Next Article in Journal
Heat Treatment Optimization for a High Strength Al–Mn–Sc Alloy Fabricated by Selective Laser Melting
Next Article in Special Issue
Predictive Modeling of Molecular Mechanisms in Hydrogen Production and Storage Materials
Previous Article in Journal
Accelerated Direct Carbonation of Steel Slag and Cement Kiln Dust: An Industrial Symbiosis Strategy Applied in the Bergamo–Brescia Area
Previous Article in Special Issue
Thermodynamic Analysis of Chemical Hydrogen Storage: Energetics of Liquid Organic Hydrogen Carrier Systems Based on Methyl-Substituted Indoles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancement of the Desorption Properties of LiAlH4 by the Addition of LaCoO3

1
Energy Storage Research Group, Faculty of Ocean Engineering Technology and Informatics, Universiti Malaysia Terengganu, Kuala Nerus 21030, Terengganu, Malaysia
2
Department of Chemical and Materials Engineering, King Abdulaziz University, P.O. Box 80204, Jeddah 21589, Saudi Arabia
*
Author to whom correspondence should be addressed.
Materials 2023, 16(11), 4056; https://doi.org/10.3390/ma16114056
Submission received: 16 April 2023 / Revised: 24 May 2023 / Accepted: 27 May 2023 / Published: 29 May 2023
(This article belongs to the Special Issue Advance Materials for Hydrogen Storage)

Abstract

:
The high hydrogen storage capacity (10.5 wt.%) and release of hydrogen at a moderate temperature make LiAlH4 an appealing material for hydrogen storage. However, LiAlH4 suffers from slow kinetics and irreversibility. Hence, LaCoO3 was selected as an additive to defeat the slow kinetics problems of LiAlH4. For the irreversibility part, it still required high pressure to absorb hydrogen. Thus, this study focused on the reduction of the onset desorption temperature and the quickening of the desorption kinetics of LiAlH4. Here, we report the different weight percentages of LaCoO3 mixed with LiAlH4 using the ball-milling method. Interestingly, the addition of 10 wt.% of LaCoO3 resulted in a decrease in the desorption temperature to 70 °C for the first stage and 156 °C for the second stage. In addition, at 90 °C, LiAlH4 + 10 wt.% LaCoO3 can desorb 3.37 wt.% of H2 in 80 min, which is 10 times faster than the unsubstituted samples. The activation energies values for this composite are greatly reduced to 71 kJ/mol for the first stages and 95 kJ/mol for the second stages compared to milled LiAlH4 (107 kJ/mol and 120 kJ/mol for the first two stages, respectively). The enhancement of hydrogen desorption kinetics of LiAlH4 is attributed to the in situ formation of AlCo and La or La-containing species in the presence of LaCoO3, which resulted in a reduction of the onset desorption temperature and activation energies of LiAlH4.

1. Introduction

Due to the world’s population growth and civilizational developments, there has been an exponential increase in energy demand. The exploration of sustainable and renewable energy such as wind energy, hydropower, electrochemical energy and solar energy is believed to relieve the burden on the current energy demand [1,2,3]. Moreover, the amount of renewable energy consumed today is increasing which raises the standard for energy storage and transportation. However, the primary energy source is still fossil fuels despite the fact that they are unsustainable and have brought serious problems to the environment and human health [4,5]. Based on these facts, all countries have begun to implement a variety of strategies in order to avoid such problems. Future carbon emissions are predicted to be significantly reduced by hydrogen [6,7]. Over the last two decades, studies on hydrogen storage have received more attention in the literature [8,9]. In addition, hydrogen has the tremendous benefit of providing clean secondary energy with high energy content, no pollution and abundant resources [10,11]. There are three ways of storing hydrogen: (i) in a liquid form [12], (ii) in compressed gas [13] and (iii) in solid-state forms [14]. However, storing hydrogen in a solid-state form is attractive due to its safety reasons. The solid-state form has two main types: (i) physisorption and (ii) chemisorption. In physisorption, the hydrogen bind in the molecular to the surface of the host materials through weak interaction [15]. This physisorption process occurs for materials such as metal-organic frameworks [16] and carbon-based materials [17]. In 2023, a study by Mishra and colleagues [18] explored novel polycrystalline carbon nanotubes (PCNTs) in order to be used for large-scale hydrogen storage applications. They also revealed that PCNTs with moderate grain size are quite effective for these applications. Liang et al. [19] also reported an effective way to enhance hydrogen storage performance by using a porous carbon nanotube (CNT). However, the physisorption process has a low enthalpy of adsorption as revealed by previous studies [20,21]. Chemisorption is favorable because it can create a better storage capacity in ambient conditions [22]. Complex hydrides (such as NaBH4 and LiAlH4) and metal hydrides (such as MgH2) are typical materials for chemisorption. Other typical materials for chemisorption in chemical hydrogen storage are ammonia borane, methane, dimethyl ether, methanol, methanol and formic acid [23]. However, these chemical hydrides materials also suffer from some drawbacks, for example, ammonia borane (NH3BH3) requires a long induction time and high temperature to release hydrogen [24] and ammonia (NH3) must be heated at high temperature (above 650 °C) to achieve complete conversion [25].
In this study, lithium aluminium hydride (LiAlH4), which are complex hydrides, have been used due to their low cost, moderate temperature to release hydrogen and high hydrogen storage capacity (10.5 wt.%) [26,27]. However, the use of LiAlH4 for hydrogen storage applications is still constrained by its sluggish kinetics and irreversibility [28,29]. Furthermore, LiAlH4 needs to be handled according to the safety data sheet and Standard Operating Procedure (SOP). LiAlH4 needs to be stored away from heat/flame and moisture/water sources and kept sealed under an inert atmosphere. Personal protective equipment including safety glasses, lab coat, gloves and covered shoes should be used. In order to overcome the disadvantages, a few techniques for enhancing its kinetics and thermodynamics have been investigated. For example, doping LiAlH4 with various kinds of catalysts or additives such as metal oxides [30,31,32,33], metal halides [34,35], metal hydrides [36], pure metals [37,38] and carbon material [39] have been performed via the ball-milling method. However, the improvement of LiAlH4 is adversely affected by the ball-milling technique as proposed by Resan and colleagues [40].
A study by Jiao and colleagues [41] revealed that excellent desorption properties of LiAlH4 were accomplished after being doped with 1 wt.% NiCo nanoalloy encapsulated in graphene layers (NiCo@G). The starting hydrogen release for the doped samples was unexpectedly lowered by 109 °C compared to pure LiAlH4 (152 °C). For the desorption kinetic performance, as-milled LiAlH4 releases 1.6 wt.% of H2 at 150 °C within 10 min. Meanwhile, faster desorption kinetics can be observed after LiAlH4 is doped with 1 wt.% NiCo@G. The desorption kinetics for the doped samples carries on rapidly with 5.8 wt.% of H2 under the same conditions. As investigated by Li et al. [42], the addition of CoFe2O4 nanopowder to LiAlH4 effectively lowered the decomposition kinetic barrier for LiAlH4. Additionally, the activation energies of LiAlH4 doped with CoFe2O4 for the first two stages were 52.4 kJ/mol and 86.5 kJ/mol, respectively. During the heating process of LiAlH4 doped with CoFe2O4, the LiAlO2, LiFeO2, Al0.52Co0.48 and Fe0.98O phases appeared in the X-ray diffraction (XRD) patterns, indicating that this reaction could alter the reaction thermodynamics by lowering the enthalpy of the desorption reaction. The improvement of LiAlH4 after the addition of CoFe2O4 as a catalyst is also supported by Wei et al. [43]. It is exciting to note that they created CoFe2O4 using a thermal decomposition method and discovered that the starts of hydrogen release as compared to pure LiAlH4 (176 °C and 279 °C for the first two stages) were lowered by 103.3 °C and 97.9 °C for the first two stages, respectively.
Many studies have observed that rare earth metals such as Nb, Y and La exhibited excellent hydriding and dehydriding kinetics of MgH2 [44,45]. Additionally, adding 10 wt.% LaFeO3 lowered the activation energies and started the hydrogen release of LiAlH4 as indicated by our earlier research [46]. A further study discovered that the desorption properties of LiAlH4 were greatly benefited by the freshly formed La or La-containing species, Fe and LiFeO2 that had been generated during the heating process. Based on the study by Zhou et al. [47], the reduction temperature in the endothermic peaks showed that the desorption properties of LiAlH4 were positively affected after the inclusion of LaF3. On the other hand, Xueping et al. [48] looked into how the addition of La2O3 affected the desorption properties of LiAlH4–NH4Cl. The result obtained shows that adding La2O3 to LiAlH4–NH4Cl will increase the rate of hydrogen release while reducing the starting time.
In this context, with the aim of combining La and Co, Cobalt lanthanum oxide (LaCoO3) was introduced to prepare LaCoO3–LiAlH4 via the ball-milling method to improve the hydrogen storage properties of LiAlH4. Interestingly, various weight percentages of LaCoO3 were investigated to study the catalytic effects and the microstructure of hydrogen storage properties of LiAlH4. To the best of the author’s knowledge, no research on LaCoO3 and LiAlH4 performance has been published.

