Next Article in Journal
Construction and Performance of Superhydrophobic Surfaces for Rusted Iron Artifacts
Next Article in Special Issue
Effect of LaCoO3 Synthesized via Solid-State Method on the Hydrogen Storage Properties of MgH2
Previous Article in Journal / Special Issue
Ni0.6Zn0.4O Synthesised via a Solid-State Method for Promoting Hydrogen Sorption from MgH2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Boosting the Dehydrogenation Properties of LiAlH4 by Addition of TiSiO4

by
Nurul Yasmeen Yusnizam
,
Nurul Amirah Ali
,
Noratiqah Sazelee
and
Mohammad Ismail
*
Energy Storage Research Group, Faculty of Ocean Engineering Technology and Informatics, Universiti Malaysia Terengganu, Kuala Nerus 21030, Malaysia
*
Author to whom correspondence should be addressed.
Materials 2023, 16(6), 2178; https://doi.org/10.3390/ma16062178
Submission received: 7 February 2023 / Revised: 21 February 2023 / Accepted: 23 February 2023 / Published: 8 March 2023
(This article belongs to the Special Issue Advance Materials for Hydrogen Storage)

Abstract

:
Given its significant gravimetric hydrogen capacity advantage, lithium alanate (LiAlH4) is regarded as a suitable material for solid-state hydrogen storage. Nevertheless, its outrageous decomposition temperature and slow sorption kinetics hinder its application as a solid-state hydrogen storage material. This research’s objective is to investigate how the addition of titanium silicate (TiSiO4) altered the dehydrogenation behavior of LiAlH4. The LiAlH4–10 wt% TiSiO4 composite dehydrogenation temperatures were lowered to 92 °C (first-step reaction) and 128 °C (second-step reaction). According to dehydrogenation kinetic analysis, the TiSiO4-added LiAlH4 composite was able to liberate more hydrogen (about 6.0 wt%) than the undoped LiAlH4 composite (less than 1.0 wt%) at 90 °C for 2 h. After the addition of TiSiO4, the activation energies for hydrogen to liberate from LiAlH4 were lowered. Based on the Kissinger equation, the activation energies for hydrogen liberation for the two-step dehydrogenation of post-milled LiAlH4 were 103 and 115 kJ/mol, respectively. After milling LiAlH4 with 10 wt% TiSiO4, the activation energies were reduced to 68 and 77 kJ/mol, respectively. Additionally, the scanning electron microscopy images demonstrated that the LiAlH4 particles shrank and barely aggregated when 10 wt% of TiSiO4 was added. According to the X-ray diffraction results, TiSiO4 had a significant effect by lowering the decomposition temperature and increasing the rate of dehydrogenation of LiAlH4 via the new active species of AlTi and Si-containing that formed during the heating process.

1. Introduction

Concerns about the energy crisis and the environment have led to an increase in the proportion of renewable energy sources in the energy system, such as wind, hydro, and other types. The supply and utilization times are typically out of sync and have some geographical restrictions. The proper secondary energy must be selected to match them, and hydrogen is a promising alternative to meet the demands for clean and sustainable energy technologies [1]. Hydrogen energy is a perfect substitute for petroleum because of its high energy density and lack of carbon emissions [2,3,4,5,6,7].
Hydrogen can be stored in a variety of ways for application purposes. As of now, compressed gas has been the most popular technique. It can also be kept as a liquid at extremely low temperatures. Other methods of storing hydrogen are via the solid-state method through physisorption and chemisorption [8]. Storing hydrogen via the solid-state method is regarded as the most promising method. Based on the US Department of Energy (DOE), by 2025, the fuel cell materials should store 5.5 wt% (gravimetric) and 40 g L−1 (volumetric) of hydrogen [9]. Complex hydride has emerged as the most promising medium for solid-state hydrogen storage based on the DOE target due to its high hydrogen storage capacity. Additionally, based on previous research, storing hydrogen is promising in metal or complex hydrides via chemisorption. This technique involves absorbing and storing hydrogen in a metal powder made of either pure metal or an alloy. The metal/complex hydride material generates heat when hydrogen is absorbed into it, and heat is required for the metal/complex hydride to generate hydrogen again. As a result of the hydrogen’s strong bonding with the metal in some materials, it requires extreme temperatures, more than 400 °C for magnesium, as an example [10,11]. This is a disadvantage of the metal/complex hydride storage method [12,13,14]. Among metal or complex hydride materials, one of the best potential candidates is said to be lithium alanate (LiAlH4) [15,16,17,18,19,20]. LiAlH4 is thought to be a promising material for storing hydrogen in solid-state form as it has a greater capacity to store hydrogen compared to other complex hydrides. Table 1 compares the properties of LiAlH4 with other complex hydrides.
LiAlH4 decomposes in three steps, as shown in Equations (1)–(3),
LiAlH4 → 1/3Li3AlH6 + 2/3Al + H2 (about 5.3 wt% of H2 liberated)
Li3AlH6 → 3LiH + Al + 3/2H2 (about 2.6 wt% of H2 liberated)
LiH + Al → LiAl + 1/2H2 (about 2.6 wt% of H2 liberated)
However, its application is hampered by its high dehydrogenation temperature, slow dehydrogenation kinetics, and reversibility issues. According to previous studies [19,26,27,28], these challenges were addressed through several modifications, such as particle size reduction utilizing the ball milling method and doping with a catalyst. Some examples of doping with catalysts in the literature [19,29,30,31,32,33,34,35] include doping with metal halides, metal oxides, and carbon-based additives.
Complex metal oxide additives, especially secondary metal oxides, are a newly emerging research area to further enhance LiAlH4’s dehydrogenation abilities [36,37,38]. The study of the catalysis of Al2TiO5 on the dehydrogenation properties of LiAlH4 demonstrated that the onset temperatures dropped significantly after milling with 5 wt% Al2TiO5 [39]. The hydrogen began to be liberated at about 90 °C and 137 °C for both steps, respectively. Meanwhile, Zhai et al. [36] investigated how MnFe2O4 affected LiAlH4’s dehydrogenation properties and discovered that the temperature at which decomposition began was about 70 °C lower than it would have been otherwise. Moreover, the research conducted by Ismail et al. [40] proved that the onset desorption temperature of the two stages of the LiAlH4 was reduced to 80 °C and 120 °C, respectively, when doped with SrTiO3, as opposed to the as-obtained LiAlH4. The LiAlH4 dehydrogenation kinetics were also enhanced as the SrTiO3-doped LiAlH4 composite could liberate about 3.0 wt% of hydrogen at 90 °C in 20 min as opposed to the undoped composite’s 0.2 wt% in the same time frame.
However, more research is necessary to determine whether other secondary metal oxide-based catalysts could lower the decomposition temperature and improve LiAlH4 dehydrogenation kinetic efficiency. In this study, we have used another secondary metal oxide to boost the dehydrogenation properties of LiAlH4. To the best of our knowledge, no research has been conducted on enhancing LiAlH4’s ability to store hydrogen using TiSiO4 as a catalyst.

