Next Article in Journal
Electrochemical Study of Na2Fe1−xMnxP2O7 (x = 0, 0.25, 0.5, 0.75, 1) as Cathode Material for Rechargeable Na-Ion Batteries
Next Article in Special Issue
A Technical Report of the Robust Affordable Next Generation Energy Storage System-BASF Program
Previous Article in Journal / Special Issue
Effects of Salt Additives to the KOH Electrolyte Used in Ni/MH Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structure, Hydrogen Storage, and Electrochemical Properties of Body-Centered-Cubic Ti40V30Cr15Mn13X2 Alloys (X = B, Si, Mn, Ni, Zr, Nb, Mo, and La)

1
Department of Chemical Engineering and Materials Science, Wayne State University, Detroit, MI 48202, USA
2
BASF/Battery Materials-Ovonic, 2983 Waterview Drive, Rochester Hills, MI 48309, USA
*
Author to whom correspondence should be addressed.
Batteries 2015, 1(1), 74-90; https://doi.org/10.3390/batteries1010074
Submission received: 2 November 2015 / Revised: 30 November 2015 / Accepted: 4 December 2015 / Published: 10 December 2015
(This article belongs to the Special Issue Nickel Metal Hydride Batteries)

Abstract

:
Structure, gaseous phase hydrogen storage, and electrochemical properties of a series of TiVCrMn-based body-centered-cubic (BCC) alloys with different partial substitutions for Mn with covalent elements (B and Si), transition metals (Ni, Zr, Nb, and Mo), and rare earth element (La) were investigated. Although the influences from substitutions on structure and gaseous phase storage properties were minor, influences on electrochemical discharge capacity were significant. The first cycle capacity ranged from 16 mAh·g−1 (Si-substituted) to 247 mAh·g−1 (Mo-substituted). Severe alloy passivation in 30% KOH electrolyte was observed, and an original capacity close to 500 mAh·g−1 could possibly be achieved by Mo-substituted alloy if a non-corrosive electrolyte was employed. Surface coating of Nafion to the Mo-substituted alloy was able to increase the first cycle capacity to 408 mAh·g−1, but the degradation rate in mAh·g−1·cycle−1 was still similar to that of standard testing. Electrochemical capacity was found to be closely related to BCC phase unit cell volume and width of the an extra small pressure plateau at around 0.3 MPa on the 30 °C pressure-concentration-temperature (PCT) desorption isotherm. Judging from its high electrochemical discharge capacity, Mo was the most beneficial substitution in BCC alloys for Ni/metal hydride (MH) battery application.

Graphical Abstract

1. Introduction

Among all metal hydride (MH) alloy families, body-centered-cubic (BCC) solid solution alloy has the highest reversible hydrogen storage at ambient temperature. Although its gaseous phase hydrogen storage capacity is very high (up to 4.0 wt%, equivalent to 1072 mAh·g−1 [1]), few electrochemical studies have been performed on the pure BCC phase MH alloy due to its strong metal-hydrogen bonding and low surface reaction activity [2,3,4,5]. Inoue and his coworker reported a TiV3.4Ni0.6 alloy achieving 360 mAh·g−1 at room temperature with a discharge rate of 50 mA·g−1 [3]. Mori and Iba improved both the capacity and cycle stability by adding Y, lanthanoids, Pd, or Pt into a TiCrVNi BCC alloy and reached 462 mAh·g−1 [4]. Yu and his coworker reported a Ti40V30Cr15Mn15 alloy with an initial capacity of 814 mAh·g−1 measured with a rate of 10 mA·g−1 at 80 °C; however, degradation was high due to surface cracking, preferential leaching of V into the KOH electrolyte, and formation of TiOx on the surface that further blocks electrochemical reaction [5]. One or more secondary phases, such as C14, C15, and/or B2, with a high grain boundary density was introduced to improve the absorption kinetics [6], facilitate formation due to its brittleness [7,8,9], and increase the surface catalytic activity [10,11,12,13,14,15] by enhancing the synergetic effect between the main and secondary phases. High phase boundary density also promotes the formation of coherent and catalytic interfaces between the BCC and secondary phases and, therefore, improves hydrogen absorption and desorption kinetics [16].
In this experiment, we focus on continuing the work on Ti40V30Cr15Mn15 alloy with an electrochemical study performed at room temperature and an examination of substitution effects from covalent elements, transition metals, and rare-earth elements on structure, gaseous phase, and electrochemical properties. The alloy formula in the current study can be summarized as Ti40V30Cr15Mn13X2, where X = B, Si, Mn, Ni, Zr, Nb, Mo, and La.

2. Experimental Setup

In this experiment, an arc melting technique was chosen for the sample preparation. The ingot size was about 12 g and the melting was performed in an Ar environment. A Varian Liberty 100 inductively-coupled plasma optical emission spectrometer (ICP-OES, Agilent Technologies, Santa Clara, CA, USA) was used to verify the chemical composition of the ingot comparing to the ratios in the raw materials. A Philips X’Pert Pro X-ray diffractometer (XRD, Philips, Amsterdam, The Netherlands) was used to study the microstructure, and a JEOL-JSM6320F scanning electron microscope (SEM, JEOL, Tokyo, Japan) with energy dispersive spectroscopy (EDS) capability was used to study the phase distribution and composition. Gaseous phase hydrogen storage characteristics for each sample were measured using a Suzuki-Shokan multi-channel pressure-concentration-temperature (PCT, Suzuki Shokan, Tokyo, Japan) system. A piece of ingot was freshly cleaved before putting in the PCT ample holder. PCT sample was first hydrided and dehydrided at 30 °C, followed by a 2 h, 400 °C degassing with a vacuum pump. PCT isotherms at 90 °C, 30 °C, and 60 °C were then measured with a 2 h, 400 °C degassing between measurements. Details of electrode preparations as well as measurement methods have been reported previously [17,18].