2. Materials and Methods

In this study, LaCoO3 was used as an additive and was synthesized using the solid-state method as discussed in our previous research [49]. A total of 0.121 g of citric acid (≥98% pure; Sigma Aldrich, St. Louis, MO, USA), 0.081 g of lanthanum oxide (≥99.9% pure; Aldrich Chemical Compound, Milwaukee, WI, USA) and 0.040 g of pure cobalt oxide (99.99% pure, Sigma Aldrich) were ground in an agate mortar and calcined at 950 °C in a furnace for 5 h. Different weight percentages of pure LaCoO3 (5, 10, 15 and 20) were milled in a planetary ball mill (NQM–0.4) with LiAlH4 (95% pure, Sigma Aldrich) for an hour (15 min milling time, 2 min resting time with 3 cycles) at a rotational speed of 400 rpm with a ball to powder ratio of 40:1. To prevent the reaction of samples with oxygen and moisture, all operations were conducted in an argon atmosphere glove box (MBRAUN UNIlab, Germany) containing low concentrations of O2 (<0.1 ppm) and H2O (<0.1 ppm) and high-purity Ar (99.99%).
The Gas Reaction Controller (GRC) evaluates quantitative analysis of the gas–solid reaction. It introduces a controlled amount of gas into the reaction chamber that holds a specimen. The pressure of the gas needs to be monitored while the temperature of the chamber is held constant or slowly changed. The instrument is connected to a computer and controlled by software (GrcLV), which performs fully automatic operations. Figure 1 below shows a schematic diagram of the working principles of the Sieverts apparatus. To determine the decomposition temperature of each sample, 150 mg of the materials were inserted into the sample holder and were heated at a rate of 5 °C/min from 30 to 250 °C. In the meantime, the desorption kinetics for all of the samples were heated for 80 min at 90 °C. This experiment was conducted by using Sievert-type apparatus (Advanced Materials Corporation, Pittsburgh, PA, USA).
Differential thermal analysis of the prepared samples was carried out using differential scanning calorimetry (Mettler Toledo, Columbus, OH, USA TGA/DSC 1) between 30–300 °C with a heating rate of 15, 20, 25 and 30 °C/min. Alumina crucibles were used as reference material. Samples weighing 6–8 mg were put into the crucible. The crucible was then sealed in a glass bottle to avoid the oxidation of the samples during the transfer from the glove box to the DSC apparatus. After the sample was placed inside the DSC, the gas inside the DSC was purged and constant argon flow (50 mL) was supplied. The designation of phase and composition of the samples were recorded by using XRD using Cu-Kα radiation at room temperature at a scanning range of 20° < 2θ < 80°. To prevent the oxidation of the samples, a small amount of the sample was spread uniformly on the sample holder and covered with scotch tape and followed by sealing with plastic wrap as described in our previous studies [50,51]. Fourier transform infrared (FTIR) spectroscopy spectra were collected at room temperature at a resolution of 4 cm−1 from 800 to 2000 cm−1. For FTIR, the samples were placed in microcentrifuge tubes (1.5 mL) during transportation from the glove box. The samples were placed on the FTIR, and the measurement was analyzed for a few seconds. We had additionally taken some precautionary steps so that the exposure of the samples to the air is minimized. Meanwhile, the surface configuration of the samples was analyzed using scanning electron microscopy (SEM; JEOL, Akishima, Tokyo, Japan (JSM-6360LA). Before each sample was analyzed, the samples were prepared on carbon tape and sprayed with a gold spray in a vacuum. Next, the samples were put in a sample container during the transfer process from the glove box to the SEM apparatus to prevent the samples from oxidizing. During the sample investigation, the samples were put inside the SEM under a vacuum state.

3. Results and Discussion

3.1. Onset Desorption Temperature

The hydrogen release curves with temperature are shown in Figure 2 for pure LiAlH4, milled LiAlH4 and LiAlH4 with different weight percents of LaCoO3. A remarkable low-temperature shift was observed for hydrogen released from LiAlH4 in the presence of LaCoO3. For pure LiAlH4, the first and second stages began at 151 °C and 182 °C, respectively with 7.39 wt.% of H2 released. However, after LiAlH4 was milled for an hour, the starting temperature release could be decreased from 151 °C to 147 °C and from 182 °C to 177 °C for the first and second stages, respectively, with 7.09 wt.% of the total H2 released. It is validated that the milling of the samples used has a positive effect on the peak of the desorption temperature of LiAlH4. This statement was also proven by Liu et al. [36]. Theoretically, LiAlH4 decomposes according to the following reactions [52,53]:
For the first stage,
3LiAlH4 → Li3AlH6 + 2Al + 3H2
and for the second stage,
Li3AlH6 → 3LiH + Al + 3/2H2
However, the temperature of onset desorption compared with milled LiAlH4 was reduced from 147 °C to 98 °C and from 177 °C to 153°C for the first and second stages, respectively, with 6.01 wt.% of H2 released when 5 wt.% of LaCoO3 was added. The desorption temperature drops to 70 °C and 105 °C for the first stage and 156 °C and 164 °C for the second stage, respectively, when 10 wt.% and 15 wt.% of LaCoO3 are added. The total H2 released for LiAlH4 + 10 wt.% LaCoO3 and LiAlH4 + 15 wt.% LaCoO3 were 5.86 wt.% and 5.72 wt.%, respectively. Increasing the amount of LaCoO3 to 20 wt.% also lowered the onset desorption temperature by 45 °C and 13 °C compared with pure LiAlH4 for the first and second stages, respectively. In spite of that, the amount of hydrogen released also decreased to 5.67 wt.%. This happens due to the zero hydrogen content in the LaCoO3; this outcome also has been testified by Xia et al. [54] and Ahmad et al. [55]. Table 1 summarizes the onset desorption temperature, hydrogen content and total hydrogen loss for pure LiAlH4, milled LiAlH4 and different weight percentages of LaCoO3 with LiAlH4 composites.
Figure 2. Onset desorption temperature of pure LiAlH4, milled LiAlH4 and LiAlH4 + x wt.% LaCoO3 (x = 5, 10, 15 and 20).
Figure 2. Onset desorption temperature of pure LiAlH4, milled LiAlH4 and LiAlH4 + x wt.% LaCoO3 (x = 5, 10, 15 and 20).
Materials 16 04056 g002