2. Materials and Methods

Sigma Aldrich (Burlington, MA, USA) provided LiAlH4 (purity 95.0%) and TiSiO4 (purity 99.8%) powders that were used without any changes. Every stage of the sample preparation process, including loading and weighing, was completed in an MBraun Unilab glove box to avoid oxidation. The sample (LiAlH4 + 10 wt% TiSiO4) was then directly milled in an NQM-0.4 planetary ball mill for 1 h (15 min milling, followed by three cycles of 2 min rest between rotations) at a rate of 400 rpm.
A Sievert-type apparatus (Advanced Materials Corporation, Pittsburgh, PA, USA) was used to study the onset dehydrogenation temperature and the kinetic property for the liberation of H2 from the undoped LiAlH4 and the TiSiO4-added LiAlH4 composite. The sample was heated from room temperature to 250 °C at a heating rate of 5 °C/min in a vacuum environment for the onset dehydrogenation temperature characterization. To determine the kinetics of H2 liberation, the sample was heated at a constant temperature of 90 °C while being exposed to an H2 pressure of 1.0 atm. Differential scanning calorimetry (DSC) from Setline STA: Simultaneous Thermal Analysis (SETARAM) was used to investigate the thermal properties of the samples. The DSC measurements were run at heating rates of 15 °C/min, 20 °C/min, 25 °C/min, and 30 °C/min. The argon gas flow rate was set to 50 mL/min, and the temperature range for this measurement was set to 300 °C. The microstructure of the sample before and after the milling process was examined using scanning electron microscopy (SEM; JEOL JSM 6360LA). The phase structure of the sample before and after the ball mill and after the dehydrogenation process was studied using an X-ray diffractometer (XRD, Rigaku Miniflex, Tokyo, Japan) and Fourier transform infrared (IR, Shimadzu Tracer-100, Tokyo, Japan).

