray photoelectron spectrophotometry was performed using a Thermo Scientific K-Alpha XPS spectrometer. The samples were run at a takeoff angle (relative to the surface) of 90°. A monochromatic Al KR X-ray source was used, with a spot area (on a 90° sample) of 400 μm. The position of the energy scale was adjusted to place the main C 1s feature (C-C) at 284.6 eV. XPSPEAK 4.1 software was used to analyze the raw data and to fit the spectra. Surface composition was examined by energy dispersive X-ray spectrometry (EDS), and the crystallographic properties were analyzed using X-ray diffraction (XRD). 2.2. Hydrogen Electrosorption Measurements. For all electrochemical experiments, a three-electrode cell system and a VoltaLab PGZ402 potentiostat were used. The counter electrode used was a Pt wire coil (10 cm2), and this was cleaned before each experiment by flame-annealing and quenching with pure water. The reference electrode was a saturated calomel electrode (SCE) connected to the cell through a salt bridge. All potentials mentioned herein are with respect to the SCE reference electrode. The working electrodes used were the prepared Ti/PdCd plates (1 cm2). The coating loads of the electrodes, for calculating the (41) Wang, J.; Thomas, D. F.; A Chen, A. Chem. Commun. 2008, 5010–5012. (42) Wang, J.; Asmussen, R. M.; Adams, B.; Thomas, D. F.; Chen, A. Chem. Mater. 2009, 21, 1716–1724.

DOI: 10.1021/la9044072

7633

Article H/PdCd ratios, were determined by weighing the Ti plates before and after being coating with the PdCd nanostructures with a mass balance accurate to 0.01 mg. It was found that all of the electrodes fabricated and studied had similar coating loads between 0.90 and 1.10 mg. The electrolyte for these studies was 0.1 M HClO4, and the solution was deaerated with ultrapure argon gas prior to measurements to remove any dissolved oxygen, and argon gas was continuously bubbled into the solution during the experiments to increase mass transport. All experiments were performed at room temperature. Data acquisition and analysis were performed using VoltMaster 4 software. Hydrogen absorption was performed at a constant potential for 10 min, which was sufficient to ensure full hydrogen saturation, determined by chronoamperometry. A scan rate of 20 mV/s was used for the cyclic voltammetry as this was found to be appropriate to desorb the full amount of hydrogen. The amount of absorbed hydrogen was calculated from the hydrogen oxidation charge, QH, obtained by the integration of voltammetric curves recorded during hydrogen desorption and the mass of the PdCd film and is reported as the H/(Pd þ Cd) atomic ratio.

3. Results and Discussion 3.1. Surface Analysis of Electrodes. After the Pd and Pd-Cd particles were deposited onto the Ti substrates, these plates were examined using SEM. High-resolution images can be seen in our previous study of these Pd-Cd nanostructures.40 A lower magnification image of the Pd-Cd(10%) surface is displayed in Figure 1, showing highly even surface coverage and small dendritic structures, which is consistent with our previous work. X-ray photoelectron spectroscopic (XPS) studies of the Pd-Cd(10%) sample verified the composition and revealed an electronic interaction between the Pd and Cd species upon alloying. Figure 2 shows the high-resolution XPS spectra for the Pd 3d and Cd 3d regions. It was found that the actual surface contains a mixture of metallic Pd and Cd as well as Pd and Cd in higher oxidation states. The Pd 3d5/2 peak is located at 335.2 eV, and the Cd 3d5/2 peak is located at 404.0 eV. XPS was also performed on a Pd sample to examine the electronic interaction upon the addition of Cd. For pure Pd, the 3d5/2 peak is located at 335.5 eV, indicating a negative shift of 0.3 eV for PdCd(10%). The negative shift of the Pd 3d peak suggests that there is a slight electron transfer from Cd to Pd, which agrees with the electronegativity series (2.20 for Pd and 1.69 for Cd). Using the area of the fitted curves for the Pd and Cd in both the metallic and higher oxidation states, the ratio of metal/metal oxide was calculated. It was found that 88% of the Pd was in its metallic state and 76% of the Cd was in its metallic state. Both of these values are quite high, indicating that the hydrothermal method of preparing the Pd-Cd nanostructures is efficient, and since only the surface layer was examined, the bulk of the Pd-Cd phase could be pure metals. The actual atomic composition of Cd was calculated, from the area under the peaks, to be 11.46%. This is very close to the nominal ratio of 90:10 palladium to cadmium. 3.2. General Cyclic Voltammetric Behavior of the Pd-Cd Nanostructures. Figure 3 shows typical cyclic voltammograms (CVs) of the thin films of the PdCd nanostructures in 0.1 M HClO4. There are three regions present in these cyclic voltammograms: the hydrogen region (1 and 2), the oxide region (3 and 4), and the double-layer region (5). In the hydrogen region, absorption/adsorption (1) and subsequent desorption (2) take place, and in the oxide region, surface oxide formation (oxygen adsorption, 3) and oxide reduction (oxide desorption, 4) occur. These two regions are separated by the double-layer region: a potential range where no faradaic current flows.38 It can be seen, however, that at this potential scan rate the double layer almost 7634 DOI: 10.1021/la9044072