3.2. Isothermal Desorption Kinetics

For practical applications, the desorption rate of hydrogen storage materials is a crucial characteristic. Hence, Figure 3 below shows the isothermal desorption kinetics for LiAlH4 with different weight percents of LaCoO3 composites compared to pure LiAlH4 and milled LiAlH4 at 90 °C within 80 min. Under the same conditions, the LiAlH4 + 5 wt.% LaCoO3 systems desorbed 3.37 wt.% of H2. Increasing the amount of LaCoO3 to 10 wt.% also released 3.37 wt.% of H2 under the same period. However, slightly faster desorption kinetics can be observed for LiAlH4 + 5 wt.% LaCoO3 systems within 30 min at 90 °C. Meanwhile, LiAlH4 + 15 wt.% LaCoO3 and LiAlH4 + 20 wt.% LaCoO3 composites release approximately 3.06 wt.% and 2.50 wt.% of H2, respectively. The individual LiAlH4 releases less than 0.30 wt.% under the same conditions. This has proven that after the addition of LaCoO3, the isothermal desorption kinetics are 10 times faster than the unsubstituted LiAlH4. The next characterization was optimized to be 10 wt.% of LaCoO3. This is because the LiAlH4 + 10 wt.% LaCoO3 system decomposes at lower temperatures for the first stages compared to another amount of LaCoO3. Even though 5 wt.% also performed well for the isothermal desorption properties, the onset desorption temperature for 10 wt.% of LaCoO3 compared to that of LiAlH4 + 5 wt.% LaCoO3 (98 °C) decreased by 28 °C for the first stages, which indicates that the temperature for the LiAlH4 + 5 wt.% LaCoO3 is quite high for the desorption process to occur. Sulaiman and Ismail [56] also studied different weight percentages of SrFe12O19 (10, 20 and 50) on the thermal desorption of LiAlH4. A further study relating the catalytic effect of SrFe12O19 on the hydrogen storage performance of LiAlH4 was performed using 10 wt.% of SrFe12O19. Next, the effects of 5, 10, 15 and 20 wt.% of CoTiO3 were investigated on the onset desorption temperature of LiAlH4 [57]. However, 10 wt.% of CoTiO3 was chosen as an optimal amount to explore the desorption properties of LiAlH4. Thus, based on the discussion above, 10 wt.% of LaCoO3 was chosen for further characterization.

3.3. Activation Energy

The DSC traces for milled LiAlH4 and LiAlH4 + 10 wt.% LaCoO3 are included in Figure 4a,b, respectively. The figure below indicates that as the heating rate increases, the temperature also increases. Figure 4c exhibits the comparison for milled LiAlH4 and LiAlH4 + 10 wt.% LaCoO3 at a heating rate of 25 °C/min. For milled LiAlH4, there are four peaks (two exothermic peaks and two endothermic peaks) [58,59]. At 174 °C, the first exothermic peak can be regarded as being caused by LiAlH4 interacting with the surface hydroxyl impurities. These characteristics are due to the presence of surface hydroxyl impurities in the alanate powder, as already described in a previous study [60]. The second exothermic peak at 220 °C was directly related to the decomposition of LiAlH4 to Li3AlH6 and Al (as in Equation (1)) and the second endothermic peak at 273 °C was believed to happen due to the decomposition of Li3AlH6 to LiH and Al (as in Equation (2)). In contrast with the unsubstituted samples, the temperature of each peak dropped after the inclusion of 10 wt.% LaCoO3. The first exothermic and endothermic peaks were 134 °C and 167 °C, respectively. Further heating led to additional exothermic and endothermic peaks at 188 °C and 242 °C, respectively. It has been shown that the addition of 10 wt.% LaCoO3 has shifted the peaks to a lower temperature for each reaction, proving that the desorption kinetics performance of LiAlH4 was improved. This could be associated with the fact that the introduction of 10 wt.% LaCoO3 leads to significant particle size refinement of LiAlH4. The lower desorption peaks of the composite can effectively contribute to the kinetic improvement of LiAlH4 via the addition of 10 wt.% LaCoO3. Furthermore, LaCoO3 is reported to have a hardness of 6.5 Mohs compared with the 4.0 Mohs of LiAlH4 [61]. The hardness of the LaCoO3 had thus broken the LiAlH4 particles. Therefore, smaller particle size could enhance the kinetic enhancement and lower the activation energy of the LiAlH4. The enhancement in the desorption kinetics performance is related to the energy barrier for hydrogen released from LiAlH4.
In order to obtain kinetic information of the hydrogen desorption from milled LiAlH4 and LiAlH4 + 10 wt.% LaCoO3, the activation energies (EA) were examined using the Kissinger equation, as shown in Equation (3) below:
ln [β/Tp2] = −EA/RTp + A
where A = linear constant, R = gas constant, Tp = peak temperature in the DSC curve and β = heating rate. The activation value can be obtained by analyzing the slope of the Kissinger plot, ln (β/Tp2) vs. 1000/Tp. The first stage of EA was applied to the decomposition of LiAlH4 as shown in Figure 5a while the second stage of EA was applied to the decomposition of Li3AlH6 as illustrated in Figure 5b. From the figure, the EA are 107 kJ/mol and 71 kJ/mol for milled LiAlH4 and LiAlH4 + 10 wt.% LaCoO3, respectively. For the second stage, milled LiAlH4 has an EA of 120 kJ/mol and this value decreased by 25 kJ/mol after the addition of 10 wt.% LaCoO3 (95 kJ/mol). This proves that adding LaCoO3 as an additive could lower the EA of LiAlH4 for both stages. A previous study by Li et al. [62] exposed that the EA of LiAlH4 are 78.2 kJ/mol and 90.8 kJ/mol, 38 kJ/mol and 42.2 kJ/mol lower than pure LiAlH4 for the first and second stages, respectively, after the addition of K2TiF6. Furthermore, Zhang et al. [63] revealed that the EA was reduced by 18 kJ/mol for the first stages and 73.3 kJ/mol compared to milled LiAlH4 (90.6 kJ/mol and 144.7 kJ/mol) after adding Li2TiO3. Adding SrFe12O19 also reduced the EA from 103 kJ/mol to 76 kJ/mol for the first stage and from 111 kJ/mol to 96 kJ/mol for the second stage compared with milled LiAlH4 [56]. Based on this discussion, it can be concluded that adding LaCoO3 could lower the EA of LiAlH4. Owing to this reduction of the EA, the desorption kinetics of LiAlH4 are enhanced.