3. Results and Discussion

Figure 1 depicts the results of the research on the onset thermal dehydrogenation and dehydrogenation kinetics processes for both added and undoped composites (LiAlH4 and 10 wt% TiSiO4 systems). Two distinguishing features of each composite’s dehydrogenation reaction are shown on the graph. As shown in Figure 1a, at about 146 °C, the as-obtained LiAlH4 has initiated the liberation of hydrogen, and further heating has led to the beginning of a second stage dehydrogenation process at about 180 °C. After 1 h of the LiAlH4 milling process, the first and second steps’ onset dehydrogenation temperatures drop to about 144 °C and 174 °C, respectively. This result demonstrates how the ball milling technique insignificantly influenced the dehydrogenation temperature of LiAlH4. Moreover, doping 10 wt% of TiSiO4 drastically lowered the decomposition temperatures for the two steps at 92 °C and 128 °C. Adding TiSiO4 as a catalyst proved to enhance the dehydrogenation performance of LiAlH4.
Figure 1b shows the isothermal dehydrogenation kinetics at 90 °C for doped and undoped samples. Less than 1.0 wt% of hydrogen was liberated in 80 min by the undoped sample. Under the same conditions, the hydrogen’s liberate capacity increased significantly to 5.7 wt% after 10 wt% TiSiO4 was added. As a result, the TiSiO4-added LiAlH4 sample desorbs at a rate that is, on average, 5 to 6 times higher than the undoped LiAlH4. According to this result, the catalyst addition of TiSiO4 improved the kinetics of LiAlH4 dehydrogenation. The outcome shows that the kinetics of LiAlH4 dehydrogenation were improved by the catalyst doping, as previous studies stated [15,41,42].
The catalytic effects of TiSiO4 on the thermal properties of LiAlH4 were validated using DSC measurements on added and undoped LiAlH4. Temperatures ranging from 25 °C to 300 °C, with an argon flow rate of 50 mL/min, were used for the investigation. Figure 2 displays the DSC curves for added and undoped LiAlH4 at a rate of 15 °C/min. The thermal characteristics of as-obtained LiAlH4 have four peaks. Two are exothermic, while the other two are endothermic. The interaction of hydroxyl impurities with LiAlH4’s surface caused the first exothermic peak to appear at 150 °C, while the melting of LiAlH4 caused the first endothermic peak to appear at 175 °C [37,40,43,44]. In the meantime, the dehydrogenation of liquid LiAlH4 causes the second exothermic peak (190 °C). Li3AlH6 decomposition (265 °C) was responsible for the second endothermic peak. These second exothermic and endothermic peaks were assigned to Equations (1) and (2), respectively. These similar peaks also occurred to the post-milled LiAlH4, but at lower temperatures.
The thermal event of LiAlH4 is lowered from four to two after the addition of TiSiO4. The exothermic event peak at around 120 °C seems to correspond to the decomposition of LiAlH4 (Equation (1)), and the endothermic event peak at around 176 °C is most likely to correspond to the decomposition of Li3AlH6 (Equation (2)). The first endothermic event associated with LiAlH4 melting has vanished from the DSC curve of TiSiO4-doped LiAlH4. The disappearance of the melting event is most likely owing to the fact that the decomposition temperature of the first stage of the doped sample is lower than the melting temperature of post-milled LiAlH4.
Different heating rates were measured using DSC to examine the impact of TiSiO4 addition on the activation energy (EA) for hydrogen desorbed from LiAlH4. The DSC curves for the added and undoped LiAlH4 samples at various heating rates (15, 20, 25, and 30 °C/min) are shown in Figure 3a,b. The Kissinger equation used to calculate the EA values of the dehydrogenation process of the 1 h milled LiAlH4 and LiAlH4–10 wt% TiSiO4 composite is as follows:
ln [β/Tp2] = −EA/RTp + A
where R is the gas constant, A is a linear constant, Tp is the peak temperature on the DSC dehydrogenation curves, EA is the activation energy, and β is the DSC heating rate. Consequently, the EA was determined using the graph’s slope, ln β / T p 2 vs. 1000/Tp. The Kissinger plots for the dehydrogenation of LiAlH4 (1st stage) and the dehydrogenation of Li3AlH6 (2nd stage) for both composites are shown in Figure 3c,d. Based on the slopes, EA for the post-milled LiAlH4 are 103 kJ/mol (1st stage dehydrogenation) and 115 kJ/mol (2nd stage dehydrogenation), respectively. Meanwhile, the EA for the first and second stages of dehydrogenation, respectively, are reduced to 68 kJ/mol and 77 kJ/mol for the added system with 10 wt% TiSiO4.
Figure 4 shows the morphology differences between the TiSiO4, added, and undoped LiAlH4 samples. Figure 4a shows the morphology for the as-obtained TiSiO4, which is a fine particle. Figure 4b depicts the as-obtained LiAlH4’s morphology, which is described as rough and asymmetrical in shape. The particle size of LiAlH4 was smaller but aggregated and inhomogeneous after a 1 h milling process (Figure 4c). The particle size of the 10 wt% TiSiO4-doped LiAlH4 sample (Figure 4d) has shrunk and is less aggregated compared to post-milled LiAlH4. The 10 wt% TiSiO4-added LiAlH4 composite has a larger surface area due to the smaller particle sizes. Previous research has shown that surface modification significantly improves the hydrogen storage properties of metals and complex hydride materials [34,45,46,47].