3. Results and Discussion

3.1. X-Ray Diffraction Structure Analysis

Eight alloys were prepared by arc melting, and their compositions were verified by ICP. XRD patterns of the alloys are shown in Figure 1. Three major peaks are detected in all alloys and belong to a BCC structure. Most of the peak intensity ratios are similar except for I(200)/I(110) in Alloy-Nb (alloy with partial replacement of Nb). Among all substitutions, elemental Nb and Mo are similar in size and both have a BCC structure; however, only Alloy-Nb has the unusually larger (200) peak. The reason for such phenomenon is not clear and requires further structural refinement analysis. In addition to the main phase, one or more secondary phases can be found in the XRD pattern of each alloy apart from Alloy-Nb.
Figure 1. XRD patterns using Cu-Kα as the radiation source for Ti40V30Cr15Mn13X2 alloys, where X = (a) B; (b) Si; (c) Mn; (d) Ni; (e) Zr; (f) Nb; (g) Mo; and (h) La. The vertical line is used to indicate shifts in the body-centered-cubic (BCC) peak (110) with respect to that in Ti40V30Cr15Mn15 alloy.
Figure 1. XRD patterns using Cu-Kα as the radiation source for Ti40V30Cr15Mn13X2 alloys, where X = (a) B; (b) Si; (c) Mn; (d) Ni; (e) Zr; (f) Nb; (g) Mo; and (h) La. The vertical line is used to indicate shifts in the body-centered-cubic (BCC) peak (110) with respect to that in Ti40V30Cr15Mn15 alloy.
Batteries 01 00074 g001
Rietveld refinement results from the XRD analysis are summarized in Table 1. Lattice parameter a of the BCC phase ranges from 3.0679 Å to 3.0839 Å, which is larger than the optimized value of 3.042 Å corresponding to the maximized hydrogen storage capacity [19], leaving room for potential improvement in storage capacity for future studies.
Table 1. Summary of X-ray diffraction (XRD) results from alloys Ti40V30Cr15Mn13X2.
Table 1. Summary of X-ray diffraction (XRD) results from alloys Ti40V30Cr15Mn13X2.
Xa of BCC phaseBCC phase abundanceSecondary phasea of secondary phasec of secondary phaseSecondary phase abundance
Åwt%ÅÅwt%
B3.070398.2TiO24.1761-1.8
Si3.067996.9TiO24.1472-3.1
Mn3.068798.9TiO24.1687-1.1
Ni3.064999.7TiO24.1567-0.3
Zr3.083998.2C144.98958.17901.8
Nb3.079099.8TiO24.1743-0.2
Mo3.077499.6TiO24.1706-0.4
La3.069398.3La2O311.302-1.7
The BCC lattice constant a is plotted against the atomic radius of the substituting element in Figure 2. Alloys substituted with transition metals show a linear relationship between the lattice constant and atomic radius (represented by the straight line in Figure 2). B and Si, with smaller atomic radii and larger electronegativity, do not shrink the BCC unit cell volume, which is possibly due to the electrons transferred from neighboring atoms and increases in the radius. Similar behavior has been found in the increase in the radius of B and Si in the Laves phase intermetallic compound [20]. La, with the largest atomic radius, does not change the BCC lattice constant, indicating La does not enter the BCC phase. According to the results of Rietveld refinement, the BCC phase abundances in all alloys are greater than 96.9%. TiO2 is the dominating secondary phase, with some exceptions, seen in Alloy-Zr (C14), Alloy-Ni (TiNi observed in the SEM/EDS analysis as discussed in the next section), and Alloy-La (La2O3). Zr is known to promote the Laves phase in BCC-predominant alloys [21,22,23,24,25,26]. TiNi is a common phase seen in C14-predominant alloys with high concentration of Ni [27,28,29,30]. The main diffraction peak of TiNi overlaps with BCC (110) and, therefore, is indistinguishable in the XRD pattern (Figure 1). La2O3 was formed since La is too large to be included in the BCC phase, agreeing with the immiscibility shown in the La-V binary phase diagram [31].
Figure 2. BCC lattice constant a vs. atomic radius of the partial substitution element X in Ti40V30Cr15Mn13X2 alloys. There is a linear dependence when X is a transition metal. Addition of the largest La does not change the BCC lattice constant because La does not dissolve in the BCC phase and, instead, forms La2O3 secondary phase. Adding relatively small B and Si with higher electronegativity do not shrink the BCC unit cell because these atoms attract electrons from neighboring metallic atoms.
Figure 2. BCC lattice constant a vs. atomic radius of the partial substitution element X in Ti40V30Cr15Mn13X2 alloys. There is a linear dependence when X is a transition metal. Addition of the largest La does not change the BCC lattice constant because La does not dissolve in the BCC phase and, instead, forms La2O3 secondary phase. Adding relatively small B and Si with higher electronegativity do not shrink the BCC unit cell because these atoms attract electrons from neighboring metallic atoms.
Batteries 01 00074 g002