3.4. Morphology and Microstructures

Figure 6 illustrates the SEM micrographs of pure LiAlH4, milled LiAlH4 and LiAlH4 + 10 wt.% LaCoO3. As shown in Figure 6a, the particles of pure LiAlH4 are shaped like blocks and are coarse, as reported by a previous study [64]. After LiAlH4 was milled for an hour, the particles became agglomerated and inhomogeneous, as shown in Figure 6b and proven by Ares and co-workers [65]. However, after the addition of LaCoO3, the particle size became small compared to pure LiAlH4 and milled LiAlH4 (Figure 6c). This result is consistent with earlier research showing that adding an additive or catalyst can reduce the particle size of LiAlH4 [66]. In addition, after LiAlH4 was doped with MnFe2O4, the samples decreased in particle size, resulting in more grain boundaries and increased surface area, which enhanced the desorption properties of LiAlH4 [67]. Therefore, it can be concluded that after the addition of 10 wt.% LaCoO3 through the ball-milling method, a smaller particle size can be observed, which may benefit the desorption behavior of the LiAlH4.

3.5. Particles Size Distribution

The data plotted in Figure 7 show the average particle size distribution (PSD) of composites with and without the addition of LaCoO3. For pure LiAlH4 (as in Figure 7a), the particle size distribution calculated using ImageJ was 39.62 µm. After an hour of the milling process of LiAlH4, the samples improve the quality of the powders in terms of reduction in the particle size, as shown in Figure 7b. The particle size of milled LiAlH4 was reduced to 0.66 µm. However, the particle size of LiAlH4 + 10 wt.% LaCoO3 is smaller if compared to the unsubstituted samples previously described in the SEM part. The average particle size of LiAlH4 + 10 wt.% LaCoO3 was 0.27 µm, as shown in Figure 7c. In summary, the progressive phenomenon of reduction in particle size is reached, indicating that the inclusion of LaCoO3 using the ball-milling method improved the desorption performance of LiAlH4.

3.6. Phase Structure

The XRD patterns of pure LiAlH4 and LiAlH4 after being milled for an hour are displayed in Figure 8a,b, respectively. At this stage, all the peaks correspond to LiAlH4, which perfectly matches the standard pattern of LiAlH4 (JCPDS card no. 73-461). However, the intensity of the sample lowered compared to pure LiAlH4. According to these patterns, the milling process will result in a reduction in the intensity of the peaks. A previous study by Rahmaninasab and colleagues [68] revealed that the milling time causes the intensity to be reduced. This is due to the increase in the lattice strain and internal energy during the milling process, as revealed by Dittrich et al. [69]. This also demonstrates that LiAlH4 is stable during the milling process, as stated in the previous studies [32,70,71]. The XRD pattern of LiAlH4 + 10 wt.% LaCoO3 is presented in Figure 8c. It was seen that the peaks related to LiAlH4 become broad and had a lower intensity when compared to unsubstituted LiAlH4. No new peaks or peaks for LaCoO3 were detected during the milling process of LaCoO3.
The FTIR spectra were carried out to confirm the presence of 10 wt.% LaCoO3, as indicated in Figure 9. Two regions of active IR peaks that are the bending mode and the stretching mode were detected in the range of 800–900 cm−1 and 1600–1800 cm−1 [72], respectively, for pure LiAlH4, milled LiAlH4 and LiAlH4 + 10 wt.% LaCoO3. After the addition of 10 wt.% LaCoO3, a new peak was detected at 1383 cm−1, suggesting that LiAlH4 was decomposed to Li3AlH6 and Al during the ball-milling process. A previous study led by Shen et al. [73] also revealed that the peaks that appear around 1400 cm−1 represent the stretching peak of Li3AlH6. Furthermore, after the addition of NiFe2O4, Wei et al. [43] discovered a weak stretching mode at about 1404 cm−1 which belongs to the Al–H stretching of [AlH6]3−.
To observe the change of phase structure of the LiAlH4 with LaCoO3 during the desorption process, an XRD scan has been conducted from 20° to 80°. As shown in Figure 10a below, after LiAlH4 + 10 wt.% LaCoO3 was heated, LiH and Al peaks were found, indicating that Equations (1) and (2) had occurred. A new peak for AlCo was identified, proving that the reaction of LiAlH4 and LaCoO3 may have occurred during the heating process. No peaks for La or La-containing species were detected as displayed in Figure 10a. However, after the amount of LaCoO3 was increased to 30 wt.% (as shown in Figure 10b), the same peaks for AlCo and LiH/Al were spotted.
From the above discussions, the enhancement of LiAlH4 may be ascribed to a few factors. The new active species of AlCo formed after the heating process acts as a real additive in enhancing the hydrogen storage performance of LiAlH4. Additionally, it has been evidenced that Co additives may benefit the desorption properties of light metal complex hydrides [74,75]. In addition, Lv and colleagues [76] managed to show that the Co element significantly improves the hydrogen storage performance of the alloy. Other than that, Ali et al. [77] indicated that the formation of Al–Co alloys during the heating process of NaAlH4 and CoTiO3 is the cause of the enhanced desorption behavior of NaAlH4. Furthermore, the starting hydrogen released by NaAlH4 compared to milled NaAlH4 (200 °C) decreased by ~65 °C after the addition of CoTiO3. Moreover, hydrogen storage desorption of LiAlH4 was improved by the presence of CoFe2O4 nanoparticles [42]. The start of hydrogen release dropped to 65 °C at the first stage and the activation energy lowered by 42.4 kJ/mol in the presence of CoFe2O4 nanoparticles. One of the reasons for the enhancement of LiAlH4 is due to the formation of Al–Co phases during the heating process of LiAlH4 and CoFe2O4 composites. However, no peak for La or La-containing species (such as La–O, Co-La or Al-La) was detected due to the low amount of LaCoO3 in the composite or La being amorphous. This result was consistent with our previous study [46]. Hence, it is possible to speculate in situ formation of AlCo and La or La-containing species during the heating process attributed to improving the desorption kinetics of LiAlH4.

4. Conclusions

In this paper, thermal desorption of LiAlH4 has been studied with and without the addition of different weight percentages of LaCoO3 additive (5, 10, 15 and 20). It was found that the addition of 10 wt.% LaCoO3 is an effective additive which substantially lowered the onset desorption temperature by 81 °C and 26 °C for LiAlH4 compared with pure LiAlH4 for the first two stages, respectively. Furthermore, the desorption kinetics are 10 times faster than the unsubstituted LiAlH4 at 90 °C for 80 min. The LaCoO3 additive considerably reduced the activation energies from 107 kJ/mol to 71 kJ/mol for the first stages and from 120 kJ/mol to 95 kJ/mol for the second stages. From our observation, the morphology of LiAlH4 + 10 wt.% LaCoO3 shows that the particle size becomes smaller and less agglomerate compared to the unsubstituted LiAlH4. The result of particle size distribution also proved that the addition of LaCoO3 additive reduced the particle size to 0.27 µm, 39.35 µm lower than pure LiAlH4 and 0.39 µm reduced from milled LiAlH4. Furthermore, the formation of new phases (AlCo and La or La-containing species) have a notable impact on enhancing the hydrogen storage behavior of LiAlH4, which can be determined by observing the reaction between LiAlH4 and LaCoO3 during the heating process.