The particle size distribution of the TiSiO4, added, and undoped LiAlH4 samples were determined using the Image J software, as demonstrated in Figure 5. Based on the histogram shown in Figure 5a–d, the estimated average particle sizes for the as-obtained TiSiO4, as-obtained LiAlH4, post-milled LiAlH4, and LiAlH4 added with 10 wt% TiSiO4 were determined to be 6.10, 81.58, 61.71, and 28.13 µm, respectively. These findings demonstrated that the addition of 10 wt% TiSiO4 to LiAlH4 significantly reduced particle size. Previous studies by Ahmad et al. [22] and Cai et al. [48] stated the added LiAlH4 increased the rate of hydrogen diffusion, which resulted in fast dehydrogenation kinetics and low activation energy. Additionally, its smaller particle size results in a larger surface area and more grain boundaries [36,38,49].
Figure 6 shows the XRD pattern of the as-obtained and post-milled LiAlH4, TiSiO4-doped LiAlH4 composite, and as-obtained TiSiO4 sample. The sample of as-obtained TiSiO4 shows the high purity of TiSiO4 [50]. For the as-obtained LiAlH4 sample, as shown in Figure 6a, the peaks of LiAlH4 are dominant, and there are no other peaks detected, proving that the as-obtained LiAlH4 sample is pure, as stated by the supplier. Furthermore, as shown in Figure 6b, even after the ball milling process of 1 h, LiAlH4 initial phase remains unchanged. According to this outcome and the findings of Sazelee et al. [15] and Balema et al. [51], LiAlH4 was stable even after being subjected to ball milling. Figure 6c shows the XRD pattern of the LiAlH4–10 wt% TiSiO4 composite. In addition to the majority peaks of LiAlH4, there is a new peak corresponding to the Al for the TiSiO4-doped LiAlH4 after 1 h milling. The appearance of Al peaks indicates that with the presence of TiSiO4, hydrogen is slightly released from LiAlH4 after milling (Equation (1)). However, the peaks that correspond to Li3AlH6 could not be detected in this pattern. Additionally, due to the small amount of TiSiO4 used or the fact that TiSiO4 became amorphous after 1 h of the milling process, the peaks of TiSiO4 could not be detected with an XRD pattern [39,49,52].
Figure 7 depicts the FTIR spectrum used to investigate the effect of TiSiO4 as a catalyst on the LiAlH4 infrared spectroscopy band and to confirm the presence of Li3AlH6 in the TiSiO4-doped LiAlH4 sample after milling. In the previous studies [15,37,53], LiAlH4 was represented by peaks with two bending modes between 800 and 900 cm−1 and two stretching modes between 1600 and 1800 cm−1. Figure 7a,b show these peaks in the as-obtained and post-milled LiAlH4 samples. After 1 h of milling (Figure 7c), a new peak at 1403 cm−1 was discovered in the TiSiO4-doped LiAlH4 sample. This new peak corresponds to the Al–H stretching mode of Li3AlH6. Although the peak of Li3AlH6 could not be detected in the XRD pattern (Figure 6c), the appearance of the Al–H stretching mode of Li3AlH6 in the FTIR result proves that the TiSiO4-doped LiAlH4 sample slightly released hydrogen during the milling process (Equation (1)).
The XRD analysis helped to identify the catalytic and reaction mechanism behind the improvement of the dehydrogenation properties of TiSiO4-added LiAlH4. Figure 8 shows the results of 10 wt% and 20 wt% TiSiO4-added LiAlH4 composites after the dehydrogenation process at 250 °C. Due to the small amount of catalyst used, it may be challenging to identify the active species in the 10 wt% TiSiO4-added LiAlH4, so 20 wt% TiSiO4 was added to LiAlH4. The presence of the Al and LiH peaks in both samples (Figure 8a) indicates that LiAlH4 has been completely dehydrogenated (Equation (2)). In addition to Al and LiH peaks, the peaks that correspond to the AlTi species can also be detected in the dehydrogenation sample. However, the peaks that correspond to the Si or Si-containing patterns could not be observed after the dehydrogenation process. This may be because the Si or Si-containing patterns were in an amorphous state. According to this research, it was hypothesized that the enhanced dehydrogenation behaviors of LiAlH4 were attributed to the formation of AlTi and Si or Si-containing species. Furthermore, there were no new peaks found in the XRD patterns of the desorbed 20 wt% TiSiO4-added LiAlH4 composite (Figure 8b), which are similar to those of the 10 wt% TiSiO4-added LiAlH4 composite (Figure 8a). The peak of Al, LiH, and AlTi did not change, and no other patterns were discovered in the dehydrogenation composite, possibly because these patterns were in an amorphous state when the dehydrogenation process was complete.
Previous studies have shown that the in-situ emergence of AlTi helps to improve the dehydrogenation behavior of LiAlH4 [39,49,54]. Strong surface reactions between the Ti atom and LiAlH4 caused a reduction in the H binding energy. The correlation between the charge transfer change between Al and H and the rapid kinetic efficiency of LiAlH4 may explain the decrease in binding energy. Although the Si or Si-containing phase was not found, these species likewise play a particular role in enhancing the dehydrogenation behavior of LiAlH4. For example, the Si-containing species have been proven to play a significant role in enhancing the hydrogen storage properties of MgH2 [55]. Another study proved that SiC could enhance the hydrogen storage properties of MgH2 by reducing grain size and increasing defect concentration in MgH2 particles [56,57]. However, more research is still needed, such as using high-resolution transmission electron microscopy and X-ray photoelectron spectroscopy to clarify the exact catalytic role of TiSiO4 on the hydrogen storage properties of LiAlH4.