3.2. Scanning Electron Microscope/Energy Dispersive Spectroscopy Microstructure Analysis

Microstructures of the alloys were studied using SEM. The back-scattering electron images (BEI) are presented in Figure 3.
Figure 3. Scanning electron microscope (SEM) back-scattering electron images (BEI) for Ti40V30Cr15Mn13X2 alloys, where X = (a) B; (b) Si; (c) Mn; (d) Ni; (e) Zr; (f) Nb; (g) Mo; and (h) La. Chemical compositions in the numbered areas measured by energy dispersive spectroscopy (EDS) are listed in Table 2.
Figure 3. Scanning electron microscope (SEM) back-scattering electron images (BEI) for Ti40V30Cr15Mn13X2 alloys, where X = (a) B; (b) Si; (c) Mn; (d) Ni; (e) Zr; (f) Nb; (g) Mo; and (h) La. Chemical compositions in the numbered areas measured by energy dispersive spectroscopy (EDS) are listed in Table 2.
Batteries 01 00074 g003
EDS, although a semi-quantitative analysis, was used to study the chemical compositions of several spots with different contrasts identified numerically in the micrographs (Figure 3), and the results are summarized in Table 2 due to convenience and availability.
Table 2. Summary of energy dispersive spectroscopy (EDS) results. All compositions are in at%. Compositions of BCC phase are in bold.
Table 2. Summary of energy dispersive spectroscopy (EDS) results. All compositions are in at%. Compositions of BCC phase are in bold.
LocationTiVCrMnXPhase
Figure 3a-141.129.316.712.90.0BCC
Figure 3a-242.228.016.213.70.0BCC
Figure 3a-342.228.216.313.30.0BCC
Figure 3a-459.523.99.37.30.0Oxide
Figure 3a-564.422.57.65.50.0Oxide
Figure 3b-138.731.915.812.11.6BCC
Figure 3b-238.432.215.812.11.5BCC
Figure 3b-349.715.412.015.47.5Oxide
Figure 3b-455.815.310.512.65.8Oxide
Figure 3b-555.617.410.211.55.2Oxide
Figure 3c-138.830.115.815.30.0BCC
Figure 3c-238.729.915.915.50.0BCC
Figure 3c-341.626.414.917.00.0BCC
Figure 3c-443.326.214.615.90.0BCC
Figure 3c-542.125.914.917.10.0BCC
Figure 3d-136.933.816.611.41.3BCC
Figure 3d-238.631.916.112.01.3BCC
Figure 3d-342.628.214.712.42.1BCC
Figure 3d-451.116.810.012.69.4TiNi
Figure 3d-557.99.15.911.216.0TiNi
Figure 3e-139.731.315.712.11.1BCC
Figure 3e-243.119.913.315.97.7C14/ZrxNiy
Figure 3e-332.919.316.720.410.6C14/ZrxNiy
Figure 3e-431.517.916.421.512.7C14/ZrxNiy
Figure 3e-539.017.314.818.410.5C14/ZrxNiy
Figure 3f-138.732.815.111.32.1BCC
Figure 3f-239.631.614.911.82.1BCC
Figure 3f-339.631.814.911.62.1BCC
Figure 3f-443.327.814.113.01.9BCC
Figure 3f-545.125.913.913.31.9BCC
Figure 3g-140.529.916.112.01.5BCC
Figure 3g-241.429.016.112.31.3BCC
Figure 3g-342.127.915.912.91.2BCC
Figure 3g-444.925.515.313.60.7BCC
Figure 3g-546.723.614.914.30.6BCC
Figure 3h-145.526.113.914.50.0BCC
Figure 3h-243.628.214.513.60.0BCC
Figure 3h-334.024.210.59.122.2La2O3
Figure 3h-416.110.41.05.666.9La2O3
Figure 3h-55.02.91.00.091.2La
Except for B and La, the substituting element is present in the BCC phase, ranging in content from 1.1 at% to 2.1 at%. EDS, although a semi-quantitative analysis, was used to study the chemical compositions of several spots with different contrasts identified numerically in the micrographs (Figure 3), and the results are summarized in Table 2 due to convenience and availability. Except for B and La, the substituting element is present in the BCC phase, ranging in content from 1.1 at% to 2.1 at%. The EDS system used for the current study cannot quantify the amount of lighter elements, such as B. According to the XRD and SEM-BEI analyses, the B-predominating phase does not exist; therefore, it is assumed that B is distributed in the BCC phase. Area with darker contrast in Alloy-B and Alloy-Si (Figure 3a-4,3a-5,3b-3,3b-4,3b-5) are small TiO2 particles embedded in the BCC matrix. Alloy-Mn, Alloy-Nb, and Alloy-Mo are uniform in composition. In Alloy-Ni, the TiNi secondary phase was found (Figure 3d-4,3d-5). The C14/ZrxNiy phase in Alloy-Zr distributes inter-granularly since the BCC phase solidifies first and pushes Zr into the C14 phase. Average electron density (e/a) of the secondary phase in Alloy-Zr is 5.06, which is below the C14/C15 threshold [32,33] and is, therefore, another piece of evidence that the secondary phase is C14 rather than C15 in addition to the findings in XRD analysis. B/A in this C14 phase is in the range of 0.97 to 1.3, which is way too low for an AB2 with a perfect B/A of 2.0. Since there is no major shift in XRD peaks of C14, these areas are not hypo-stoichiometric AB2. Therefore, other ZrxNiy secondary phase must also co-exist in this C14 phase, as in the case of AB2-predominated alloys [34,35]. Since the B/A ratios of the components of ZrxNiy (Zr7Ni10, Zr9Ni11, TiNi, and ZrNi) are all below 2.0, their existence will lower the B/A ratio in this region. In Alloy-La, La does not precipitate in the main BCC phase. La either forms a large metallic inclusion (Figure 3h-5) or an oxide suspended uniformly in the BCC matrix (Figure 3h-3)/near the edge of La metallic clusters (Figure 3h-4). The zero-solubility of La in BCC explains why the addition of La does not change the BCC lattice constant (Figure 2).

3.3. Gaseous Phase Study

PCT analysis was used to characterize the gaseous phase hydrogen storage properties of alloys in this study. The chamber containing the sample was filled with 7 MPa of hydrogen at 30 °C, and then the absorption amount was calculated, followed by a PCT desorption measurement at the same temperature. The sample was degassed at 400 °C for 2 h with a mechanical vacuum pump, and then a full 60 °C absorption-desorption PCT was measured. The sample was degassed at 400 °C for 2 h again, followed by a 90 °C PCT measurement. Finally, it was degassed at 400 °C for 2 h, and a last 30 °C PCT measurement was conducted. Absorption and desorption isotherms measured at 30 °C, 60 °C, and 90 °C together with the initial 30 °C desorption isotherm are shown in Figure 4. Information obtained from the PCT study is summarized in Table 3. Most of the alloys show similar gaseous phase properties. Except for Alloy-Zr (3.12 wt%), the pristine alloys have similar maximum storage capacities in the range of 3.30 wt% to 3.55 wt%. A storage capacity of 3.50 wt% can be translated into an electrochemical discharge capacity 938 mAh·g−1 based on 1 wt% of hydrogen storage is equivalent to 268 mAh·g−1. Maximum storage capacities measured at 30 °C and 60 °C after 400 °C degassing show the following trend: substitution of B > Mo ~ Nb ~ La > Ni ~ Mn ~ Si > Zr, which demonstrate very weak correlations to the BCC unit cell volume (correlation factors R2 = 0.18 and 0.22 for storage capacities at 30 °C and 60 °C, respectively, indicating larger BCC unit cell corresponds to lower capacity) that were opposite to was expected. In general, reversibility of these alloys (ratio of revisable capacity down to 0.001 MPa and maximum capacity) is much worse than that of AB2 or AB5 MH alloy because of the fact the first pressure plateau between BCC and body-center-tetragonal (BCT) phases is too low to be observed with our PCT apparatus. While Alloy-B shows the best reversibility at 30 °C, Alloy-Mn and Alloy-Ni have better reversibility at 60 °C than others. Average reversible 30 °C storage capacity is about 0.7 wt%, which is equivalent to an electrochemical discharge capacity of 188 mAh·g−1. The 90 °C desorption plateau pressure of Alloy-Ni is much higher than those of other alloys. Hysteresis of the PCT isotherm is defined as ln(Pa/Pd), where Pa and Pd are the absorption and desorption equilibrium pressures at 2.0 wt% hydrogen storage, respectively. In this series of alloys, only PCT hysteresis at 90 °C can be measured. All substitutions show similar or slightly lower hysteresis, except for Si. PCT hysteresis is mainly from the energy required to elastically deform the lattice near the metal/MH interface during hydrogenation. Most substitutions increase the chemical disorder and reduce the PCT hysteresis. Nb has the same BCC crystal structure as Mn, therefore its effects on the degree of disorder and PCT hysteresis are limited. Adding Si with covalent bonding may stiffen the lattice, requiring higher energy to expand the MH phase in the host metal.
Due to the low desorption plateau pressure in these alloys, the regular thermodynamic calculation cannot be performed. Instead, the absorption equilibrium pressures at 2.0 wt% hydrogen storage at 60 °C and 90 °C were used to estimate the changes in enthalpy (ΔH) and entropy (ΔS) by the equation:
ΔG = ΔHTΔS = RTlnP
where R is the ideal gas constant and T is the absolute temperature. Results of these calculations are listed in Table 3. Compared to the base Alloy-Mn, all substitutions decrease ΔH except for Zr, which indicates that Zr decreases the hydride stability. In the case of Alloy-Zr, addition of the C14 phase in the alloy facilitates hydrogen absorption through the synergetic effect between the storage and catalytic phases [36] and destabilizes the hydride. ΔS, usually calculated with the desorption isotherm, is an indication of how far the MH system is from a perfect and ordered situation. The theoretical value of ΔS is the entropy of hydrogen gas, which is close to −130 J·mol−1·K−1 [37]. In our calculation, all substitutions decrease ΔS to below −135 J·mol−1·K−1 except for Zr, which is an indication that a more ordered MH system was formed. Alloy-Zr shows a relatively high value of ΔS, suggesting a more disordered MH system was formed, possibly due to the interaction between the main BCC and C14 secondary phases.
One interesting feature in the PCT isotherms caught one of the authors’ (Ouchi) attention: several alloys—Alloy-B, Alloy-Zr, Alloy-Nb, and Alloy-Mo—show a small plateau near 0.3 MPa on the 30 °C desorption curve while others do not. This plateau, although very small (about 0.10 wt% to 0.16 wt%), is at a pressure just above one atmosphere (0.1 MPa) and can be from a catalytic phase that has not been reported previously. The importance of this phase with respect to electrochemical performance will be discussed in the discussion section of this paper.
Table 3. Summary of gaseous phase hydrogen storage properties. PCT: pressure-concentration-temperature.
Table 3. Summary of gaseous phase hydrogen storage properties. PCT: pressure-concentration-temperature.
XInitial maximum capacity30 °C maximum capacity30 °C reversible capacity60 °C maximum capacity60 °C reversible capacity90 °C desorption pressure @2.0 wt%90 °C hysteresis @2.0 wt%−∆H−∆SPCT plateau @0.3 MPa
wt%wt%wt%wt%wt%MPakJ·mol−1J·mol−1·K−1
B3.483.381.733.431.000.0111.767181Yes
Si3.303.080.523.081.120.0112.556156No
Mn3.473.110.493.021.400.0132.037107No
Ni3.393.160.633.161.560.0281.858165No
Zr3.122.760.532.591.090.0161.92263Yes
Nb3.483.250.593.190.860.0112.064176Yes
Mo3.423.240.563.291.000.0121.866178Yes
La3.553.190.493.190.740.0091.667178No
Figure 4. PCT isotherms measured at 30 °C (both before and after 400 °C degassing), 60 °C, and 90 °C for Ti40V30Cr15Mn13X2 alloys, where X = (a) B; (b) Si; (c) Mn; (d) Ni; (e) Zr; (f) Nb; (g) Mo; and (h) La. Open and solid symbols are for absorption and desorption curves, respectively.
Figure 4. PCT isotherms measured at 30 °C (both before and after 400 °C degassing), 60 °C, and 90 °C for Ti40V30Cr15Mn13X2 alloys, where X = (a) B; (b) Si; (c) Mn; (d) Ni; (e) Zr; (f) Nb; (g) Mo; and (h) La. Open and solid symbols are for absorption and desorption curves, respectively.
Batteries 01 00074 g004aBatteries 01 00074 g004b

3.4. Electrochemical Measurement in 30% KOH

In a flooded half-cell, the electrochemical properties of MH alloys in this study were studied. Electrodes were made with powder after the PCT measurement and degassed four times at 400 °C for 2 h. No alkaline pretreatment was applied before the half-cell measurement. The charge condition was 10 h with a current density of 50 mA·g−1 and discharged at the same rate initially and followed by two pulls at 12 mA·g−1 and 4 mA·g−1 with a cut-off voltage at 0.9 V versus the counter electrode. The 500 mAh total charge input was based on the maximum reversible gaseous phase capacity being 1.73% (429 mAh·g−1). The charge and discharge voltage curves for Alloy-Mo are shown in Figure 5. The high resistance through the poor-conducting TiO2 surface resulted from the highly-corrosive 30% KOH electrolyte may cause the large charge and discharge overpotentials [5].
Figure 5. The first cycle charge and discharge voltage profiles for Ti40V30Cr15Mn13Mo2.
Figure 5. The first cycle charge and discharge voltage profiles for Ti40V30Cr15Mn13Mo2.
Batteries 01 00074 g005
Capacities totaled at 50, 12, and 4 mA·g−1 are listed in Table 4.
Table 4. Summary of electrochemical hydrogen storage properties.
Table 4. Summary of electrochemical hydrogen storage properties.
X1st cycle capacity @ 50 mA·g−11st cycle capacity @ 12 mA·g−11st cycle capacity @ 4 mA·g−1
mAh·g−1mAh·g−1mAh·g−1
B81163179
Si81416
Mn122024
Ni334761
Zr64130144
Nb416879
Mo152234247
La233941
About 50% of the capacity was obtained at the highest rate. The total capacity (totaled at 4 mA·g−1) for the first six cycles of each alloy is plotted in Figure 6a. All substitutions for Mn, except for Si, show improvement in the first cycle capacity. The first cycle capacity demonstrates the trend of substitution of Mo > B > Zr > Nb > Ni > La > Mn > Si. Alloy-Si has the highest hysteresis, indicating its proneness to pulverization [38] and, thus, poor electrochemical performance. Alloy-Mo shows the highest discharge capacity at 247 mAh·g−1. Partial Mo substitution in Ti-Cr MH alloy was reported previously, and it stabilized the BCC structure and improved the gaseous phase properties [39]. As seen in Figure 6a, capacity drops to almost nothing at the second cycle due mainly to the highly corrosive nature of 30% KOH electrolyte. The large amount of over-charge may also contribute to the severe capacity degradation. The original capacity of Alloy-Mo without corrosion from KOH can be extrapolated and is about double that obtained from the first cycle; in other words, electrochemical capacity close to 500 mAh·g−1 is possible if corrosion and passivation can be prevented from the use of non-corrosive electrolyte. In cycles two to six, Alloy-Ni with the TiNi phase shows the highest discharge capacity since the TiNi phase protects some portions of the bulk from being completely corroded. TiNi was found to increase the cycle stability of Laves phase MH alloys in a previous study [32].
Figure 6. (a) Discharge capacities measured at 4 mA·g−1 in 30% KOH for Ti40V30Cr15Mn13X2 alloys and (b) discharge capacities measured at 4 mA·g−1 for Ti40V30Cr15Mn13Mo2 in modified electrolytes and with Nafion treatment.
Figure 6. (a) Discharge capacities measured at 4 mA·g−1 in 30% KOH for Ti40V30Cr15Mn13X2 alloys and (b) discharge capacities measured at 4 mA·g−1 for Ti40V30Cr15Mn13Mo2 in modified electrolytes and with Nafion treatment.
Batteries 01 00074 g006
A SEM micrograph taken from the surface of Alloy-Mo after six electrochemical cycles reveals severe pulverization (Figure 7a), and the EDS spectrum taken from the surface shows no trace of oxide (Figure 7b). It may be too thin and compact to be detected in the current study, but surface TiO2 formed after cycling was reported previously in a BCC alloy [5] and can hinder the surface electrochemical reaction completely.
Figure 7. (a) SEM surface micrograph exhibiting the severe pulverization and (b) EDS spectrum showing negligible amount of oxygen on the surface after six electrochemical cycles with Ti40V30Cr15Mn13Mo2.
Figure 7. (a) SEM surface micrograph exhibiting the severe pulverization and (b) EDS spectrum showing negligible amount of oxygen on the surface after six electrochemical cycles with Ti40V30Cr15Mn13Mo2.
Batteries 01 00074 g007
In order to improve the cycle stability, preferential leaching, and surface passivation issues need to be addressed. Therefore, the effects of Y2O3 and ZnO additions in electrolyte and Nafion surface coating on electrode are investigated. It was found that by mixing Y2O3 powder to AB5 alloy powder, dissolution of the alloy’s constituent elements and formation of rare earth hydroxides were suppressed by an yttrium protective film on the alloy surface [40,41]. A small amount of Y2O3 dissolved in the alkaline electrolyte, and the yttrium complex ions were then adsorbed on or chemically bound to the surface of the alloy powder. ZnO, an amphoteric species, is nearly insoluble in water but soluble in acid and base. ZnO’s effect as an electrolyte additive on electrochemical degradation performance will be interesting to observe. One drawback of Y2O3 addition is that the yttrium protective layer lowers the catalytic activity of the alloy surface and, consequently, reduces the surface charge transfer reaction [40]. In order to possibly solve such an issue, Nafion, which is hydrogen-permeable with great chemical stability, was also adopted in the current study. Nafion applied as a protective coating increased the first cycle capacity by up to 75% in an MgNi-based alloy [42]. For the electrolyte modification study, two electrolytes were made by adding 2 wt% Y2O3 and 2 wt% ZnO (both based on the amount of KOH) into 30% KOH and used in the half-cell measurements with Alloy-Mo. For the surface modification study, Alloy-Mo electrode was coated with Nafion (perfluorosulfonic acid-PTFT copolymer, 5 wt% in water and 1-propanol) by dipping it for 2 min in ethanol diluted Nafion solution, where the ethanol to Nafion solution ratio was 5:1 by weight, and then heat treating for 2 h at 120 °C under argon atmosphere. The total capacity of each treatment compared to that of standard 30% KOH is plotted in Figure 6b. The first cycle capacity is improved with the additions of Y2O3 and ZnO from 247 mAh·g−1 to 320 mAh·g−1, but it still drops dramatically at the second cycle due to pulverization. The additives were incorporated differently than they were the previous study, where pasted electrode was made using the mixture of additive, alloy powder, and water [40]. The wet method can perhaps provide a better distribution of additive and more complete protection for the alloy. Among all treatments, Nafion coating on electrode is the most effective in improving the first cycle capacity and achieves 408 mAh·g−1, but degradation after the first cycle is still severe. After cycling, alloy particle pulverization (as seen from the SEM micrograph in Figure 7a) creates new surfaces that are not protected by the Nafion coating. Combination of surface coating (original robust protection) and electrolyte modification (continuous protection throughout cycling) may be advantageous for further corrosion and passivation inhibitions.

3.5. Discussion

To further study the correlations between electrochemical discharge capacity and various properties, correlation factors R2 from linear regression are calculated and listed in Table 5. All correlations with gaseous phase properties are insignificant.
Table 5. Correlation factor (R2) between electrochemical discharge capacity and various gaseous phase hydrogen storage properties. A R2 value closer to one indicates better linear correlation between the two variables.
Table 5. Correlation factor (R2) between electrochemical discharge capacity and various gaseous phase hydrogen storage properties. A R2 value closer to one indicates better linear correlation between the two variables.
Correlation factorInitial maximum capacity30 °C maximum capacity30 °C reversible capacity60 °C maximum capacity60 °C reversible capacity90 °C desorption pressure @2 wt%90 °C hysteresis @2 wt%ΔHBCC lattice constant aPlateau width
R2 =0.020.030.170.040.060.010.190.020.290.70
Correlation with BCC lattice constant is only marginally significant (R2 = 0.29) and plotted in Figure 8a, showing the enlargement of unit cell results in an increase in electrochemical capacity but not in a strictly linear relationship.
One significant correlation can be established between electrochemical capacity and occurrence of a small plateau near 0.3 MPa on the 30 °C desorption isotherm. Among all alloys, Alloy-B, Alloy-Zr, Alloy-Nb, and Alloy-Mo have this plateau at around 0.3 MPa and also the highest electrochemical discharge capacity. This plateau is from an intermediate hydride phase that can be catalytic and improve the electrochemical reaction. Electrochemical discharge capacity is plotted against width of the plateau at around 0.3 MPa (0.16 wt% for Alloy-B, 0.09 wt% for Alloy-Zr, 0.08 wt% Alloy-Nb, and 0.10 wt% for Alloy-Mo, which is defined as the width of concentration difference between the extrapolations from two neighboring isotherms) in Figure 8b and shows a strong correlation with R2 = 0.70. Furthermore, correlation with transition metal substitution is even stronger as seen from the straight line connecting points from Alloy-Mo, Alloy-Zr, and Alloy-Nb. Electrochemical discharge capacity is mainly dominated by the catalytic phase formed during hydrogenation.
Figure 8. (a) Discharge capacities measured at 4 mA·g−1 vs. BCC lattice constant and (b) vs. width of 0.3 MPa pressure plateau. The straight line in Figure 8b is to illustrate the linear correlation between capacity and plateau width at 0.3 MPa of the transition metal substitution.
Figure 8. (a) Discharge capacities measured at 4 mA·g−1 vs. BCC lattice constant and (b) vs. width of 0.3 MPa pressure plateau. The straight line in Figure 8b is to illustrate the linear correlation between capacity and plateau width at 0.3 MPa of the transition metal substitution.
Batteries 01 00074 g008

4. Conclusions

Various hydrogen storage properties in gaseous phase and in electrochemistry of a series of TiVCrMn-based BCC alloys with different partial substitutions for Mn with B, Si, Ni, Zr, Nb, Mo, and La were investigated. All substitutions went into the BCC phase except for La. While Ni promoted the formation of TiNi secondary phase that provided better cycle stability, Zr promoted the formation of C14 secondary phase, which did not affect any of the properties significantly due to its small abundance. Correlations between gaseous phase properties and lattice constant were not clear.
A newly discovered catalytic phase formed during hydrogenation was found to be very critical for the electrochemical discharge capacity performance. This phase enabled the electrochemical application of BCC-only alloys without contributions from secondary phases. The highest discharge capacity of 247 mAh·g−1 was obtained from Ti40V30Cr15Mn13Mo2 alloy with both a catalytic hydride phase at around 0.3 MPa and an enlarged BCC unit cell. Further improvement in electrochemical capacity of this alloy reached as high as 408 mAh·g−1 when a protective Nafion coating was applied on the electrode. Substitutions of B, Nb, and Zr also improved the electrochemical capacity but at a lesser degree.

Acknowledgments

This work is financially supported by Advanced Research Projects Agency-Energy (ARPA-E) under the robust affordable next generation EV-storage (RANGE) program (DE-AR0000386).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Young, K.; Fetcenko, M.A.; Ouchi, T.; Im, J.; Ovshinsky, S.R.; Li, F.; Reinhout, M. Hydrogen Storage Materials Having Excellent Kinetics, Capacity, and Cycle Stability. U.S. Patent 7,344,676, 18 March 2008. [Google Scholar]
  2. Lee, H.; Chourashiya, M.G.; Park, C.; Park, C. Hydrogen storage and electrochemical properties of the Ti0.32Cr0.43–xyV0.25FexMny (x = 0–0.055, y = 0–0.080) alloys and their composites with MmNi3.99Al0.29Mn0.3Co0.6 alloy. J. Alloys Compd. 2013, 566, 37–42. [Google Scholar] [CrossRef]
  3. Inoue, H.; Arai, S.; Iwakura, C. Crystallographic and electrochemical characterization of TiV4–xNix alloys for nickel-metal hydride batteries. Electrochim. Acta 1996, 41, 937–939. [Google Scholar] [CrossRef]
  4. Mori, T.; Iba, H. Hydrogen-absorbing Alloy and Hydrogen-absorbing Alloy Electrode. U.S. Patent 6,338,764, 15 January 2002. [Google Scholar]
  5. Yu, X.B.; Wu, Z.; Xia, B.J.; Xu, N.X. A Ti-V-based bcc phase alloy for use as metal hydride electrode with high discharge capacity. J. Chem. Phys. 2004, 121, 987–990. [Google Scholar] [CrossRef] [PubMed]
  6. Chen, N.; Li, R.; Zhu, Y.; Liu, Y.; Pan, H. Electrochemical hydrogenation and dehydrogenation mechanisms of the Ti-V base multiphase hydrogen storage electrode alloy. Acta Metal. Sin. 2004, 40, 1200–1204. [Google Scholar]
  7. Iba, H.; Akiba, E. The relation between microstructure and hydrogen absorbing property in Laves phase-solid solution multiphase alloys. J. Alloys Compd. 1995, 231, 508–512. [Google Scholar] [CrossRef]
  8. Rönnebro, E.; Noréus, D.; Sakai, T.; Tsukahara, M. Structural studies of a new Laves phase alloy (Hf,Ti)(Ni,V)2 and its very stable hydride. J. Alloys Compd. 1995, 231, 90–94. [Google Scholar] [CrossRef]
  9. Tsukahara, M.; Takahashi, K.; Mishima, T.; Isomura, A.; Sakai, T. V-based solid solution alloys with Laves phase network: Hydrogen absorption properties and microstructure. J. Alloys Compd. 1996, 236, 151–155. [Google Scholar] [CrossRef]
  10. Qiu, S.; Chu, H.; Zhang, Y.; Sun, D.; Song, X.; Sun, L.; Xu, F. Electrochemical kinetics and its temperature dependence behaviors of Ti0.17Zr0.08V0.35Cr0.10Ni0.30 alloy electrode. J. Alloys Compd. 2009, 471, 453–456. [Google Scholar] [CrossRef]
  11. Young, K.; Nei, J.; Wong, D.F.; Wang, L. Structural, hydrogen storage, and electrochemical properties of Laves phase-related body-centered-cubic solid solution metal hydride alloys. Int. J. Hydrog. Energy 2014, 39, 21489–21499. [Google Scholar] [CrossRef]
  12. Young, K.; Wong, D.F.; Wang, L. Effect of Ti/Cr content on the microstructures and hydrogen storage properties of Laves phase-related body-centered-cubic solid solution alloys. J. Alloys Compd. 2015, 622, 885–893. [Google Scholar] [CrossRef]
  13. Young, K.; Ouchi, T.; Nei, J.; Meng, T. Effects of Cr, Zr, V, Mn, Fe, and Co to the hydride properties of Laves phase-related body-centered-cubic solid solution alloys. J. Power Sources 2015, 281, 164–172. [Google Scholar] [CrossRef]
  14. Yan, Y.; Chen, Y.; Liang, H.; Zhou, X.; Wu, C.; Tao, M.; Pang, L. Hydrogen storage properties of V–Ti–Cr–Fe alloys. J. Alloys Compd. 2008, 454, 427–431. [Google Scholar] [CrossRef]
  15. Huot, J.; Akiba, E.; Ogura, T.; Ishido, Y. Crystal structure, phase abundance and electrode performance of Laves phase compounds (Zr, A) V0.5Ni1.1Mn0.2Fe0.2 (A = Ti, Nb or Hf). J. Alloys Compd. 1995, 218, 101–109. [Google Scholar] [CrossRef]
  16. Iba, H.; Akiba, E. Hydrogen absorption and modulated structure in Ti–V–Mn alloys. J. Alloys Compd. 1997, 253–254, 21–24. [Google Scholar] [CrossRef]
  17. Young, K.; Fetcenko, M.A.; Li, F.; Ouchi, T. Structural, thermodynamic, and electrochemical properties of TixZr1x(VNiCrMnCoAl)2 C14 Laves phase alloys. J. Alloys Compd. 2008, 464, 238–247. [Google Scholar] [CrossRef]
  18. Young, K.; Fetcenko, M.A.; Koch, J.; Morii, K.; Shimizu, T. Studies of Sn, Co, Al, and Fe additives in C14/C15 Laves alloys for NiMH battery application by orthogonal arrays. J. Alloys Compd. 2009, 486, 559–569. [Google Scholar] [CrossRef]
  19. Yoshida, M.; Akiba, E. Hydrogen absorbing-desorbing properties and crystal structure of the Zr–Ti–Ni–Mn–V AB2 Laves phase alloys. J. Alloys Compd. 1995, 224, 121–126. [Google Scholar] [CrossRef]
  20. Gakkai, N.K. Hi Kagaku Ryouronteki Kinzoku Kagobutu; Maruzen: Tokyo, Japan, 1975; p. 296. (In Japanese) [Google Scholar]
  21. Huot, H.; Akiba, E.; Ishido, Y. Crystal structure of multiphase alloys (Zr,Ti)(Mn,V)2. J. Alloys Compd. 1995, 231, 85–89. [Google Scholar] [CrossRef]
  22. Kuriiwa, T.; Tamura, T.; Amemiya, T.; Fuda, T.; Kamegawa, A.; Takamura, H.; Okada, M. New V-based alloys with high protium absorption and desorption capacity. J. Alloys Compd. 1999, 293–295, 433–436. [Google Scholar] [CrossRef]
  23. Young, K.; Ouchi, T.; Fetcenko, M.A. Roles of Ni, Cr, Mn, Sn, Co, and Al in C14 Laves phase alloys for NiMH battery application. J. Alloys Compd. 2009, 476, 774–781. [Google Scholar] [CrossRef]
  24. Bendersky, L.A.; Wang, K.; Levin, I.; Newbury, D.; Young, K.; Chao, B.; Creuziger, A. Ti12.5Zr21V10Cr8.5MnxCo1.5Ni46.5–x AB2-type metal hydride alloys for electrochemical storage application: Part 1. Structural characteristics. J. Power Sources 2012, 218, 474–486. [Google Scholar] [CrossRef]
  25. Young, K.; Ouchi, T.; Huang, B.; Reichman, B.; Blankenship, R. Improvement in −40 °C electrochemical properties of AB2 metal hydride alloy by silicon incorporation. J. Alloys Compd. 2013, 575, 65–72. [Google Scholar] [CrossRef]
  26. Young, K.; Reichman, B.; Fetcenko, M.A. Electrochemical performance of AB2 metal hydride alloys measured at −40 °C. J. Alloys Compd. 2013, 580, S349–S352. [Google Scholar] [CrossRef]
  27. Smith, J.F.; Lee, K.J. La-V (Lanthanum-Vanadium). In Binary Alloy Phase Diagram, 2nd ed.; Massalski, T.B., Okamoto, H., Subramanian, P.R., Kacprzak, L., Eds.; ASM International: Geauga County, OH, USA, 1990; Volume 3, pp. 2437–2439. [Google Scholar]
  28. Nei, J.; Young, K.; Salley, S.O.; Ng, K.Y.S. Determination of C14/C15 phase abundance in Laves phase alloys. Mater. Chem. Phys. 2012, 136, 520–527. [Google Scholar] [CrossRef]
  29. Nei, J.; Young, K.; Regmi, R.; Lawes, G.; Salley, S.O.; Ng, K.Y.S. Gaseous phase hydrogen storage and electrochemical properties of Zr8Ni21, Zr7Ni10, Zr9Ni11, and ZrNi metal hydride alloys. Int. J. Hydrog. Energy 2012, 37, 16042–16055. [Google Scholar] [CrossRef]
  30. Züttle, A. Materials for hydrogen storage. Mater. Today 2003, 6, 24–33. [Google Scholar] [CrossRef]
  31. Huang, T.; Li, J.; Yu, J.; Liu, Z.; Mao, S.; Zhang, Y.; Sun, G.; Han, J.; Ren, H.; Chen, J. Influence of partial substitution of Mo for Cr on structure and hydrogen storage characteristics of non-stoichiometric Laves phase TiCrB0.9 alloy. Int. J. Hydrog. Energy 2013, 38, 11955–11963. [Google Scholar] [CrossRef]
  32. Young, K.; Nei, J.; Ouchi, T.; Fetcenko, M.A. Phase abundances in AB2 metal hydride alloys and their correlations to various properties. J. Alloys Compd. 2011, 509, 2277–2284. [Google Scholar] [CrossRef]
  33. Johnston, R.L.; Hoffmann, R. Structure-bonding relationships in the Laves phases. Z. Anorg. Allg. Chem. 1992, 616, 105–120. [Google Scholar] [CrossRef]
  34. Boettinger, W.J.; Newbury, D.E.; Wang, K.; Bendersky, L.A.; Chiu, C.; Kattner, U.R.; Young, K.; Chao, B. Examination of multiphase (Zr,Ti)(V,Cr,Mn,Ni)2 Ni-MH electrode alloys: Part I. Dendritic solidification structure. Metall. Mater. Trans. A 2010, 41, 2033–2047. [Google Scholar] [CrossRef]
  35. Bendersky, L.A.; Wang, K.; Boettinger, W.J.; Newbury, D.E.; Young, K.; Chao, B. Examination of multiphase (Zr,Ti)(V,Cr,Mn,Ni)2 Ni-MH electrode alloys: Part II. Solid-state transformation of the interdendritic B2 phase. Metall. Mater. Trans. A 2010, 41, 1891–1906. [Google Scholar] [CrossRef]
  36. Wong, D.F.; Young, K.; Nei, J.; Wang, L.; Ng, K.Y.S. Effects of Nd-addition on the structural, hydrogen storage, and electrochemical properties of C14 metal hydride alloys. J. Alloys Compd. 2015, 647, 507–518. [Google Scholar] [CrossRef]
  37. Lide, D.R. CRC Handbook of Chemistry and Physics, 74th ed.; CRC Press: Boca Raton, FL, USA, 1993; pp. 6–22. [Google Scholar]
  38. Young, K.; Ouchi, T.; Fetcenko, M.A. Pressure-composition-temperature hysteresis in C14 Laves phase alloys: Part 1. Simple ternary alloys. J. Alloys Compd. 2009, 480, 428–433. [Google Scholar] [CrossRef]
  39. Kamegawa, A.; Shirasaki, K.; Tamura, T.; Kuriiwa, T.; Takamura, H.; Okada, M. Crystal structure and protium absorption properties of Ti-Cr-X alloys. Mater. Trans. 2002, 43, 470–473. [Google Scholar] [CrossRef]
  40. Kaiya, H.; Ookawa, T. Improvement in cycle life performance of high capacity nickel-metal hydride battery. J. Alloys Compd. 1995, 231, 598–603. [Google Scholar] [CrossRef]
  41. Kong, L.; Chen, B.; Young, K.; Koch, J.; Chan, A.; Li, W. Effects of Al- and Mn-contents in the negative MH alloy on the self-discharge and long-term storage properties of Ni/MH battery. J. Power Sources 2012, 213, 128–139. [Google Scholar] [CrossRef]
  42. Kim, S.; Chourashiya, M.G.; Park, C.; Park, C. Electrochemical performance of NAFION coated electrodes of hydriding combustion synthesized MgNi based composite hydride. Mater. Lett. 2013, 93, 81–84. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Young, K.-H.; Ouchi, T.; Huang, B.; Nei, J. Structure, Hydrogen Storage, and Electrochemical Properties of Body-Centered-Cubic Ti40V30Cr15Mn13X2 Alloys (X = B, Si, Mn, Ni, Zr, Nb, Mo, and La). Batteries 2015, 1, 74-90. https://doi.org/10.3390/batteries1010074

AMA Style

Young K-H, Ouchi T, Huang B, Nei J. Structure, Hydrogen Storage, and Electrochemical Properties of Body-Centered-Cubic Ti40V30Cr15Mn13X2 Alloys (X = B, Si, Mn, Ni, Zr, Nb, Mo, and La). Batteries. 2015; 1(1):74-90. https://doi.org/10.3390/batteries1010074

Chicago/Turabian Style

Young, Kwo-Hsiung, Taihei Ouchi, Baoquan Huang, and Jean Nei. 2015. "Structure, Hydrogen Storage, and Electrochemical Properties of Body-Centered-Cubic Ti40V30Cr15Mn13X2 Alloys (X = B, Si, Mn, Ni, Zr, Nb, Mo, and La)" Batteries 1, no. 1: 74-90. https://doi.org/10.3390/batteries1010074

Article Metrics

Back to TopTop