Author Contributions

Writing—original draft, N.S. and M.I.; methodology, N.S.; writing—review and editing, N.A.A., S.-U.R., H.S.B., H.A., A.A.T. and U.S.; Visualization, N.A.A., S.-U.R., H.A., A.A.T. and U.S.; supervision, M.I. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was funded by Institutional Fund Projects under grant no. (IFPIP: 407-135-1443). The authors gratefully acknowledge the technical and financial support provided by the Ministry of Education and King Abdulaziz University, DSR, Jeddah, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

This research work was funded by Institutional Fund Projects under grant no. (IFPIP: 407-135-1443). The authors gratefully acknowledge technical and financial support provided by the Ministry of Education and King Abdulaziz University, DSR, Jeddah, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Du, W.; Wang, X.; Zhan, J.; Sun, X.; Kang, L.; Jiang, F.; Zhang, X.; Shao, Q.; Dong, M.; Liu, H.; et al. Biological cell template synthesis of nitrogen-doped porous hollow carbon spheres/MnO2 composites for high-performance asymmetric supercapacitors. Electrochim. Acta 2019, 296, 907–915. [Google Scholar] [CrossRef]
  2. Wu, D.; Xie, X.; Zhang, J.; Ma, Y.; Hou, C.; Sun, X.; Yang, X.; Zhang, Y.; Kimura, H.; Du, W. Embedding NiS nanoflakes in electrospun carbon fibers containing NiS nanoparticles for hybrid supercapacitors. Chem. Eng. J. 2022, 446, 137262. [Google Scholar] [CrossRef]
  3. Wu, D.; Xie, X.; Ma, Y.; Zhang, J.; Hou, C.; Sun, X.; Yang, X.; Zhang, Y.; Kimura, H.; Du, W. Morphology controlled hierarchical NiS/carbon hexahedrons derived from nitrilotriacetic acid-assembly strategy for high-performance hybrid supercapacitors. Chem. Eng. J. 2022, 433, 133673. [Google Scholar] [CrossRef]
  4. Guo, X.; Liang, C.; Umar, M.; Mirza, N. The impact of fossil fuel divestments and energy transitions on mutual funds performance. Technol. Forecast. Soc. Change 2022, 176, 121429. [Google Scholar] [CrossRef]
  5. Hansen, T.A. Stranded assets and reduced profits: Analyzing the economic underpinnings of the fossil fuel industry’s resistance to climate stabilization. Renew. Sust. Energ. Rev. 2022, 158, 112144. [Google Scholar] [CrossRef]
  6. Hienuki, S.; Noguchi, K.; Shibutani, T.; Fuse, M.; Noguchi, H.; Miyake, A. Risk identification for the introduction of advanced science and technology: A case study of a hydrogen energy system for smooth social implementation. Int. J. Hydrogen Energy 2020, 45, 15027–15040. [Google Scholar] [CrossRef]
  7. Chapman, A.; Itaoka, K.; Farabi-Asl, H.; Fujii, Y.; Nakahara, M. Societal penetration of hydrogen into the future energy system: Impacts of policy, technology and carbon targets. Int. J. Hydrogen Energy 2020, 45, 3883–3898. [Google Scholar] [CrossRef]
  8. Lebrouhi, B.E.; Djoupo, J.J.; Lamrani, B.; Benabdelaziz, K.; Kousksou, T. Global hydrogen development—A technological and geopolitical overview. Int. J. Hydrogen Energy 2022, 47, 7016–7048. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Zheng, J.; Lu, Z.; Song, M.; He, J.; Wu, F.; Zhang, L. Boosting the hydrogen storage performance of magnesium hydride with metal organic framework-derived Cobalt@Nickel oxide bimetallic catalyst. Chin. J. Chem. Eng. 2022, 52, 161–171. [Google Scholar] [CrossRef]
  10. Zhang, J.; Hou, Q.; Guo, X.; Yang, X. Achieve high-efficiency hydrogen storage of MgH2 catalyzed by nanosheets CoMoO4 and rGO. J. Alloys Compd. 2022, 911, 165153. [Google Scholar] [CrossRef]
  11. Nasser, M.; Megahed, T.F.; Ookawara, S.; Hassan, H. Techno-economic assessment of clean hydrogen production and storage using hybrid renewable energy system of PV/Wind under different climatic conditions. Sustain. Energy Technol. Assess. 2022, 52, 102195. [Google Scholar] [CrossRef]
  12. Nasri, A.; Jaleh, B.; Shabanlou, E.; Nasrollahzadeh, M.; Ali Khonakdar, H.; Kruppke, B. Ionic liquid-based (nano)catalysts for hydrogen generation and storage. J. Mol. Liq. 2022, 365, 120142. [Google Scholar] [CrossRef]
  13. Wang, H.; Zhao, Y.; Dong, X.; Yang, J.; Guo, H.; Gong, M. Thermodynamic analysis of low-temperature and high-pressure (cryo-compressed) hydrogen storage processes cooled by mixed-refrigerants. Int. J. Hydrogen Energy 2022, 47, 28932–28944. [Google Scholar] [CrossRef]
  14. Sürer, M.G.; Arat, H.T. Advancements and current technologies on hydrogen fuel cell applications for marine vehicles. Int. J. Hydrogen Energy 2022, 47, 19865–19875. [Google Scholar] [CrossRef]
  15. Sahoo, R.K.; Sahu, S. Transition from chemisorption to physisorption of H2 on Ti functionalized [2,2,2]paracyclophane: A computational search for hydrogen storage. J. Energy Storage 2023, 63, 106951. [Google Scholar] [CrossRef]
  16. Zhu, Z.W.; Zheng, Q.R. Investigation of cryo-adsorption hydrogen storage capacity of rapidly synthesized MOF-5 by mechanochemical method. Int. J. Hydrogen Energy 2023, 48, 5166–5174. [Google Scholar] [CrossRef]
  17. Pandey, A.P.; Shaz, M.A.; Sekkar, V.; Tiwari, R.S. Synergistic effect of CNT bridge formation and spillover mechanism on enhanced hydrogen storage by iron doped carbon aerogel. Int. J. Hydrogen Energy 2022, in press, corrected proof. [Google Scholar] [CrossRef]
  18. Mishra, S.; Luhadiya, N.; Kundalwal, S.I. Atomistic insights into the H2 adsorption and desorption behavior of novel Li-functionalized polycrystalline CNTs. Carbon 2023, 207, 23–35. [Google Scholar] [CrossRef]
  19. Liang, H.; Du, X.; Li, J.; Sun, L.; Song, M.; Li, W. Manipulating active sites on carbon nanotube materials for highly efficient hydrogen storage. Appl. Surf. Sci. 2023, 619, 156740. [Google Scholar] [CrossRef]