4. Conclusions

In conclusion, the addition of TiSiO4 improved the dehydrogenation properties of LiAlH4. The introduction of TiSiO4 to LiAlH4 reduced the dehydrogenation temperatures for the first and second stages (start decomposing at 92 °C and 128 °C, respectively). Dehydrogenation kinetic analysis at 90 °C revealed that after 2 h, the TiSiO4-added LiAlH4 composite was able to liberate more hydrogen (about 6.00 wt%) than the undoped LiAlH4 composite (less than 1.00 wt%). The EA for hydrogen liberated from LiAlH4 was reduced after TiSiO4 was added. According to the Kissinger equation, the EA for hydrogen liberated in the two-step dehydrogenation of post-milled LiAlH4 was 103 and 115 kJ/mol, respectively. The EA were lowered to 68 and 77 kJ/mol, respectively, after milling LiAlH4 with 10 wt% of TiSiO4. The addition of 10 wt% of TiSiO4 also caused the LiAlH4 particles to shrink and become less aggregated. According to the XRD results, TiSiO4 significantly enhanced the dehydrogenation properties of LiAlH4 by forming active AlTi and Si-containing species during the heating process.

Author Contributions

N.Y.Y., investigations, formal analysis, and writing—original draft preparation; N.A.A., writing—review and editing; N.S., writing—review and editing; M.I., supervision, writing—original draft preparation. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Universiti Malaysia Terengganu through Golden Goose Research Grant (GGRG) (VOT 55190).

