Acessibilidade / Reportar erro

Electrochemical hydrogen-storage properties of La0.78Mg0.22Ni2.67Mn0.11Al0.11Co0.52-M1Ni3.5Co0.6Mn0.4Al0.5 composites

Abstracts

For improving the electrochemical properties of nonstoichiometric AB3-type La0.78Mg0.22Ni2.67Mn0.11Al0.11Co0.52 alloy as negative electrode of Ni-MH battery, its related composites La0.78Mg0.22Ni2.67Mn0.11Al0.11Co0.52-x wt.% M1Ni3.5Co0.6Mn0.4Al0.5 (x = 0, 10, 20, 30) were prepared. Analysis by X-ray diffractometry (XRD) revealed that the composites consist mainly of LaNi5 and La2Ni7 phases. Despite the small decrease in the maximum discharge capacity, the cycle performance was significantly enhanced. Linear polarization (LP), anodic polarization (AP) and potential step discharge experiments revealed that the electrochemical kinetics increases first and then decreases with increasing x.

AB3-type alloy; composite electrode material; cycle stability


Com o objetivo de aprimorar as propriedades eletroquímicas da liga do tipo AB3 La0,78Mg0,22Ni2,67Mn0,11Al0,11Co0,52 como eletrodo negativo de bateria Ni-MH, seus compósitos relacionados La0,78Mg0,22Ni2,67Mn0,11Al0,11Co0,52-x wt.% M1Ni3,5Co0,6Mn0,4Al0,5 (x = 0, 10, 20, 30) foram preparados. A análise por difratometria de raios X (XRD) revelou que os compósitos consistem principalmente das fases LaNi5 e La2Ni7. Apesar da pequena diminuição na capacidade de descarga máxima, o rendimento do ciclo foi significativamente melhorado. Experimentos de polarização linear (LP), polarização anódica (AP) e descarga de potencial degrau revelaram um aumento inicial da cinética eletroquímica que na sequência diminui com o aumento de x.


ARTICLE

Electrochemical hydrogen-storage properties of La0.78Mg0.22Ni2.67Mn0.11Al0.11Co0.52-M1Ni3.5Co0.6Mn0.4Al0.5 composites

Hongxia HuangI,II,* * e-mail: hhxhunan@126.com ; Guohui LiI; Shuxin ZhuangIII

IGuangxi Scientific Experiment Center of Mining, Metallurgy and Environment, College of Chemistry and Bioengineering, Guilin University of Technology, Guilin 541004, People's Republic of China

IIKey Lab of New Processing Technology for Nonferrous Metals & Materials Ministry of Education, Guilin University of Technology, Guilin 541004, People's Republic of China

IIISchool of Material Scienceand engineering, Xiamen University of Technology, Xiamen 361024, People's Republic of China

ABSTRACT

For improving the electrochemical properties of nonstoichiometric AB3-type La0.78Mg0.22Ni2.67Mn0.11Al0.11Co0.52 alloy as negative electrode of Ni-MH battery, its related composites La0.78Mg0.22Ni2.67Mn0.11Al0.11Co0.52-x wt.% M1Ni3.5Co0.6Mn0.4Al0.5 (x = 0, 10, 20, 30) were prepared. Analysis by X-ray diffractometry (XRD) revealed that the composites consist mainly of LaNi5 and La2Ni7 phases. Despite the small decrease in the maximum discharge capacity, the cycle performance was significantly enhanced. Linear polarization (LP), anodic polarization (AP) and potential step discharge experiments revealed that the electrochemical kinetics increases first and then decreases with increasing x.