  20. Desai, F.J.; Uddin, M.N.; Rahman, M.M.; Asmatulu, R. A critical review on improving hydrogen storage properties of metal hydride via nanostructuring and integrating carbonaceous materials. Int. J. Hydrogen Energy 2023, in press, corrected proof. [Google Scholar] [CrossRef]
  21. Meduri, S.; Nandanavanam, J. Materials for hydrogen storage at room temperature—An overview. Mater. Today Proc. 2023, 72, 1–8. [Google Scholar] [CrossRef]
  22. Zhang, X.; Sun, Y.; Xia, G.; Yu, X. Light-weight solid-state hydrogen storage materials characterized by neutron scattering. J. Alloys Compd. 2022, 899, 163254. [Google Scholar] [CrossRef]
  23. Pawelczyk, E.; Łukasik, N.; Wysocka, I.; Rogala, A.; Gębicki, J. Recent progress on hydrogen storage and production using chemical hydrogen carriers. Energies 2022, 15, 4964. [Google Scholar] [CrossRef]
  24. Akbayrak, S.; Özkar, S. Ammonia borane as hydrogen storage materials. Int. J. Hydrogen Energy 2018, 43, 18592–18606. [Google Scholar] [CrossRef]
  25. Andersson, J.; Grönkvist, S. Large-scale storage of hydrogen. Int. J. Hydrogen Energy 2019, 44, 11901–11919. [Google Scholar] [CrossRef]
  26. Sazelee, N.; Ismail, M. Recent advances in catalyst-enhanced LiAlH4 for solid-state hydrogen storage: A review. Int. J. Hydrogen Energy 2021, 46, 9123–9141. [Google Scholar] [CrossRef]
  27. Feng, W.; Xue Ping, Z.; Xue Qin, M.; Yue Hui, D.; Xiao Yan, Z.; Zi Wei, X. Catalytic effects of FGO-Ni addition on the dehydrogenation properties of LiAlH4. Int. J. Hydrogen Energy 2022, 47, 26458–26467. [Google Scholar] [CrossRef]
  28. Sazelee, N.; Ali, N.; Yahya, M.; Mustafa, N.; Halim Yap, F.; Mohamed, S.; Ghazali, M.; Suwarno, S.; Ismail, M. Recent advances on Mg–Li–Al systems for solid-state hydrogen storage: A Review. Front. Energy Res. 2022, 10, 875405. [Google Scholar] [CrossRef]
  29. Graetz, J.; Wegrzyn, J.; Reilly, J.J. Regeneration of lithium aluminum hydride. J. Am. Chem. Soc. 2008, 130, 17790–17794. [Google Scholar] [CrossRef]
  30. Li, Z.; Li, P.; Wan, Q.; Zhai, F.; Liu, Z.; Zhao, K.; Wang, L.; Lü, S.; Zou, L.; Qu, X. Dehydrogenation improvement of LiAlH4 catalyzed by Fe2O3 and Co2O3 nanoparticles. J. Phys. Chem. C 2013, 117, 18343–18352. [Google Scholar] [CrossRef]
  31. Ali, N.A.; Idris, N.H.; Sazelee, N.A.; Yahya, M.S.; Halim Yap, F.A.; Ismail, M. Catalytic effects of MgFe2O4 addition on the dehydrogenation properties of LiAlH4. Int. J. Hydrogen Energy 2019, 44, 28227–28234. [Google Scholar] [CrossRef]
  32. Sazelee, N.A.; Yahya, M.S.; Ali, N.A.; Idris, N.H.; Ismail, M. Enhancement of dehydrogenation properties in LiAlH4 catalysed by BaFe12O19. J. Alloys Compd. 2020, 835, 155183. [Google Scholar] [CrossRef]
  33. Ismail, M.; Ali, N.A.; Sazelee, N.A.; Suwarno, S. Catalytic effect of Al2TiO5 on the dehydrogenation properties of LiAlH4. Int. J. Hydrogen Energy 2022, 47, 31903–31910. [Google Scholar] [CrossRef]
  34. Cao, Z.; Ma, X.; Wang, H.; Ouyang, L. Catalytic effect of ScCl3 on the dehydrogenation properties of LiAlH4. J. Alloys Compd. 2018, 762, 73–79. [Google Scholar] [CrossRef]
  35. Ismail, M.; Zhao, Y.; Yu, X.B.; Dou, S.X. Effects of NbF5 addition on the hydrogen storage properties of LiAlH4. Int. J. Hydrogen Energy 2010, 35, 2361–2367. [Google Scholar] [CrossRef]
  36. Liu, S.S.; Li, Z.B.; Jiao, C.L.; Si, X.L.; Yang, L.N.; Zhang, J.; Zhou, H.Y.; Huang, F.-L.; Gabelica, Z.; Schick, C. Improved reversible hydrogen storage of LiAlH4 by nano-sized TiH2. Int. J. Hydrogen Energy 2013, 38, 2770–2777. [Google Scholar] [CrossRef]
  37. Varin, R.A.; Zbroniec, L.; Czujko, T.; Wronski, Z.S. The effects of nanonickel additive on the decomposition of complex metal hydride LiAlH4 (lithium alanate). Int. J. Hydrogen Energy 2011, 36, 1167–1176. [Google Scholar] [CrossRef]
  38. Xueping, Z.; Ping, L.; Fuqiang, A.; Guoqing, W.; Xuanhui, Q. Effects of Ti and Fe additives on hydrogen release from lithium alanate. Rare Metal. Mat. Eng. 2008, 37, 400–403. [Google Scholar] [CrossRef]
  39. Tan, C.Y.; Tsai, W.T. Effects of Ni and Co-decorated MWCNTs addition on the dehydrogenation behavior and stability of LiAlH4. Int. J. Hydrogen Energy 2015, 40, 14064–14071. [Google Scholar] [CrossRef]
  40. Resan, M.; Hampton, M.D.; Lomness, J.K.; Slattery, D.K. Effect of TixAly catalysts on hydrogen storage properties of LiAlH4 and NaAlH4. Int. J. Hydrogen Energy 2005, 30, 1417–1421. [Google Scholar] [CrossRef]
  41. Jiao, C.; Sun, L.; Xu, F.; Liu, S.S.; Zhang, J.; Jiang, X.; Yang, L. NiCo nanoalloy encapsulated in graphene layers for improving hydrogen storage properties of LiAlH4. Sci. Rep. 2016, 6, 27429. [Google Scholar] [CrossRef] [PubMed]
  42. Li, Z.; Zhai, F.; Wan, Q.; Liu, Z.; Shan, J.; Li, P.; Volinsky, A.A.; Qu, X. Enhanced hydrogen storage properties of LiAlH4 catalyzed by CoFe2O4 nanoparticles. RSC Adv. 2014, 4, 18989–18997. [Google Scholar] [CrossRef]
  43. Wei, S.; Liu, J.; Xia, Y.; Zhang, H.; Cheng, R.; Sun, L.; Xu, F.; Huang, P.; Rosei, F.; Pimerzin, A.A.; et al. Remarkable catalysis of spinel ferrite XFe2O4 (X = Ni, Co, Mn, Cu, Zn) nanoparticles on the dehydrogenation properties of LiAlH4: An experimental and theoretical study. J. Mater. Sci. Technol. 2022, 111, 189–203. [Google Scholar] [CrossRef]
  44. Wu, C.; Wang, Y.; Liu, Y.; Ding, W.; Sun, C. Enhancement of hydrogen storage properties by in situ formed LaH3 and Mg2NiH4 during milling MgH2 with porous LaNiO3. Catal. Today 2018, 318, 113–118. [Google Scholar] [CrossRef]