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

Universiti Malaysia Terengganu is fully acknowledged for supporting this study through Golden Goose Research Grant (GGRG) (VOT 55190). N.A.A. and N.S. are grateful for the scholarship provided by Universiti Malaysia Terengganu (SIPP and BUMT).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ye, Y.; Lu, J.; Ding, J.; Wang, W.; Yan, J. Numerical simulation on the storage performance of a phase change materials based metal hydride hydrogen storage tank. Appl. Energy 2020, 278, 115682. [Google Scholar] [CrossRef]
  2. Huang, Y.; Cheng, Y.; Zhang, J. A review of high density solid hydrogen storage materials by pyrolysis for promising mobile applications. Ind. Eng. Chem. Res. 2021, 60, 2737–2771. [Google Scholar] [CrossRef]
  3. Rivard, E.; Trudeau, M.; Zaghib, K. Hydrogen storage for mobility: A review. Materials 2019, 12, 1973. [Google Scholar] [CrossRef] [Green Version]
  4. Moore, J.; Shabani, B. A critical study of stationary energy storage policies in Australia in an international context: The role of hydrogen and battery technologies. Energies 2016, 9, 674. [Google Scholar] [CrossRef]
  5. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. J. Nat. 2001, 414, 353–358. [Google Scholar] [CrossRef] [PubMed]
  6. Mustafa, N.S.; Yahya, M.S.; Itam Sulaiman, N.N.; Yap, F.A.H.; Ismail, M. Enhanced the hydrogen storage properties and reaction mechanisms of 4MgH2+ LiAlH4 composite system by addition with TiO2. Int. J. Energy Res. 2021, 45, 21365–21374. [Google Scholar] [CrossRef]
  7. Salehabadi, A.; Umar, M.F.; Ahmad, A.; Ahmad, M.I.; Ismail, N.; Rafatullah, M. Carbon-based nanocomposites in solid-state hydrogen storage technology: An overview. Int. J. Energy. Res. 2020, 44, 11044–11058. [Google Scholar] [CrossRef]
  8. Egeland-Eriksen, T.; Hajizadeh, A.; Sartori, S. Hydrogen-based systems for integration of renewable energy in power systems: Achievements and perspectives. Int. J. Hydrog. Energy 2021, 46, 31963–31983. [Google Scholar] [CrossRef]
  9. DOE Technical Targets for Onboard Hydrogen Storage for Light-Duty Vehicles. Available online: https://www.energy.gov/eere/fuelcells/doe-technical-targets-onboard-hydrogen-storage-light-duty-vehicles (accessed on 19 February 2023).
  10. Aziz, M.; Wijayanta, A.T.; Nandiyanto, A.B.D. Ammonia as effective hydrogen storage: A review on production, storage and utilization. Energies 2020, 13, 3062. [Google Scholar] [CrossRef]
  11. Ouyang, L.; Liu, F.; Wang, H.; Liu, J.; Yang, X.-S.; Sun, L.; Zhu, M. Magnesium-based hydrogen storage compounds: A review. J. Alloys Compd. 2020, 832, 154865. [Google Scholar] [CrossRef]
  12. Ouyang, L.; Chen, K.; Jiang, J.; Yang, X.-S.; Zhu, M. Hydrogen storage in light-metal based systems: A review. J. Alloy. Compd. 2020, 829, 154597. [Google Scholar] [CrossRef]
  13. Ali, N.A.; Ismail, M. Advanced hydrogen storage of the Mg–Na–Al system: A review. J. Magnes. Alloy 2021, 9, 1111–1122. [Google Scholar] [CrossRef]
  14. Ali, N.A.; Sazelee, N.A.; Ismail, M. An overview of reactive hydride composite (RHC) for solid-state hydrogen storage materials. Int. J. Hydrog. Energy 2021, 46, 31674–31698. [Google Scholar] [CrossRef]
  15. Sazelee, N.A.; Yahya, M.S.; Idris, N.H.; Din, M.M.; Ismail, M. Desorption properties of LiAlH4 doped with LaFeO3 catalyst. Int. J. Hydrog. Energy 2019, 44, 11953–11960. [Google Scholar] [CrossRef]
  16. Ali, N.A.; Yahya, M.S.; Mustafa, N.S.; Sazelee, N.A.; Idris, N.H.; Ismail, M. Modifying the hydrogen storage performances of NaBH4 by catalyzing with MgFe2O4 synthesized via hydrothermal method. Int. J. Hydrog. Energy 2019, 44, 6720–6727. [Google Scholar] [CrossRef]
  17. Halim Yap, F.A.; Ali, N.A.; Idris, N.H.; Ismail, M. Catalytic effect of MgFe2O4 on the hydrogen storage properties of Na3AlH6–LiBH4 composite system. Int. J. Hydrog. Energy 2018, 43, 20882–20891. [Google Scholar] [CrossRef]
  18. Juahir, N.; Mustafa, N.S.; Yap, F.A.H.; Ismail, M. Study on the hydrogen storage properties and reaction mechanism of NaAlH4–Mg(BH4)2 (2:1) with and without TiF3 additive. Int. J. Hydrog. Energy 2015, 40, 7628–7635. [Google Scholar] [CrossRef]
  19. 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]
  20. Zheng, S.; Fang, F.; Zhou, G.; Chen, G.; Ouyang, L.; Zhu, M.; Sun, D. Hydrogen storage properties of space-confined NaAlH4 nanoparticles in ordered mesoporous silica. Chem. Mater. 2008, 20, 3954–3958. [Google Scholar] [CrossRef]
  21. Sigma-Aldrich. Available online: https://www.sigmaaldrich.com/MY/en/campaigns/connected-lab-instrumentation (accessed on 20 February 2023).
  22. Ahmad, M.A.N.; Sazelee, N.A.; Ali, N.A.; Ismail, M. Enhancing the dehydrogenation properties of LiAlH4 using K2NiF6 as additive. Int. J. Hydrog. Energy 2022, 47, 24843–24851. [Google Scholar] [CrossRef]
  23. Fu, J. Advanced Doping Techniques and Dehydrogenation Properties of Transition Metal-Doped LiAlH4 for Fuel Cell Systems. Ph.D Thesis, Dresden University of Technology, Dresden, Germany, 2015. [Google Scholar]
  24. Orimo, S.-I.; Nakamori, Y.; Eliseo, J.R.; Züttel, A.; Jensen, C.M. Complex hydrides for hydrogen storage. Chem. Rev. 2007, 107, 4111–4132. [Google Scholar] [CrossRef]