Adams et al.

Figure 1. Typical SEM image of the Pd-Cd(10%) electrode surface at 6000 magnification.

Figure 2. XPS spectra of the Pd 3d (a) and Cd 3d (b) regions for the Pd-Cd(10%) sample. The dots, dashed lines, and red, blue, and black solid lines represent the raw data, baseline, individual components (zero and high oxidation states), and total fit, respectively.

disappears for the PdCd(10%) and PdCd(15%) nanostructures as the current due to the desorption of hydrogen carries into this range. It can also be noted that only one set of peaks are visible for hydrogen absorption/adsorption and desorption. This is thought to be because of the diffusion of dissolved hydrogen from the thicker nanostructured films to the surface as compared to the LVEs studied by others which only consist of electrodeposited monolayers.26 The current due to the absorbed (and desorption of Langmuir 2010, 26(10), 7632–7637

Adams et al.

Article

Figure 3. Cyclic voltammograms of the Pd-Cd nanostructures in 0.1 M HClO4 at a scan rate of 20 mV/s.

absorbed) hydrogen from the bulk of the metal hides current due to the adsorption of (and desorption of adsorbed) hydrogen or hydrogen in the R phase of the bulk metal. It should also be noted that, with an increasing amount of Cd, the adsorption/absorption current becomes increasingly negative and there is a negative shift in the peak potential, indicating that the addition of Cd decreases the adsorption energies for the hydrogen underpotential deposition on Pd and that the hydrogen is more weakly bonded on the PdCd nanostructures. The area under both the adsorption and desorption peaks becomes larger with more Cd, indicating a larger amount of hydrogen can be electrosorbed. This is consistent with our previous voltametric studies.40 3.3. Sweep Rate Dependence of Hydrogen Electrosorption in the PdCd Nanostructures. The dependence of the sweep rate on the amount of hydrogen which was absorbed into the PdCd nanostructures was examined. Shown in Figure 4a are the cyclic voltammograms of the hydrogen region at various sweep rates for the PdCd(10%) electrode. The double-layer region becomes clearly visible here as the sweep rate is decreased. The charge due to hydrogen desorption was calculated from the anodic part of the curves and is displayed in Figure 4b. The inset shows the H/(Pd þ Cd) ratio which takes into consideration the mass of the PdCd(10%) coating. The amount of electrosorbed hydrogen increases considerably with a slower sweep rate from a H/(Pd þ Cd) ratio of less than 0.1 with 20 mV/s to about 0.35 for 1 mV/s. In order to better study the hydrogen-sorbing capabilities of the PdCd nanostructures and to attain the maximum amount of hydrogen which can be adsorbed/absorbed, an extremely slow sweep rate is necessary. The electrodes were saturated with hydrogen by holding the potential for a period of time prior to running the CVs. 3.4. Potential and Composition Dependence of Hydrogen Electrosorption in the PdCd Nanostructures. The electrosorption of hydrogen into the PdCd nanostructures was examined at various potentials, from 0 to -325 mV (vs SCE). The potential was first held at a constant potential; cyclic voltammetry was run immediately after. The potential was held for 10 min, which was much longer than necessary, to ensure complete saturation of hydrogen. Figure 5 shows the anodic sweeps (desorption of hydrogen) of the CVs of the Pd and PdCd(10%) nanostructures after complete saturation with hydrogen at various potentials. For Pd, there is a sharp transition between -225 and -250 mV. The amount of desorbed hydrogen at potentials more positive than -225 mV is negligible compared to that at -250 mV and below. For PdCd(10%), the transition is more gradual, and there Langmuir 2010, 26(10), 7632–7637

Figure 4. (a) Cyclic voltammograms of Pd-Cd(10%) in 0.1 M HClO4 at various scan rates. (b) Charge due to hydrogen desorption versus sweep rate calculated by integrating the area under the anodic peaks in (a). Inset shows the H/(Pd þ Cd) ratio versus sweep rate calculated based on the coating load of the Pd-Cd(10%) electrode used.

is still a significant amount of hydrogen desorbed above -225 mV. This transition is the R to β phase transition where both phases exist mutually. For Pd, at potentials more positive than -225 mV, there is exclusive R-phase hydrogen, and at potentials e-250 mV there is exclusive β-phase hydrogen. The H/(Pd þ Cd) ratio was calculated using the same procedure described in section 3.3 for electrodes containing up to 15 at. % Cd at various cathodic potential limits ranging from -325 to 0 mV (Figure 6). Pure Pd exhibits the highest maximum value of H/metal ratio of ∼0.66 at the potential of -325 mV. The maximum H/(Pd þ Cd) ratio is seen at -325 mV for all of the PdCd compositions studied. At potentials more negative than this, a substantial amount of hydrogen evolution was observed, and this caused cracking and loss of the nanostructured coatings to the electrolyte solution. For the electrodes with a very small amount of Cd (i.e., 0 to 5 at. %), a slight decrease in the amount of hydrogen, which may be electrosorbed, is observed from -325 to -250 mV (β phase hydrogen), at which point there is a sharp decrease (β to R phase transition), and only R phase hydrogen is seen at potentials more positive than -225 mV. For the PdCd(10%) and PdCd(15%) nanostructures, the phase transition is very gradual and can barely be distinguished. For these particular electrodes, there is a considerable amount of hydrogen absorbed in the R phase. The H/(Pd þ Cd) ratio does not reach minimal values until 0 mV for PdCd(10%) and PdCd(15%). This is an excellent trait for a hydrogen selective membrane material, as the abrupt phase transition of Pd is what causes embrittlement. DOI: 10.1021/la9044072

7635

Article

Adams et al.

Figure 7. H/(Pd þ Cd) values versus Cd composition at various potentials indicated in the figure.

Figure 5. Anodic sweeps of the desorption of hydrogen from (a) Pd and (b) Pd-Cd(10%) after holding the potential at various cathodic limits for 10 min in 0.1 M HClO4. The scan rate was 20 mV/s.

Figure 6. H/(Pd þ Cd) ratio versus potential for Pd-Cd nano-

potential, it was found that Pd-Cd(15%) has a hydrogen sorption capacity of ∼17 times that of pure Pd. 3.5. Kinetics of the Electrosorption of Hydrogen in the PdCd Nanostructures. Figure 8a presents a series of chronoamperograms for pure Pd after a potential step from the doublelayer region (300 mV) to different potentials in the hydrogen region. After this step to a more negative potential, current begins to flow due to the reduction of Hþ ions to H atoms adsorbed on the surface of the Pd and subsequent entry of the H atoms into the Pd (or PdCd) crystal lattice.38 This forms a bulk phase of absorbed hydrogen in the R or β phase depending on the applied potential. This adsorption/absorption process can be described by the following equations:25,29 Hþ þ M þ e - hMHads

ð1Þ

MHads þ Msubsurface hMsurface þ MHabsðsubsurfaceÞ

ð2Þ

MHabsðsubsurfaceÞ f MHabsðbulk, R þ βÞ

ð3Þ

It should be noted that after a certain amount of time the current reaches a steady state. The current eventually falls to zero at potentials where hydrogen evolution does not occur, as can be expected once the Pd (or PdCd) layer becomes completely saturated with hydrogen. For potentials lower than ∼-250 mV, hydrogen evolution occurs and a negative current results in. This is determined by the rate of the hydrogen evolution reaction at the surface of the electrode according to one of the following equations:25,29

structures with various compositions as indicated in the figure.

The dependence of the H/(Pd þ Cd) ratio on the composition of Cd in the PdCd nanostructures is displayed in Figure 7. It can clearly be seen in this figure that while the amount of β phase hydrogen decreases with an increasing amount of Cd, the amount of R phase hydrogen increases substantially. At potentials more positive than -225 mV (R phase hydrogen region), the solubility of hydrogen is much higher for PdCd alloys than pure Pd. This potential region is analogous to low hydrogen gas pressures in hydrogen sorption measurements carried out in the gas phase.34 This enhanced solubility effect with the addition of Cd is what was seen in our previous work,40 performed at -200 mV. At this 7636 DOI: 10.1021/la9044072

MHads þ Hþ þ e - hM þ H2

ð4Þ

2MHads h2M þ H2

ð5Þ

From the chronoamperometric curves, the time to obtain steady state electrode saturation with hydrogen was determined at various potentials and is displayed in Figure 8b. With a decrease in the electrode potential, the absorption time initially increases until a sharp maximum occurs and then decreases. The time maxima occur at the same potential (-250 mV) for all compositions of PdCd, except the PdCd(15%) which occurs at Langmuir 2010, 26(10), 7632–7637

Adams et al.

Article

t-E plot represents the potential at which the slowest absorption step occurs, revealing that the R to β phase transition is the ratelimiting step in the absorption process rather than surface kinetics or bulk diffusion. With increasing amounts of Cd, the time maxima are decreased so much that with PdCd(15%) only a slight hump is observed, rather than the sharp peak. This fits with the R to β phase transition being the slowest step, as it was shown in section 3.4 that the phase transition is diminished with the addition of Cd. The enhanced solubility of hydrogen in PdCd alloys with respect to pure Pd is attributed almost solely to the larger size of the Cd atoms, similar to PdAg.16-21

4. Conclusions

Figure 8. Chronoamperometric curves for the electrosorption of hydrogen on the Pd electrode recorded in 0.1 M HClO4 after pretreatment at 300 mV (a), the time required for obtaining steady state saturation of the electrodes with absorbed hydrogen with respect to potential (b), and an overlay showing the H/(Pd þ Cd) ratio (triangles) and time to reach steady state (circles) versus potential for the Pd and Pd-Cd(15%) electrodes (c).

-225 mV. A comparison of the potential for time maxima with the respective H/(Pd þ Cd) ratio reveals that the potentials are always within the phase transition region (Figure 8c), which is consistent with the previous studies.34,36 This maximum on the

Langmuir 2010, 26(10), 7632–7637

In summary, we have systemically studied the electrosorption of hydrogen into nanostructured films of Pd-Cd alloys ranging from 0 to 15 at. % Cd. The nanostructures were fabricated using a simple hydrothermal reduction technique. Cyclic voltammetric studies showed that the sweep rate has a major influence on the amount of hydrogen which can be absorbed into the Pd-Cd nanostructures. By decreasing the sweep rate, a larger amount of hydrogen can be electrosorbed, indicating that, if only cyclic voltammetric studies are performed, the amount of electrosorbed hydrogen is limited by the sweep rate. A combination of chronoamperometry and cyclic voltammetry proved to be a more effective method for examining the maximum amount of hydrogen which can be electrosorbed into the Pd-Cd nanostructures. Both potential and composition were found to be major factors in the H/(Pd þ Cd) ratio. The maximum H/(Pd þ Cd) ratio was found to be the highest for pure Pd (0.66). The maximum was found to always occur at -325 mV for every composition, but the value decreased with an increasing amount of Cd. Two distinct phases (R and β), with a sharp phase transition between -225 and -250 mV, were observed for pure Pd; however, this phase transition was diminished with increasing amounts of Cd. In particular, compositions of PdCd alloys with a higher Cd content (i.e., 10 and 15 at. %) barely displayed any phase transition and were capable of electrosorbing much more hydrogen at higher potentials (in the R phase) than pure Pd. This region is analogous to lower pressures of hydrogen sorption in the gas phase. Our study has shown that the phase transition is the slowest step in the hydrogen absorption process, and therefore, with the doping of Cd, the kinetics is much quicker. The combination of enhanced solubility of hydrogen and diminishing of the R and β phase transition in PdCd alloys relative to pure Pd makes the Pd-Cd nanostructures attractive for hydrogen selective membranes and hydrogen dissociation catalysts. Acknowledgment. This work was supported by a Discovery Grant from the Natural Sciences and Engineering Research Council of Canada (NSERC). A. Chen acknowledges NSERC and the Canada Foundation of Innovation (CFI) for the Canada Research Chair Award in Material and Environmental Chemistry. We also thank the Surface Interface Ontario/Chemical Engineering & Applied Chemistry at the University of Toronto for carrying out the XPS analysis of the PdCd samples.

DOI: 10.1021/la9044072

7637