  45. Sazelee, N.A.; Idris, N.H.; Md Din, M.F.; Yahya, M.S.; Ali, N.A.; Ismail, M. LaFeO3 synthesised by solid-state method for enhanced sorption properties of MgH2. Results Phys. 2020, 16, 102844. [Google Scholar] [CrossRef]
  46. Sazelee, N.A.; Yahya, M.S.; Idris, N.H.; Md Din, M.F.; Ismail, M. Desorption properties of LiAlH4 doped with LaFeO3 catalyst. Int. J. Hydrogen Energy 2019, 44, 11953–11960. [Google Scholar] [CrossRef]
  47. Zhou, S.; Zou, J.; Zeng, X.; Ding, W. Effects of REF3 (RE = Y, La, Ce) additives on dehydrogenation properties of LiAlH4. Int. J. Hydrogen Energy 2014, 39, 11642–11650. [Google Scholar] [CrossRef]
  48. Xueping, Z.; Qiuhua, M.; Zhen, F.; Feng, M.; Shenglin, L.; Xin, F.; Guo, X.; Jiaojiao, Z. Effect of La3+ on dehydrogenation capacity of the LiAlH4− NH4Cl system. Rare Metal. Mat. Eng. 2014, 43, 2075–2078. [Google Scholar] [CrossRef]
  49. Sazelee, N.; Md Din, M.F.; Ismail, M.; Rather, S.U.; Bamufleh, H.S.; Alhumade, H.; Taimoor, A.A.; Saeed, U. Effect of LaCoO3 synthesized via solid-state method on the hydrogen storage properties of MgH2. Materials 2023, 16, 2449. [Google Scholar] [CrossRef]
  50. Halim Yap, F.A.; Mustafa, N.S.; Ismail, M. A study on the effects of K2ZrF6 as an additive on the microstructure and hydrogen storage properties of MgH2. RSC Adv. 2015, 5, 9255–9260. [Google Scholar] [CrossRef]
  51. Ismail, M. Effect of adding different percentages of HfCl4 on the hydrogen storage properties of MgH2. Int. J. Hydrogen Energy 2021, 46, 8621–8628. [Google Scholar] [CrossRef]
  52. Luo, Y.; Wang, Q.; Li, J.; Xu, F.; Sun, L.; Zou, Y.; Chu, H.; Li, B.; Zhang, K. Enhanced hydrogen storage/sensing of metal hydrides by nanomodification. Mater. Today Nano 2020, 9, 100071. [Google Scholar] [CrossRef]
  53. He, Q.; Zhu, D.; Wu, X.; Dong, D.; Xu, M.; Tong, Z. Hydrogen desorption properties of LiBH4/xLiAlH4 (x = 0.5, 1, 2) composites. Molecules 2019, 24, 1861. [Google Scholar] [CrossRef]
  54. Xia, Y.; Zhang, H.; Sun, Y.; Sun, L.; Xu, F.; Sun, S.; Zhang, G.; Huang, P.; Du, Y.; Wang, J. Dehybridization effect in improved dehydrogenation of LiAlH4 by doping with two-dimensional Ti3C2. Mater. Today Nano 2019, 8, 100054. [Google Scholar] [CrossRef]
  55. Ahmad, M.A.N.; Sazelee, N.A.; Ali, N.A.; Ismail, M. Enhancing the dehydrogenation properties of LiAlH4 using K2NiF6 as additive. Int. J. Hydrogen Energy 2022, 47, 24843–24851. [Google Scholar] [CrossRef]
  56. Sulaiman, N.N.; Ismail, M. Catalytic effect of SrFe12O19 on the hydrogen storage properties of LiAlH4. Int. J. Hydrogen Energy 2017, 42, 19126–19134. [Google Scholar] [CrossRef]
  57. Ali, N.A.; Ahmad, M.A.N.; Yahya, M.S.; Sazelee, N.; Ismail, M. Improved dehydrogenation properties of LiAlH4 by addition of nanosized CoTiO3. Nanomaterials 2022, 12, 3921. [Google Scholar] [CrossRef]
  58. Wei, S.; Xue, S.; Huang, C.; Che, B.; Zhang, H.; Sun, L.; Xu, F.; Xia, Y.; Cheng, R.; Zhang, C. Multielement synergetic effect of NiFe2O4 and h-BN for improving the dehydrogenation properties of LiAlH4. Inorg. Chem. Front. 2021, 8, 3111–3126. [Google Scholar] [CrossRef]
  59. Li, Z.L.; Zhai, F.Q.; Qiu, H.C.; Wan, Q.; Li, P.; Qu, X.H. Dehydrogenation characteristics of ZrC-doped LiAlH4 with different mixing conditions. Rare Met. 2020, 39, 383–391. [Google Scholar] [CrossRef]
  60. Ismail, M.; Sinin, A.; Sheng, C.; Nik, W. Desorption behaviours of lithium alanate with metal oxide nanopowder additives. Int. J. Electrochem. Sci. 2014, 9, 495. [Google Scholar]
  61. Ismail, M.; Sazelee, N.A.; Ali, N.A.; Suwarno, S. Catalytic effect of SrTiO3 on the dehydrogenation properties of LiAlH4. J. Alloys Compd. 2021, 855, 157475. [Google Scholar] [CrossRef]
  62. Li, Z.; Liu, S.; Si, X.; Zhang, J.; Jiao, C.; Wang, S.; Liu, S.; Zou, Y.-J.; Sun, L.; Xu, F. Significantly improved dehydrogenation of LiAlH4 destabilized by K2TiF6. Int. J. Hydrogen Energy 2012, 37, 3261–3267. [Google Scholar] [CrossRef]
  63. Zhang, T.; Isobe, S.; Wang, Y.; Liu, C.; Hashimoto, N.; Takahashi, K. Enhanced hydrogen desorption properties of LiAlH4 by doping lithium metatitanate. Phys. Chem. Chem. Phys. 2016, 18, 27623–27629. [Google Scholar] [CrossRef] [PubMed]
  64. Varin, R.; Zbroniec, L. Decomposition behavior of unmilled and ball milled lithium alanate (LiAlH4) including long-term storage and moisture effects. J. Alloys Compd. 2010, 504, 89–101. [Google Scholar] [CrossRef]
  65. Ares, J.R.; Aguey Zinsou, K.F.; Porcu, M.; Sykes, J.M.; Dornheim, M.; Klassen, T.; Bormann, R. Thermal and mechanically activated decomposition of LiAlH4. Mater. Res. Bull. 2008, 43, 1263–1275. [Google Scholar] [CrossRef]
  66. Li, L.; Qiu, F.; Wang, Y.; Xu, Y.; An, C.; Liu, G.; Jiao, L.; Yuan, H. Enhanced hydrogen storage properties of TiN–LiAlH4 composite. Int. J. Hydrogen Energy 2013, 38, 3695–3701. [Google Scholar] [CrossRef]
  67. Zhai, F.; Li, P.; Sun, A.; Wu, S.; Wan, Q.; Zhang, W.; Li, Y.; Cui, L.; Qu, X. Significantly improved dehydrogenation of LiAlH4 destabilized by MnFe2O4 nanoparticles. J. Phys. Chem. C 2012, 116, 11939–11945. [Google Scholar] [CrossRef]
  68. Rahmaninasab, M.A.; Raygan, S.; Abdizadeh, H.; Pourabdoli, M.; Mirghaderi, S.H. Properties of activated MgH2+ mischmetal nanostructured composite produced by ball-milling. Mater. Renew. Sustain. Energy 2018, 7, 15. [Google Scholar] [CrossRef]
  69. Dittrich, M.; Schumacher, G. Evolution of crystallite size, lattice parameter and internal strain in Al precipitates during high energy ball milling of partly amorphous Al87Ni8La5 alloy. Mater. Sci. Eng. A 2014, 604, 27–33. [Google Scholar] [CrossRef]