  25. Liu, Y.; Liang, C.; Zhou, H.; Gao, M.; Pan, H.; Wang, Q. A novel catalyst precursor K2TiF6 with remarkable synergetic effects of K, Ti and F together on reversible hydrogen storage of NaAlH4. Chem. Commun. 2011, 47, 1740–1742. [Google Scholar] [CrossRef]
  26. Andreasen, A.; Vegge, T.; Pedersen, A.S. Dehydrogenation kinetics of as-received and ball-milled LiAlH4. J. Solid State Chem. 2005, 178, 3672–3678. [Google Scholar] [CrossRef]
  27. Liu, S.-S.; Sun, L.-X.; Zhang, Y.; Xu, F.; Zhang, J.; Chu, H.-L.; Fan, M.Q.; Zhang, T.; Song, X.Y.; Grolier, J.P. Effect of ball milling time on the hydrogen storage properties of TiF3-doped LiAlH4. Int. J. Hydrog. Energy 2009, 34, 8079–8085. [Google Scholar] [CrossRef]
  28. Liu, X.; McGrady, G.S.; Langmi, H.W.; Jensen, C.M. Facile cycling of Ti-doped LiAlH4 for high performance hydrogen storage. J. Am. Chem. Soc. 2009, 131, 5032–5033. [Google Scholar] [CrossRef] [PubMed]
  29. Li, Z.; Li, P.; Wan, Q.; Zhai, F.; Liu, Z.; Zhao, K.; Wang, L.; Lü, S.; Zou, L.; Qu, X.; et al. Dehydrogenation improvement of LiAlH4 catalyzed by Fe2O3 and Co2O3 nanoparticles. J. Phys. Chem. C 2013, 117, 18343–18352. [Google Scholar] [CrossRef]
  30. Resan, M.; Hampton, M.D.; Lomness, J.K.; Slattery, D.K. Effects of various catalysts on hydrogen release and uptake characteristics of LiAlH4. Int. J. Hydrog. Energy 2005, 30, 1413–1416. [Google Scholar] [CrossRef]
  31. Balema, V.P.; Wiench, J.W.; Dennis, K.W.; Pruski, M.; Pecharsky, V.K. Titanium catalyzed solid-state transformations in LiAlH4 during high-energy ball-milling. J. Alloys Compd. 2001, 329, 108–114. [Google Scholar] [CrossRef] [Green Version]
  32. Shenglin, L.; Qiuhua, M.; Xueping, Z.; Xin, F.; Guo, X.; Jiaojiao, Z. Influences of Y2O3 doping on hydrogen release property of LiAlH4. Rare Metal Mat. Eng. 2014, 43, 287–290. [Google Scholar] [CrossRef] [Green Version]
  33. 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]
  34. Varin, R.A.; Zbroniec, L. Fast and slow dehydrogenation of ball milled lithium alanate (LiAlH4) catalyzed with manganese chloride (MnCl2) as compared to nanometric nickel catalyst. J. Alloys Compd. 2011, 509, S736–S739. [Google Scholar] [CrossRef]
  35. Sazelee, N.A.; Ismail, M. Recent advances in catalyst-enhanced LiAlH4 for solid-state hydrogen storage: A review. Int. J. Hydrog. Energy 2021, 46, 9123–9141. [Google Scholar] [CrossRef]
  36. 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]
  37. Ismail, M.; Zhao, Y.; Yu, X.; Nevirkovets, I.; Dou, S. Significantly improved dehydrogenation of LiAlH4 catalysed with TiO2 nanopowder. Int. J. Hydrog. Energy 2011, 36, 8327–8334. [Google Scholar] [CrossRef]
  38. Xuanhui, Q.; Ping, L.; Zhang, L.; Ahmad, M. Hydrogen sorption improvement of LiAlH4 catalyzed by Nb2O5 and Cr2O3 nanoparticles. J. Phys. Chem. C 2011, 115, 13088–13099. [Google Scholar]
  39. Ismail, M.; Ali, N.A.; Sazelee, N.A.; Suwarno, S. Catalytic effect of Al2TiO5 on the dehydrogenation properties of LiAlH4. Int. J. Hydrog. Energy 2022, 47, 31903–31910. [Google Scholar] [CrossRef]
  40. 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]
  41. Li, L.; An, C.; Wang, Y.; Xu, Y.; Qiu, F.; Wang, Y.; Jiao, L.; Yuan, H. Enhancement of the H2 desorption properties of LiAlH4 doping with NiCo2O4 nanorods. Int. J. Hydrog. Energy 2014, 39, 4414–4420. [Google Scholar] [CrossRef]
  42. 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.; et al. Improved reversible hydrogen storage of LiAlH4 by nano-sized TiH2. Int. J. Hydrog. Energy 2013, 38, 2770–2777. [Google Scholar] [CrossRef]
  43. Ismail, M.; Zhao, Y.; Yu, X.B.; Ranjbar, A.; Dou, S.X. Improved hydrogen desorption in lithium alanate by addition of SWCNT–metallic catalyst composite. Int. J. Hydrog. Energy 2011, 36, 3593–3599. [Google Scholar] [CrossRef]
  44. Ismail, M.; Zhao, Y.; Yu, X.B.; Dou, S.X. Effects of NbF5 addition on the hydrogen storage properties of LiAlH4. Int. J. Hydrog. Energy 2010, 35, 2361–2367. [Google Scholar] [CrossRef]
  45. Czujko, T.; Oleszek, E.E.; Szot, M. New aspects of MgH2 morphological and structural changes during high-energy ball milling. Materials 2020, 13, 4550. [Google Scholar] [CrossRef]
  46. Révész, Á.; Gajdics, M. Improved H-storage performance of novel Mg-based nanocomposites prepared by high-energy ball milling: A review. Energies 2021, 14, 6400. [Google Scholar] [CrossRef]
  47. Varin, R.A.; Czujko, T.; Wronski, Z. Particle size, grain size and γ-MgH2 effects on the desorption properties of nanocrystalline commercial magnesium hydride processed by controlled mechanical milling. Nanotechnology 2006, 17, 3856. [Google Scholar] [CrossRef]
  48. Cai, J.; Zang, L.; Zhao, L.; Liu, J.; Wang, Y. Dehydrogenation characteristics of LiAlH4 improved by in-situ formed catalysts. J. Energy Chem. 2016, 25, 868–873. [Google Scholar] [CrossRef]
  49. 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] [PubMed]
  50. Devrim, Y.; Erkan, S.; Bac, N.; Eroglu, I. Improvement of PEMFC performance with Nafion/inorganic nanocomposite membrane electrode assembly prepared by ultrasonic coating technique. Int. J. Hydrog. Energy 2012, 37, 16748–16758. [Google Scholar] [CrossRef]
  51. Balema, V.P.; Pecharsky, V.K.; Dennis, K.W. Solid state phase transformations in LiAlH4 during high-energy ball-milling. J. Alloys Compd. 2000, 313, 69–74. [Google Scholar] [CrossRef] [Green Version]
  52. Xia, Y.; Zhang, H.; Sun, Y.; Sun, L.; Xu, F.; Sun, S.; Zhang, G.; Huang, P.; Du, Y.; Wang, J.; et al. Dehybridization effect in improved dehydrogenation of LiAlH4 by doping with two-dimensional Ti3C2. Mater. Today Nano 2019, 8, 100054. [Google Scholar] [CrossRef]
  53. Rafi ud, d.; Zhang, L.; Ping, L.; Xuanhui, Q. Catalytic effects of nano-sized TiC additions on the hydrogen storage properties of LiAlH4. J. Alloy. Compd. 2010, 508, 119–128. [Google Scholar] [CrossRef]
  54. Zang, L.; Cai, J.; Zhao, L.; Gao, W.; Liu, J.; Wang, Y. Improved hydrogen storage properties of LiAlH4 by mechanical milling with TiF3. J. Alloy. Compd. 2015, 647, 756–762. [Google Scholar] [CrossRef]
  55. Ismail, M.; Yahya, M.S.; Sazelee, N.A.; Ali, N.A.; Yap, F.A.H.; Mustafa, N.S. The effect of K2SiF6 on the MgH2 hydrogen storage properties. J. Magnes. Alloy 2020, 8, 832–840. [Google Scholar] [CrossRef]
  56. Ranjbar, A.; Guo, Z.; Yu, X.; Attard, D.; Calka, A.; Liu, H.-K. Effects of SiC nanoparticles with and without Ni on the hydrogen storage properties of MgH2. Int. J. Hydrog. Energy 2009, 34, 7263–7268. [Google Scholar] [CrossRef]
  57. Kurko, S.; Rašković, Ž.; Novaković, N.; Mamula, B.P.; Jovanović, Z.; Baščarević, Z.; Novaković, J.G.; Matović, L. Hydrogen storage properties of MgH2 mechanically milled with α and β SiC. Int. J. Hydrog. Energy 2011, 36, 549–554. [Google Scholar] [CrossRef]
Figure 1. (a) TPD curve of as-obtained LiAlH4, post-milled LiAlH4, and LiAlH4 added with 10 wt% TiSiO4 at 250 °C and (b) the dehydrogenation kinetics performance at a constant temperature of 90 °C for the added and undoped composites.
Figure 1. (a) TPD curve of as-obtained LiAlH4, post-milled LiAlH4, and LiAlH4 added with 10 wt% TiSiO4 at 250 °C and (b) the dehydrogenation kinetics performance at a constant temperature of 90 °C for the added and undoped composites.
Materials 16 02178 g001
Figure 2. DSC data recorded at a rate of 15 °C/min for as-obtained LiAlH4, post-milled LiAlH4, and LiAlH4 added with 10 wt% TiSiO4.
Figure 2. DSC data recorded at a rate of 15 °C/min for as-obtained LiAlH4, post-milled LiAlH4, and LiAlH4 added with 10 wt% TiSiO4.
Materials 16 02178 g002
Figure 3. DSC dehydrogenation curves at different heating rates for (a) post-milled LiAlH4, (b) LiAlH4–10 wt% TiSiO4 and Kissinger’s plot for the (c) first and (d) second dehydrogenation stage of post-milled LiAlH4 and LiAlH4 added with 10 wt% TiSiO4.
Figure 3. DSC dehydrogenation curves at different heating rates for (a) post-milled LiAlH4, (b) LiAlH4–10 wt% TiSiO4 and Kissinger’s plot for the (c) first and (d) second dehydrogenation stage of post-milled LiAlH4 and LiAlH4 added with 10 wt% TiSiO4.
Materials 16 02178 g003
Figure 4. The morphological structures of (a) as-obtained TiSiO4, (b) as-obtained LiAlH4, (c) post-milled LiAlH4, and (d) LiAlH4 added with 10 wt% TiSiO4.
Figure 4. The morphological structures of (a) as-obtained TiSiO4, (b) as-obtained LiAlH4, (c) post-milled LiAlH4, and (d) LiAlH4 added with 10 wt% TiSiO4.
Materials 16 02178 g004
Figure 5. Particle size distribution histograms of the (a) as-obtained TiSiO4, (b) as-obtained LiAlH4, (c) post-milled LiAlH4, and (d) LiAlH4 added with 10 wt% TiSiO4.
Figure 5. Particle size distribution histograms of the (a) as-obtained TiSiO4, (b) as-obtained LiAlH4, (c) post-milled LiAlH4, and (d) LiAlH4 added with 10 wt% TiSiO4.
Materials 16 02178 g005
Figure 6. XRD results of (a) as-obtained LiAlH4, (b) post-milled LiAlH4, (c) LiAlH4 added with 10 wt% of TiSiO4, and (d) as-obtained TiSiO4.
Figure 6. XRD results of (a) as-obtained LiAlH4, (b) post-milled LiAlH4, (c) LiAlH4 added with 10 wt% of TiSiO4, and (d) as-obtained TiSiO4.
Materials 16 02178 g006
Figure 7. FTIR pattern of the (a) as-obtained LiAlH4, (b) post-milled LiAlH4, and (c) LiAlH4 added with 10 wt% of TiSiO4.
Figure 7. FTIR pattern of the (a) as-obtained LiAlH4, (b) post-milled LiAlH4, and (c) LiAlH4 added with 10 wt% of TiSiO4.
Materials 16 02178 g007
Figure 8. XRD pattern of the dehydrogenation at 250 °C of the added LiAlH4 with (a) 10 wt% and (b) 20 wt% of TiSiO4.
Figure 8. XRD pattern of the dehydrogenation at 250 °C of the added LiAlH4 with (a) 10 wt% and (b) 20 wt% of TiSiO4.
Materials 16 02178 g008
Table 1. Properties of LiAlH4 and other complex hydrides [21,22,23,24,25].
Table 1. Properties of LiAlH4 and other complex hydrides [21,22,23,24,25].
MaterialsCost of Material
($ g−1)
Desorption KineticsGravimetric
(wt%)
Volumetric
(g L−1)
LiAlH43.7~0.04 wt%
(within 180 min, 90 °C)
10.596.7
NaAlH410.8~0.01 wt%
(within 150 min, 140 °C)
7.494.8
KBH41.41NA7.487.1
NaNH26.19NA5.271.9
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

Yusnizam, N.Y.; Ali, N.A.; Sazelee, N.; Ismail, M. Boosting the Dehydrogenation Properties of LiAlH4 by Addition of TiSiO4. Materials 2023, 16, 2178. https://doi.org/10.3390/ma16062178

AMA Style

Yusnizam NY, Ali NA, Sazelee N, Ismail M. Boosting the Dehydrogenation Properties of LiAlH4 by Addition of TiSiO4. Materials. 2023; 16(6):2178. https://doi.org/10.3390/ma16062178

Chicago/Turabian Style

Yusnizam, Nurul Yasmeen, Nurul Amirah Ali, Noratiqah Sazelee, and Mohammad Ismail. 2023. "Boosting the Dehydrogenation Properties of LiAlH4 by Addition of TiSiO4" Materials 16, no. 6: 2178. https://doi.org/10.3390/ma16062178

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