Keywords: AB3-type alloy, composite electrode material, cycle stability

RESUMO

Com o objetivo de aprimorar as propriedades eletroquímicas da liga do tipo AB3 La0,78Mg0,22Ni2,67Mn0,11Al0,11Co0,52 como eletrodo negativo de bateria Ni-MH, seus compósitos relacionados La0,78Mg0,22Ni2,67Mn0,11Al0,11Co0,52-x wt.% M1Ni3,5Co0,6Mn0,4Al0,5 (x = 0, 10, 20, 30) foram preparados. A análise por difratometria de raios X (XRD) revelou que os compósitos consistem principalmente das fases LaNi5 e La2Ni7. Apesar da pequena diminuição na capacidade de descarga máxima, o rendimento do ciclo foi significativamente melhorado. Experimentos de polarização linear (LP), polarização anódica (AP) e descarga de potencial degrau revelaram um aumento inicial da cinética eletroquímica que na sequência diminui com o aumento de x.

Introduction

The discharge capacity of currently commercial AB5-type hydrogen-storage alloys has reached 320-340 mAh g-1 at room temperature. It seems to be difficult to further improve the capacity of the AB5-type alloys since the theoretical capacity of LaNi5 is 372 mAh g-1. Therefore, it is extremely important to develop new alloy types with higher capacity and longer life cycle for exacting the competition ability of Ni-MH batteries.

Recently, the new series of AB3-type R-Mg-Ni-based (where R is a rare earth or Y, Ca) hydrogen-storage alloys were extensively investigated as promising candidates for negative electrode material of Ni-MH rechargeable batteries due to their relatively high hydrogen-storage capacity (360-410 mAh g-1) and low production costs.1 Dong et al.2 reported that the La0.67Mg0.33Ni2.75Co0.25 alloy electrode exhibited a discharge capacity of 388 mAh g-1, which was much higher than AB5-type rare earth-based electrode alloys. However, these R-Mg-Ni-based alloys cannot meet the need of practical applications due to serious degradation of capacity during cyclic process.

Ball milling method is effective in the surface modification and improvement of the electrochemical properties of the hydrogen-storage alloys. For example, it was found that the maximum discharge capacity of Ml0.90Mg0.10Ni3.08Mn0.13Co0.63Al0.14-4 wt.% AB5 electrode reached 406 mAh g-1 and that the LaNi5 alloy could increase the discharge capacity of Ti-V-based BCC phase alloy.3,4 Chu et al.5 reported that the cyclic life of the La0.7Mg0.3Ni3.5 electrode was improved with addition of the TiNi0.56Co0.44 alloy.6

Herein, in order to obtain R-Mg-Ni-based hydrogen-storage alloys with excellent overall electrochemical properties, the La0.78Mg0.22Ni2.67Mn0.11Al0.11Co0.52-x wt.% M1Ni3.5Co0.6Mn0.4Al0.5 (where M1 denotes mischmetal, x = 0, 10, 20, 30) composites were prepared, which resulted in the appreciable elevation of the cycle stability. The microstructure and electrochemical properties of composites were also studied.

Experimental

Preparation and structural characterization

The La0.78Mg0.22Ni2.67Mn0.11Al0.11Co0.52 alloy (represented as A hereafter) and M1Ni3.5Co0.6Mn0.4Al0.5 alloy (represented as B hereafter, M1 consists of 37.7% La, 38.9% Ce, 6.3% Pr and 17.1% Nd) were prepared by induction melting under argon atmosphere and remelted four times for homogeneity. The purity of all the constituent metal elements was over 99.0%. The ingots were mechanically crushed and ground into 300 mesh powders for the preparation of the composite alloys. The A-x wt.% B (x = 0, 10, 20, 30) alloy and composites were prepared by ball milling. x wt.% (x = 0, 10, 20, 30) M1Ni3.5Co0.6Mn0.4Al0.5 alloy was mixed homogenously into La0.78Mg0.22Ni2.67Mn0.11Al0.11Co0.52 alloy, and ground by QM-ISP planetary ball miller under argon atmosphere for 1 h.

The structural data of the alloys was collected by powder X-ray diffraction (Rigaku D/max-2550, Cu Kα radiation, λ at 1.54178 Å).

Electrochemical measurements

The testing electrodes were prepared by cold pressing the alloy powder with carbonyl nickel powder in a weight ratio of 1:2 into a pellet of 15 mm in diameter under a pressure of 10 MPa and then pressed between two pieces of foam nickel. The charge/discharge studies were carried out by a Land 5.3 B Battery Test Instrument in a standard open tri-electrode electrolysis cell consisting of a working electrode (the MH pellet electrode for studying), a sintered Ni(OH)2/NiOOH counter electrode, and a Hg/HgO reference electrode immersed in the 6 mol dm-3 KOH electrolyte. The electrodes were charged for 5 h at a current density of 100 mA g-1, rested for 5 min and then discharged to the cut-off potential of -0.6 V vs. Hg/HgO reference electrode at a current density of 50 mA g-1.

PARSTAT 2273 electrochemical system was used for the kinetic tests. The anodic polarization was plotted at a rate of 0.5 mV s-1 from 0 to 600 mV (vs. open circuit potential) at 50% depth of discharge (DOD). The linear polarization testing was performed at a rate of 0.1 mV s-1 from -5 to 5 mV (vs. open circuit potential) at 50% DOD. Electrochemical impedance spectroscopy measurement was carried out in the frequency range of 1 × 103 to 1 × 10-4 Hz with ac amplitude of 5 mV at 50% DOD. The potential step discharge experiments were performed at 100% charge state, a +500 mV potential step was applied and the discharge time was 3600 s.

Results and Discussions

Phase structure

Figure 1 exhibits the XRD patterns for B, x = 0 and x = 10 alloys. It shows that the M1Ni3.5Co0.6Mn0.4Al0.5 alloy presents the single CaCu5-type hexagonal LaNi5 phase. The x = 0 and the x = 10 alloys consist of multiphase structures, the main phases are LaNi5 and La2Ni7 phases, indicating that the crystal structure of the alloy is not affected by appropriate milling.


Discharge capacity and cycle stability

Figure 2 illustrates the discharge capacity vs. cycle number for the A-x wt.% B (x = 0, 10, 20, 30) alloy electrodes. For comparison, the same relationship for the M1Ni3.5Co0.6Mn0.4Al0.5 alloy (represented as B) is also presented. The activation number (Na) and the maximum discharge capacity (Cmax) are listed in Table 1. The activation capability is closely related to the phase structure, surface characteristics, grain size and interstitial dimensions of the alloy. It is obvious that all the composite electrodes have excellent activation property and can be fully activated within three cycles, being mainly ascribed to their multiphase structures. The phase boundary cannot only act as a buffer for releasing strain energy but also provides good tunnels for the diffusion of hydrogen atoms, which was confirmed by literature.7


The oxidation and corrosion of the alloy composition in corrosive electrolyte is an extremely important factor that leads to the efficacy loss of the Ni-MH battery. In order to investigate the cycling durability of the alloy electrodes, the capacity retention (Sn) is defined and calculated by the following equation:

where Sn is the capacity retention at n cycles, Cnth is the discharge capacity at n cycles and Cmax is the maximum discharge capacity. The capacity retention rate at the 50th cycle (S50) of the alloy electrodes is shown in Table 1.

Cmax of the M1Ni3.5Co0.6Mn0.4Al0.5 alloy electrode is 315 mAh g-1, and the capacity retention (S50) is 88.6%. With a maximum discharge capacity of 371 mAh g-1 for x = 0 electrode, its cycle stability is rather poor. After 50 cycles, S50 is only 55.3%. As for the composite electrodes, Cmax decreases to 369 mAh g-1 (x = 10), 358 mAh g-1 (x = 20) and 356 mAh g-1 (x = 30), respectively. The hydrogen-storage capacity of the La2Ni7 phase is higher than that of LaNi5 phase,8 so the drop of the maximum discharge capacity can be attributed to the abundance decrease of the La2Ni7 phase with the addition of M1Ni3.5Co0.6Mn0.4Al0.5 alloy.

The cycle stability of the composite electrodes is significantly elevated, especially for x = 20 electrode, which exhibits the capacity retention rate (S50) of 76.5%. The improvement of the cycle performance is related to the change of the phase abundances of the alloys. The abundance of LaNi5 phase increases with the addition of M1Ni3.5Co0.6Mn0.4Al0.5 alloy, which is beneficial to the improvement of the cycle stability owing to the high Ni content of LaNi5 phase. Ni has excellent electrocatalytic activity and good corrosion resistance in an alkaline solution.9

Anodic polarization (AP)

Figure 3 shows the anodic polarization curves for the A-x wt.% B (x = 0, 10, 20, 30) alloy electrodes at 50% DOD. The anodic current density increases firstly with the overpotential increase, and then reaches a limiting value defined as the limiting current density (IL), which is mainly controlled by the hydrogen diffusion in the bulk of alloy. The hydrogen diffusion is higher as the limiting current density is larger. The values of IL obtained from Figure 3 are also summarized in Table 1. IL increases from 1204.0 mA g-1 (x = 0) to 1639.6 mA g-1 (x = 10), and then decreases to 1358.4 mA g-1 (x = 30). Liu et al.10 reported that the kinetics may be affected by the relative content of the two phases in the alloys. When the relative content of the LaNi5 and La2Ni7 phases come to a proper value, the electrochemical catalytic-activity of the alloy electrodes is enhanced due to the synergy of the two phases.11 Therefore, IL increases and then decreases with the x increase.


Linear polarization (LP)

Figure 4 illustrates the linear polarization curves for the A-x wt.% B (x = 0, 10, 20, 30) alloy electrodes. When the overpotential is changed within a small range, there is a good linear dependence between the current density and the overpotential. The exchange current density (I0) is an important kinetic parameter, which is used to evaluate the charge-transfer reaction rate on the electrode surface and can be calculated by the following equation:12


where Id, R, T, F and η denote the applied current density, gas constant, absolute temperature, Faraday constant and overpotential, respectively. As listed in Table 1, I0 increases after addition of M1Ni3.5Co0.6Mn0.4Al0.5 alloy, which is related to the change in the abundance of LaNi5 phase. The LaNi5 phase has higher electrocatalytic activity due to its higher Ni content, which cannot only work as a hydrogen absorber, but also as a catalytic site during the charge/discharge process.13 The exchange current density of the electrodes firstly increases from 191.6 mA g-1 (x = 0) to 282.9 mA g-1 (x = 10), and then decreases to 237.0 mA g-1 (x = 30) with the increase of x value. The variation of I0 is completely consistent with that of IL, which further confirms that the alloy electrode exhibits a comparatively better kinetics when the La2Ni7 and LaNi5 phases exist in a proper proportion.

Potentiostatic discharge

Figure 5 shows the curves of anodic current vs. discharge time curves for the A-x wt.% B (x = 0, 10, 20, 30) alloy electrodes at full charge state. It can be seen that after application of the overpotential, the current-time responses in the semilogarithmic plot (log ivs.t) can be divided into two domains. In the first region, the current density decreases rapidly due to the fast consumption of the hydrogen on the surface of the alloy particles, while the current density decreases slowly in a linear fashion in the second region. In the second region, the current is mainly controlled by the hydrogen diffusion velocity, and the hydrogen is supplied from the bulk alloy proportionally to the concentration gradient of hydrogen. According to a spherical diffusion model,14 the hydrogen diffusion coefficient (D) can be obtained from the slope of the linear section by the following equations:15


where D, C0, Cs, a, d, t and i denote the average coefficient of diffusion of hydrogen (cm2 s-1), the initial hydrogen concentration in the bulk of the alloy (which is considered uniform, mol cm-3), the hydrogen concentration on the surface of the alloy particles (which is considered uniform, mol cm-3), the alloy particle radius (cm), the density of the hydrogen-storage alloys (g cm-3), the discharge time (s) and the diffusion current density (A g-1), respectively. With an assumed average particle radius of a = 15 µm, the values of D calculated by the equation 4 are tabulated in Table 1. The hydrogen diffusion coefficient increases from 2.24 × 10-10 cm2 s-1 (x = 0) to 4.01 × 10-10 cm2 s-1 (x = 10), and then decreases to 2.33 × 10-10 cm2 s-1 (x = 30). It indicates that the hydrogen diffusivity firstly increases and then drops with the addition of AB5-type M1Ni3.5Co0.6Mn0.4Al0.5 alloy, in agreement with the limiting current density.

Electrochemical impedance spectra (EIS)

The charge-transfer resistance (Rct) of the alloy electrode can be determined by the electrochemical impedance spectra. Figure 6 illustrates the EIS curves for the A-x wt.% B (x = 0, 10, 20, 30) alloy electrodes at 50% DOD. It can be seen that each spectrum consists of a smaller semicircle in the high-frequency region and a larger semicircle in the low-frequency region followed by a straight line. According to the study of Kuriyama et al.,16 the high-frequency semicircle represents the contact resistance between the current collector and the alloy electrode, the low-frequency semicircle represents the charge-transfer resistance, and the straight line is namely Warburg impedance. It can be seen in Figure 6 that the contact impedance remains almost unchanged, but the radius of the large arc in the low-frequency region is in the order of x = 0 > x = 30 > x = 20 > x = 10, indicating that the charge-transfer resistance increases firstly and then decreases with increasing x, and the electrode with x = 10 exhibits the best electrochemical reaction rate.


High rate dischargeability (HRD)

Figure 7 presents the high rate dischargeability for the A-x wt.% B (x = 0, 10, 20, 30) alloy electrodes. The value of HRD is calculated according to the following equation:


where Cd and C50 represent the discharge capacities at the current densities of Id and 50 mA g-1, respectively. The values with the discharge current density of 600 mA g-1 are listed in Table 1. The M1Ni3.5Co0.6Mn0.4Al0.5 alloy presents the single LaNi5 phase, and the A-x wt.% B (x = 0, 10, 20, 30) alloys consist of multiphase structures. The phase boundary may decrease the lattice distortion and strain energy. Moreover, the phase boundary can provide good tunnels for the diffusion of hydrogen atoms, therefore, the kinetics of A-x wt.% B (x = 0, 10, 20, 30) is improved. For the composite electrodes, the HRD600 value increases from 84.2 % (x = 0) to 95.8% (x = 10), and then decreases to 88.8% (x = 30). It is well known that HRD is closely related to the kinetics of the alloy, which is mainly controlled by both the charge transfer resistance at the surface of alloy electrode and the hydrogen diffusion resistance inside the bulk of alloy.17 A proper ratio of La2Ni7 to LaNi5 phases is probably the reason why the electrode with x = 10 presents the best kinetic performance.

Figure 8 shows HRD600 as a function of I0 (obtained from linear polarization) for the A-x wt.% B (x = 0, 10, 20, 30) alloy electrodes. There is a linear relationship between HRD600 and I0. It is accepted that in case the charge-transfer reaction on the surface of the alloy electrodes is the rate-determining factor, a linear dependence of the HRD on the exchange current density would be obtained.18 Accordingly, it is concluded that the HRD is essentially controlled by electrochemical reaction of hydrogen on alloy surface at the discharge current density of 600 mA g-1.


The electrochemical kinetics is mainly determined by both the charge-transfer resistance on the surface of alloy electrode and hydrogen atom diffusion within the bulk of the alloy. The former can be characterized by I0 or Rct, while the latter can be characterized by the hydrogen diffusion coefficient or the limiting current density. In the present study, I0, IL and D firstly increase and then decrease with increasing x from 0 to 30. Thus, the kinetics of the investigated electrodes increases and then decreases with the increasing of M1Ni3.5Co0.6Mn0.4Al0.5 alloy content.

Conclusions

The La0.78Mg0.22Ni2.67Mn0.11Al0.11Co0.52-x wt.% M1Ni3.5Co0.6Mn0.4Al0.5 (x = 0, 10, 20, 30) composite alloys have a multi-phase structure of LaNi5 and La2Ni7 phases. With the addition of M1Ni3.5Co0.6Mn0.4Al0.5 alloy, the discharge capacity slightly drops, while the cycle stability is obviously enhanced. The kinetics of alloy electrodes firstly increases and then decreases with x increase. The linear relationship between HRD and I0 indicates that charge-transfer reaction on the surface should be the rate-determining factor for the electrochemical kinetics when the discharge current density is 600 mA g-1.

Acknowledgements

This work was financially supported by the Education Department and Scientific Foundation of Guangxi (201203YB088).

Submitted: February 17, 2013

Published online: July 12, 2013

  • 1. Zhang, Y. H.; Dong, X. P.; Li, B. W.; Ren, H. P.; Wu, Z. W.; Wang, X. L.; J. Alloys Compd 2008,465,422.
  • 2. Dong, Z. W.; Wu, Y. M.; Ma, L. Q.; Wang, L. D.; Shen, X. D.; Wang, L. M.; Int. J. Hydrogen Energy 2011,36,3050.
  • 3. Hu, L.; Han, S. M.; Li, J. H.; Zhu, X. L.; Li, Y.; Rare Met 2008,27,429.
  • 4. Yu, X. B.; Walker, G. S.; Grant, D. M.; Wu, Z.; Xia, B. J.; Shen, J.; Appl. Phys. Lett 2005,87,133121.
  • 5. Chu, H. L.; Qiu, S. J.; Sun, L. X.; Zhang, Y.; Xu, F.; Zhu, M.; Hu, W. Y.; Int. J. Hydrogen Energy 2008,33,755.
  • 6. Huang, H. X.; Huang, K. L.; Chen, D. Y.; Li, S. Q.; Zhuang, S. X.; J. Mater. Sci 2010,45,1123.
  • 7. Zhang, Y. H.; Li, B. W.; Ren, H. P.; Cai, Y.; Dong, X. P.; Wang, X. L.; Int. J. Hydrogen Energy 2007,32,4627.
  • 8. Oesterreicher, H.; Clinton, J.; Bittner, H.; Mater. Res. Bull 1976,11,1241.
  • 9. Wang, Y.; Wang, X.; Gao, X. P.; Shen, P. W.; Int. J. Hydrogen Energy 2007, 32,4180.
  • 10. Liu, Y. F.; Jin, Q. W.; Gao, M. X.; Zhu, Y. F.; Zhang, Z. H.; Pan, H. G.; Rare Met. Mater. Eng 2003, 32,942.
  • 11. Zhang, Z.; Han, S. M.; Li, Y.; Jing, T. F.; J. Alloys Compd 2007,431,208.
  • 12. Notten, P. H. L.; Hokkeling, P.; J. Electrochem. Soc 1991,138,1877.
  • 13. Li, Y.; Han, S. M.; Li, J. H.; Hu, L.; Electrochim. Acta 2007,52,5945.
  • 14. Nishima, T.; Ura, H.; Uchida, I.; J. Electrochem. Soc 1997,144,1273.
  • 15. Zheng, G.; Popov, B. N.; White, R. E.; J. Electrochem. Soc 1995,142,2695.
  • 16. Kuriyama, N.; Sakai, T.; Miyamura, H.; Uehara, I.; Ishikawa, H.; Iwasaki, T.; J. Alloys Compd 1993,202,183.
  • 17. Li, Y.; Han, S. M.; Li, J. H.; Zhu, X. L.; Hu, L.; J. Alloys Compd 2008,458,357.
  • 18. Iwakura, C.; Oura, T.; Inoue, H.; Matsuoka, M.; Electrochim. Acta 1996,41,117.
  • *
    e-mail:
  • Publication Dates

    • Publication in this collection
      23 Aug 2013
    • Date of issue
      Aug 2013

    History

    • Received
      17 Feb 2013
    • Accepted
      12 July 2013
    Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
    E-mail: office@jbcs.sbq.org.br