  70. Li, P.; Li, Z.; Zhai, F.; Wan, Q.; Li, X.; Qu, X.; Volinsky, A.A. NiFe2O4 nanoparticles catalytic effects of improving LiAlH4 dehydrogenation properties. J. Phys. Chem. C 2013, 117, 25917–25925. [Google Scholar] [CrossRef]
  71. Ismail, M.; Zhao, Y.; Yu, X.B.; Nevirkovets, I.P.; Dou, S.X. Significantly improved dehydrogenation of LiAlH4 catalysed with TiO2 nanopowder. Int. J. Hydrogen Energy 2011, 36, 8327–8334. [Google Scholar] [CrossRef]
  72. Pratthana, C.; Aguey-Zinsou, K.F. Surfactant induced synthesis of LiAlH4 and NaAlH4 nanoparticles for hydrogen storage. Appl. Sci. 2022, 12, 4742. [Google Scholar] [CrossRef]
  73. Shen, X.; Zhang, X.; Xiao, Q.; Liu, H. Catalytical enhancement on hydrogen production from LiAlH4 by Fe–Fe2O3 addition. Int. J. Hydrogen Energy 2022, 47, 16964–16977. [Google Scholar] [CrossRef]
  74. Wang, Y.; Ren, Q.; Wang, Y.; Li, L.; Song, D.; Jiao, L.; Yuan, H. A facile two-step synthesis and dehydrogenation properties of NaAlH4 catalyzed with Co–B. Int. J. Hydrogen Energy 2010, 35, 11004–11008. [Google Scholar] [CrossRef]
  75. Hu, J.; Pohl, A.; Wang, S.; Rothe, J.; Fichtner, M. Additive effects of LiBH4 and ZrCoH3 on the hydrogen sorption of the Li-Mg-NH hydrogen storage system. J. Phys. Chem. C 2012, 116, 20246–20253. [Google Scholar] [CrossRef]
  76. Lv, L.; Lin, J.; Yang, G.; Ma, Z.; Xu, L.; He, X.; Han, X.; Liu, W. Hydrogen storage performance of LaNi3.95Al0.75Co0.3 alloy with different preparation methods. Prog. Nat. Sci. Mater. 2022, 32, 206–214. [Google Scholar] [CrossRef]
  77. Ali, N.A.; Ismail, M.; Nasef, M.M.; Jalil, A.A. Enhanced hydrogen storage properties of NaAlH4 with the addition of CoTiO3 synthesised via a solid-state method. J. Alloys Compd. 2023, 934, 167932. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram showing the working principles of the Sieverts apparatus for determining hydrogen uptake/release.
Figure 1. Schematic diagram showing the working principles of the Sieverts apparatus for determining hydrogen uptake/release.
Materials 16 04056 g001
Figure 3. Isothermal desorption kinetics of pure LiAlH4, milled LiAlH4 and LiAlH4 + x wt.% LaCoO3 (x = 5, 10, 15 and 20).
Figure 3. Isothermal desorption kinetics of pure LiAlH4, milled LiAlH4 and LiAlH4 + x wt.% LaCoO3 (x = 5, 10, 15 and 20).
Materials 16 04056 g003
Figure 4. DSC traces of (a) milled LiAlH4, (b) LiAlH4 + 10 wt.% LaCoO3 at various heating rates and (c) DSC traces at 25 °C/min of milled LiAlH4 and LiAlH4 + 10 wt.% LaCoO3.
Figure 4. DSC traces of (a) milled LiAlH4, (b) LiAlH4 + 10 wt.% LaCoO3 at various heating rates and (c) DSC traces at 25 °C/min of milled LiAlH4 and LiAlH4 + 10 wt.% LaCoO3.
Materials 16 04056 g004
Figure 5. Activation energies of milled LiAlH4 and LiAlH4 + 10 wt.% LaCoO3 for (a) first stages and (b) second stages.
Figure 5. Activation energies of milled LiAlH4 and LiAlH4 + 10 wt.% LaCoO3 for (a) first stages and (b) second stages.
Materials 16 04056 g005
Figure 6. SEM images of (a) pure LiAlH4, (b) milled LiAlH4 and (c) LiAlH4 + 10 wt.% LaCoO3.
Figure 6. SEM images of (a) pure LiAlH4, (b) milled LiAlH4 and (c) LiAlH4 + 10 wt.% LaCoO3.
Materials 16 04056 g006
Figure 7. Particle size distributions of (a) pure LiAlH4, (b) milled LiAlH4 and (c) LiAlH4 + 10 wt.% LaCoO3.
Figure 7. Particle size distributions of (a) pure LiAlH4, (b) milled LiAlH4 and (c) LiAlH4 + 10 wt.% LaCoO3.
Materials 16 04056 g007
Figure 8. XRD pattern of (a) pure LiAlH4, (b) milled LiAlH4 and (c) LiAlH4 + 10 wt.% LaCoO3.
Figure 8. XRD pattern of (a) pure LiAlH4, (b) milled LiAlH4 and (c) LiAlH4 + 10 wt.% LaCoO3.
Materials 16 04056 g008
Figure 9. FTIR pattern of (a) pure LiAlH4, (b) milled LiAlH4 and (c) LiAlH4 + 10 wt.% LaCoO3.
Figure 9. FTIR pattern of (a) pure LiAlH4, (b) milled LiAlH4 and (c) LiAlH4 + 10 wt.% LaCoO3.
Materials 16 04056 g009
Figure 10. XRD pattern of (a) LiAlH4 + 10 wt.% LaCoO3 and (b) LiAlH4 + 30 wt.% LaCoO3 after desorption at 250 °C.
Figure 10. XRD pattern of (a) LiAlH4 + 10 wt.% LaCoO3 and (b) LiAlH4 + 30 wt.% LaCoO3 after desorption at 250 °C.
Materials 16 04056 g010
Table 1. Onset desorption temperature, hydrogen content and hydrogen loss from each sample.
Table 1. Onset desorption temperature, hydrogen content and hydrogen loss from each sample.
SamplesOnset Desorption Temperature
(1st Stage) (°C)
Onset Desorption Temperature
(2nd Stage) (°C)
Hydrogen Content
(wt.%)
Hydrogen Loss
(wt.%)
Pure LiAlH41511827.39-
Milled LiAlH41471777.090.30
LiAlH4 + 5 wt.% LaCoO3981536.011.38
LiAlH4 + 10 wt.% LaCoO3701565.861.53
LiAlH4 + 15 wt.% LaCoO31051645.721.67
LiAlH4 + 20 wt.% LaCoO31061695.671.72
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sazelee, N.; Ali, N.A.; Ismail, M.; Rather, S.-U.; Bamufleh, H.S.; Alhumade, H.; Taimoor, A.A.; Saeed, U. Enhancement of the Desorption Properties of LiAlH4 by the Addition of LaCoO3. Materials 2023, 16, 4056. https://doi.org/10.3390/ma16114056

AMA Style

Sazelee N, Ali NA, Ismail M, Rather S-U, Bamufleh HS, Alhumade H, Taimoor AA, Saeed U. Enhancement of the Desorption Properties of LiAlH4 by the Addition of LaCoO3. Materials. 2023; 16(11):4056. https://doi.org/10.3390/ma16114056

Chicago/Turabian Style

Sazelee, Noratiqah, Nurul Amirah Ali, Mohammad Ismail, Sami-Ullah Rather, Hisham S. Bamufleh, Hesham Alhumade, Aqeel Ahmad Taimoor, and Usman Saeed. 2023. "Enhancement of the Desorption Properties of LiAlH4 by the Addition of LaCoO3" Materials 16, no. 11: 4056. https://doi.org/10.3390/ma16114056

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop