Next Article in Journal
Acknowledgment to Reviewers of Reactions in 2020
Next Article in Special Issue
Low-Temperature Methane Partial Oxidation over Pd Supported on CeO2: Effect of the Preparation Method and Precursors
Previous Article in Journal
Hydrocracking of Polyethylene to Jet Fuel Range Hydrocarbons over Bifunctional Catalysts Containing Pt- and Al-Modified MCM-48
Previous Article in Special Issue
Smart Designs of Anti-Coking and Anti-Sintering Ni-Based Catalysts for Dry Reforming of Methane: A Recent Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Nanoconfinement on the Hydrogen Release Processes from Sodium Alanate

1
Institute of Chemistry, University of Tartu, Ravila 14a, 50411 Tartu, Estonia
2
Department of Applied Physics, KTH Royal Institute of Technology, 10691 Stockholm, Sweden
*
Author to whom correspondence should be addressed.
Reactions 2021, 2(1), 1-9; https://doi.org/10.3390/reactions2010001
Submission received: 7 December 2020 / Revised: 11 January 2021 / Accepted: 14 January 2021 / Published: 18 January 2021
(This article belongs to the Special Issue Hydrogen Production and Storage)

Abstract

:
Sodium alanate (NaAlH4) is a prospective H2 storage material for stationary and mobile applications, as NaAlH4 contains 7.4 wt% of H2, and it is possible to do multiple H2 release and accumulation cycles. Nanoconfinement is a potential solution to enhance the H2 release properties of NaAlH4. To optimize the supporting material and the synthesis method used for the nanoconfinement of NaAlH4, a better understanding of the influence of nanoconfinement on the H2 release processes is necessary. Thus, the H2 release from bulk, purely nanoconfined, and intermediate NaAlH4 is measured at different temperature ramp rates, and the characteristic parameters for each hydrogen release process are determined. Activation energies for each process are determined using the Kissinger method, and the effect of nanoconfinement on the activation energies is analysed. The impact of nanoconfinement on the H2 release processes from NaAlH4 and the limitations of each process in case of bulk and nanoconfined NaAlH4 are presented and discussed. Nanoconfinement of NaAlH4 decreases activation energies of the initial reversible H2 release steps to between 30 and 45 kJ mol−1 and increased the activation energy of the last irreversible H2 release step to over 210 kJ mol−1.

1. Introduction

Hydrogen is a potential energy carrier, as H2 can be produced directly from water through electrolysis, H2 can be directly converted into electrical energy with up to 60% efficiency with fuel cells, and the only by-product during conversion into electricity is water [1,2]. Thus, using H2 to accumulate cheap electricity from renewable sources during high winds and/or intense solar radiation would help balance the power grid and be a viable alternative to electric battery cars. One of the hindrances to the application of a hydrogen-based energy economy is the lack of a cheap and safe hydrogen storage technology. The current commercial standard for hydrogen storage in mobile applications is storing hydrogen in pressurized tanks of up to 700 bar [3]. As the compression of H2 to such pressures, p, consumes at least 10% [4] of the stored energy, such high pressures have inherent hazards, and even then, the H2 energy density is only 1.31 kWh L−1 at 25 °C, alternative H2 storage methods are of high interest.
The storage of hydrogen in a chemically bound form inside complex metal hydrides offers the advantage of high H2 density, storage at ambient temperature, T, and pH2 = 1 bar, and potential cyclability through the application of pH2 [5]. One of the most promising complex metal hydrides for H2 storage is sodium alanate (NaAlH4), which incorporates 7.4 wt% of H2. The release of H2 from the bulk phase can be described through three decomposition steps (Equations (1)–(3)). The thermodynamical equilibrium T of decomposition (i.e., the release of H2) for the first two decomposition reactions (Equations (1) and (2)) of NaAlH4 are at comparatively low temperatures (80 and 150 °C, respectively) [6]. Alas, the release of H2 from the bulk phase is kinetically limited in the solid phase, and thus, the first H2 release step from bulk NaAlH4 starts at 183 °C (Equation (1)) with the melting of NaAlH4. H2 is released (from Equation (1) to Equation (3)) over a wide T range (Figure 1 and Reference [7]). The third H2 release step (Equation (3)) is of no use for H2 storage as the reverse reaction has not been achieved at pH2 and T feasible for current gas storage systems. Thus, the main problems of a NaAlH4-based H2 storage system are the high temperature of H2 release caused by the kinetic limitation, the irreversibility of the third decomposition step (Equation (3)), and the segregation of decomposition products (Al, NaH, and Na3AlH6) into separate phases, which decreases the cycling lifetime of NaAlH4-based H2 storage systems.
3NaAlH4(l) → Na3AlH6(s) + 2Al(s) + 3H2Tstart = 183 °C
Na3AlH6(s) → 3NaH(s) + Al(s) + 1½H2Tstart = 235 °C
3NaH(s) → 3Na(l) + 1½H2Tstart = 270 °C
The T of H2 release from NaAlH4 can be lowered through nanoscaling, where the higher surface area of sub-micrometer particles promotes the release of H2 from NaAlH4 at a T closer to that of thermodynamical equilibrium through the elimination of kinetical limitation present in case of bulk material, i.e., diffusion of H2 inside the bulk phase. There are two main methods for nanoscaling NaAlH4: reducing the average particle size through ball-milling or nanoconfining nanoscale particles unto a supporting porous material [7,8,9,10,11]. Whilst ball-milling is relatively easy to perform, and the T of H2 release is considerably lowered during multiple hydrogenation/dehydrogenation cycles, the nanoscale particles agglomerate together at increased T, and the improved H2 release properties are lost over multiple dehydrogenation/hydrogenation cycles [12]. By using a supporting porous material for nanoconfinement of NaAlH4, the starting T of H2 release is lowered to ambient conditions [8,13]. In addition, the agglomeration of NaAlH4 and the segregation of decomposition products during dehydrogenation into separate phases is limited by the confinement of NaAlH4, e.g., during dehydrogenation of bulk NaAlH4, particles of Al form in the micrometer size. In contrast, the size of Al particles formed during the dehydrogenation of confined NaAlH4 is lowered to tens of nanometers [13,14]. Nanoscale NaAlH4 does not release H2 according to Equations (1)–(3), where the direct decomposition of NaAlH4 to NaH has been shown by Gao et al. [7], and the H2 release curves are inherently different, whereas the H2 released from different decomposition steps (Equations (1)–3)) from nanoconfined NaAlH4 are visually inseparable [7,8,13].
To fine-tune the H2 release properties of nanoconfined NaAlH4, it is essential to understand the energetical barrier associated with each H2 release step. It has been shown by Baldé et al. [8] that the decrease in NaAlH4 particle size decreases the activation energy, Ea, of the decomposition process and is dependent on the particle size, where the Ea decreases from 116 kJ mol−1, in the case of bulk material, to 58 kJ mol−1, in the case of nanoparticles from 2 to 10 nm.
In this work, the influence of nanoconfinement on the H2 release processes from NaAlH4 is presented in depth. Bulk NaAlH4, NaAlH4 deposited as bulk material on a carbon support material, and completely nanoconfined NaAlH4 materials are synthesized, and H2 release curves at constant T ramp rates, β = dT/dt, are measured and analyzed. By comparing Ea values of H2 release from different phases of NaAlH4, i.e., bulk, deposited as bulk on a supporting material, and nanoconfined, a more complete picture of the effect of nanoconfinement on the H2 release processes from NaAlH4 is obtained and presented.

2. Materials and Methods

2.1. Synthesis of NaAlH4/Carbon Composites

NaAlH4 is deposited on a high surface area, specific surface area = 1840 m2 g−1, microporous, most pores with widths under 2 nm, and high porosity, pore volume = 0.82 cm3 g−1, dry carbon material RP-20 (Kuraray, Japan) through the solution impregnation method. Tetrahydrofuran, THF (anhydrous, ≥99.9%, Sigma-Aldrich, Germany), is used as the solvent, and a 0.05 gNaAlH4 mLTHF−1 solution is used for all samples. Bulk NaAlH4 (90%, Sigma-Aldrich, Germany) was dissolved and filtrated through a glass microfiber filter (GF/B, Whatman, UK) to remove any insoluble impurities. Samples with 5, 60, and 100 mass% of deposited NaAlH4 are synthesized to yield the nanoconfined, bulk-deposited, and bulk NaAlH4, respectively, and are denominated as such. In the case of the bulk NaAlH4 sample, recrystallization is performed at identical conditions to get rid of any impurities, but without the carbon support material. All syntheses, sample storage, and sample preparation for further analysis were performed in an Ar-filled (5.0, Linde) glovebox (MBraun LABmaster sp, Germany). A full description of the synthesis process and the characterization of the synthesized materials is provided in [13].

2.2. Temperature Controlled Decomposition

The H2 release curves at constant β were measured with the Autochem 2950HP (Micromeritics, USA) chemisorption analyzer. The amount of released H2 was determined with a thermal conductivity detector, where a 50 mL min−1 N2 (6.0, Linde) gas flow was used as the carrier gas. β from 0.5 °C min−1 to 10 °C min−1 were applied. A full description of the sample preparation and measurement routine is brought in [13]. Data reduction, data analysis, and H2 release peak fitting were performed with the OriginPro Version 2016 (OriginLab Corporation, Northampton, MA, USA) software.

3. Results

3.1. Hydrogen Release from NaAlH4/Carbon Composites

The bulk, bulk-deposited, and nanoconfined NaAlH4 exhibit distinct H2 release curve shapes (Figure 1) in the case of all β used. In the case of the bulk NaAlH4, a low amount of H2 is released at ~170 °C, which is just under the T for melting of bulk NaAlH4. The release of H2 at ~170 °C is most likely the release of H2 from the surface layer and defect sites, for which the melting T is lower, and thus, H2 is released before the melting of the bulk phase. Starting from T = 200 °C, H2 is released from the bulk of the material through the decomposition steps Equations (1) and (2), and at T > 280 °C H2 is released from the last decomposition step, Equation (3). At T > 375 °C, all of the H2 has been released as the material has been completely decomposed.
In the case of bulk-deposited NaAlH4, a very low amount of H2 is already released, starting at T < 90 °C. The T at which the initial H2 release starts depends strongly on the β applied and is released at an almost constant rate. This H2 release step is caused by the small amount of NaAlH4 deposited in an amorphous state (highly polycrystalline structure) and/or as separate nanosized particles during the recrystallization process. Starting from T = 165 °C, a high amount of H2 is released over a narrow T range and is followed by a high amount of released H2 over a wide T range up to ~200 °C, after which a low amount of H2 is released at a constant rate up to T ~ 375 °C. These three steps correspond to NaAlH4 decomposition: Equation (1) from the surface, Equations (1) and (2) from bulk, and Equation (3), respectively. For a better deconvolution of the H2 release from different decomposition steps and processes, the whole H2 release curve has been fitted with distribution functions (presented in the next section).
In the case of nanoconfined NaAlH4 H2 release starts already at ambient conditions, T < 23 °C in case of all applied β. H2 is mainly released at T < 200 °C, with the shape of the H2 release curve strongly dependent on the applied β. In addition, a very low amount of H2 is released at T > 400 °C, which is most likely H2 released from the decomposition of NaH. Thus, deconvolution of the H2 release peaks through fitting is necessary to understand the decomposition processes of nanoconfined NaAlH4.
The amount of H2 released, compared to the theoretical amount of H2 in the NaAlH4 of the sample, yields the H2 content efficiency of the investigated materials. The average H2 content efficiencies are ~100%, ~66%, and ~40% for bulk, bulk-deposited, and nanoconfined, respectively. Thus, the amount of recovered H2 from NaAlH4 decreases with the deposition process, especially with the nanoconfinement of NaAlH4. The decrease in H2 content efficiency is likely caused by the decomposition of NaAlH4 during synthesis, where the step-by-step addition of NaAlH4 in THF solution to the carbon support promotes the decomposition of NaAlH4, especially at the initial addition steps where NaAlH4 is deposited as nanosized particles. In addition, as the H2 release from nanoconfined NaAlH4 begins at ambient conditions, it is very likely that some of the nanoconfined NaAlH4 decomposes during storage at ambient conditions, i.e., at 21 °C and in Ar gas environment.

3.2. Modelling of Decomposition Processes

To better understand the decomposition processes of NaAlH4 in different phases, e.g., bulk, bulk-deposited, and nanoconfined, the H2 release curves were fitted with a combination of gaussian and bigaussian peak functions (Figure 2). The bigaussian, i.e., an asymmetric gaussian profile distribution peak function which has two independent peak widths, w1 and w2, at x < xc (Equation (4)) and xxc (Equation (5)), respectively, where xc is the position of the peak center
y = y0 + H × exp(−0.5 × (xxc/w1)2) if (x < xc)
y = y0 + H × exp(−0.5 × (xxc/w2)2) if (xxc)
where y0 is the baseline and H is the peak height.
The bigaussian profile of the H2 release is caused by the kinetic limitation of some decomposition steps, where the increase in the H2 release amount with the initial rise in T is quick. After reaching maximum H2 release from the process at temperature Tmax, the H2 release process continues over a wide T range as H2 release from the decomposition process is kinetically hindered.
From the fitting of the H2 release curves, Tmax is obtained for each decomposition process. The range of obtained Tmax, determined from H2 release curves measured at different β, is presented in Table 1 with the corresponding designations for each process. It is possible to calculate the Ea of a first-order reaction using the Kissinger method [8,15]
ln(β/Tmax2) = ln(Z × R/Ea)− Ea/(R × Tmax),
where R is the ideal gas constant, and Z is the Arrhenius pre-exponential factor. By plotting ln(β/Tmax2) vs. 1/Tmax, the slope will yield −Ea/R, and thus, Ea can be calculated. The Ea values of the deconvoluted H2 release processes have been calculated from the slope of the Kissinger plot (Figure 3, Table 1).
In the case of bulk NaAlH4, all the activation energies are relatively high, >100 kJ mol−1, as the H2 release processes are kinetically hindered by the diffusion of H2 out of the bulk phase and by the thermal conductivity of the bulk phase, where the H2 release processes from NaAlH4 are endothermic [16], and thus preventively cooling the surrounding material upon decomposition. Even though the temperature of initial H2 release from the first decomposition step depends on the applied constant β, the low amount of H2 released hinders proper fitting. Thus, the obtained Tmax does not remarkably depend on applied constant β. Therefore, the Ea could not be calculated and is most likely caused by the release of H2 from the surface of NaAlH4 particles over a wide range of T values under the melting T of bulk NaAlH4. The highest Ea from the bulk material is for the second peak, which corresponds to Equation (1) reaction step from the bulk of the material. This H2 release step is over a wide T range, as the H2 diffusion length can vary widely in large NaAlH4 particles, and thus H2 release from inside the NaAlH4 particles occurs at remarkably higher T values. Simultaneously with H2 release from the Equation (1) reaction step, H2 release from the Equation (2) reaction step starts. The H2 release from the Equation (2) reaction step is discernably taking place in two distinct stages, as H2 from the surface is released first, and then the H2 from inside the particles is released. H2 release from the last irreversible reaction step, Equation (3), is clearly separatable and has a relatively high Ea, 150 kJ mol−1, as the decomposition products from preliminary H2 release steps limit the kinetics of H2 release from this step.
A similar case can be seen for bulk-deposited, but the Tmax of all H2 release steps are lowered, and the Ea values of the last two steps are also lower. This is most likely caused by the partial confinement of the decomposition products inside the porous carbon structure, enhancing the kinetics of the H2 release step corresponding to Equation (2). This restructuring and the disappearance of clear crystalline phases upon the dehydrogenation of bulk-deposited NaAlH4 has been shown before [10,13].
In the case of nanoconfined NaAlH4, three separate H2 release steps are identified under one continuous H2 release peak, starting from ambient T and continuing up to 200 °C, where the H2 release step starting with Tmax from 51 to 57 °C is discernible only at a high applied constant β. All these H2 release steps have relatively low Ea values, compared to bulk and bulk-deposited, are overlapping, and do not clearly correspond to a discrete H2 release reaction. Baldé et al. have shown that the Tmax of H2 release decreases remarkably with the decrease in the nanoparticle size [8], and Gao et al. have shown that nanosized NaAlH4 decomposes directly into NaH [7]. Thus, the indiscernibility of the H2 release steps of nanoconfined NaAlH4 is very likely caused by a distribution of differently sized NaAlH4 nanoparticles, which decompose at different T values directly into NaH, going through the Equations (1) and (2) reactions in one step. In addition, the decomposition of the nanoconfined NaAlH4 at ambient conditions, based on H2 content efficiency, makes the assignment of the H2 release peaks a nontrivial task. Regardless, the calculated Ea values, from 30 to 45 kJ mol−1, of H2 release steps discernible at all applied constant β from nanoconfined NaAlH4 are notably lower than the value of 58 kJ mol−1 achieved for 2–10 nm NaAlH4 particles [8]. The H2 release step detected at the highest T, most likely from the Equation (3) reaction step, releases a low amount of H2, and the Tmax of this step is 100 °C higher than that of bulk and bulk-deposited materials and has an increased Ea. Thus, it is possible that nanoconfined NaAlH4 does not decompose completely to Na and Al under the investigated conditions, as, for some reason, nanosized NaH is more stable than bulk NaH. This would improve the cyclability and usability of nanoconfined-NaAlH4-based H2 storage systems, as the decomposition of NaH is irreversible and, thus, unwelcome. The increased stability of the nanoconfined NaH phase must be investigated further.
The addition of fitting the H2 release curves to deconvolute different decomposition steps before applying the Kissinger method has clearly shown that various activation barriers are present even in the case of nanoconfinement. This increases our understanding of the fundamental processes involved in releasing H2 from NaAlH4, especially when nanoconfined, which may prove vital when improving and optimizing such materials for H2 storage applications.
The achieved activation energies for the reversible steps are considerably lower than most activation energies of mixed hydride [17,18,19], doped hydride [20,21,22,23,24,25], and nanoconfined hydride [9,26,27] systems described in the literature. Those include systems that utilize NaAlH4 [9,19,24,25,27], but also other metal and complex hydrides [17,18,20,21,22,23,26]. Most activation energies for bulk hydrides are well over 100 kJ mol−1. The achieved improved energies in the referred papers are mostly larger, by at least a factor of two, than the ones presented in this paper, with relatively few coming into the range of 30–45 kJ mol−1 [22,27]. This is remarkable, taking into account the simplicity of the system and its synthesis. Doping involves adding transition metals or compounds, which often adds a synthesis step of varying complexity. Mixtures may have side reactions, which are complex to analyze and may result in unwanted products or irreversible alloys [17,18,19]. Using a scaffold material for confinement improves cycling stability [13], which is essential for practical applications. In this work, the results have been achieved using a simple commercial microporous carbon, which offers lower Ea values than laboratory-made materials [26,27] and potentially higher loadings than, e.g., metal-organic frameworks [9]. Doping, ball-milling, and melt-infiltration, which even occurs naturally upon cycling above the melting temperature of NaAlH4, is expected to improve the material’s hydrogen storage properties even further.

4. Conclusions

The deconvolution of H2 release processes through curve-fitting procedures to determine Tmax when applying the Kissinger method significantly improved the amount and quality of the information received about the H2 release processes in bulk, bulk-deposited, and nanoconfined NaAlH4. The coating of porous carbon particles with a thick layer of NaAlH4, bulk-deposited, is shown to decrease the Tmax of all H2 release steps, whereas only the Ea values of the last steps are lowered. This indicates that whilst a supporting material reduces the T at which H2 is released, the H2 release processes are still limited by the H2 diffusion out of the bulk phase. In the case of nanoconfined NaAlH4, H2 release starts at ambient T, and the Ea of H2 release processes is remarkably lowered in comparison to bulk and bulk-deposited NaAlH4. None of the H2 release steps are limited by H2 diffusion in the case of nanoconfined NaAlH4. Even though the nanoconfined NaAlH4 did not have a discrete particle size, based on the discernability of multiple H2 release steps, very low Ea values from 30 to 45 kJ mol−1 are determined. These Ea values are a significant improvement on the vast majority of previously used techniques for nanoconfinement. The H2 release at ambient T and the very low Ea of the nanoconfined NaAlH4, in addition to limiting the final, irreversible decomposition step, highlights the possible utility of nanoconfinement for the improvement of NaAlH4-based H2 storage materials, where the critical problem is synthesizing nanoconfined composites with a high total H2 yield per mass and/or volume.

Author Contributions

Conceptualization, R.P.; methodology, R.P.; formal analysis, R.P., K.T.; investigation, R.P.; resources, R.P.; writing—original draft preparation, R.P.; writing—review and editing, K.T.; funding acquisition, R.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the EU through the European Regional Development Fund (Centers of Excellence, TK141 2014-2020.4.01.15-0011) and by the Estonian Research Council Grant, grant number PUTJD957.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Von Helmolt, R.; Eberle, U. Fuel cell vehicles: Status 2007. J. Power Sources 2007, 165, 833–843. [Google Scholar] [CrossRef]
  2. Peighambardoust, S.; Rowshanzamir, S.; Amjadi, M. Review of the proton exchange membranes for fuel cell applications. Int. J. Hydrogen Energy 2010, 35, 9349–9384. [Google Scholar] [CrossRef]
  3. Hua, T.Q.; Ahluwalia, R.; Peng, J.-K.; Kromer, M.; Lasher, S.; McKenney, K.; Law, K.; Sinha, J. Technical assessment of compressed hydrogen storage tank systems for automotive applications. Int. J. Hydrogen Energy 2011, 36, 3037–3049. [Google Scholar] [CrossRef] [Green Version]
  4. Makridis, S.S. Hydrogen storage and compression. In Methane and Hydrogen for Energy Storage; Carriveau, R., Ting, D.S.-K., Eds.; Institution of Engineering and Technology: London, UK, 2016; pp. 1–28. [Google Scholar]
  5. Schüth, F.; Bogdanović, B.; Felderhoff, M. Light metal hydrides and complex hydrides for hydrogen storage. Chem. Commun. 2004, 20, 2249–2258. [Google Scholar] [CrossRef] [PubMed]
  6. Ke, X.; Tanaka, I. Decomposition reactions forNaAlH4, Na3AlH6, and NaH: First-principles study. Phys. Rev. B 2005, 71, 024117. [Google Scholar] [CrossRef]
  7. Gao, J.; Adelhelm, P.; Verkuijlen, M.H.W.; Rongeat, C.; Herrich, M.; Van Bentum, P.J.M.; Gutfleisch, O.; Kentgens, A.P.M.; De Jong, K.P.; De Jongh, P.E. Confinement of NaAlH4 in Nanoporous Carbon: Impact on H2 Release, Reversibility, and Thermodynamics. J. Phys. Chem. C 2010, 114, 4675–4682. [Google Scholar] [CrossRef]
  8. Baldeé, C.P.; Hereijgers, B.P.C.; Bitter, J.H.; De Jong, K.P. Sodium Alanate Nanoparticles—Linking Size to Hydrogen Storage Properties. J. Am. Chem. Soc. 2008, 130, 6761–6765. [Google Scholar] [CrossRef] [Green Version]
  9. Bhakta, R.K.; Maharrey, S.; Stavila, V.; Highley, A.; Alam, T.; Majzoub, E.; Allendorf, M. Thermodynamics and kinetics of NaAlH4 nanocluster decomposition. Phys. Chem. Chem. Phys. 2012, 14, 8160–8169. [Google Scholar] [CrossRef]
  10. Zheng, S.; Fang, F.; Zhou, G.; Chen, G.; Ouyang, L.; Zhu, M.; Sun, D. Hydrogen Storage Properties of Space-Confined NaAlH4Nanoparticles in Ordered Mesoporous Silica. Chem. Mater. 2008, 20, 3954–3958. [Google Scholar] [CrossRef]
  11. Bendyna, J.K.; Dyjak, S.; Notten, P.H. The influence of ball-milling time on the dehydrogenation properties of the NaAlH4–MgH2 composite. Int. J. Hydrogen Energy 2015, 40, 4200–4206. [Google Scholar] [CrossRef] [Green Version]
  12. Andrei, C.; Walmsley, J.C.; Brinks, H.; Holmestad, R.; Srinivasan, S.; Jensen, C.; Hauback, B. Electron-microscopy studies of NaAlH4 with TiF3 additive: Hydrogen-cycling effects. Appl. Phys. A 2005, 80, 709–715. [Google Scholar] [CrossRef]
  13. Palm, R.; Kurig, H.; Aruväli, J.; Lust, E. NaAlH4 /microporous carbon composite materials for reversible hydrogen storage. Microporous Mesoporous Mater. 2018, 264, 8–12. [Google Scholar] [CrossRef]
  14. Baldé, C.P.; Hereijgers, B.P.C.; Bitter, J.H.; De Jong, K.P. Facilitated Hydrogen Storage in NaAlH4 Supported on Carbon Nanofibers. Angew. Chem. 2006, 118, 3581–3583. [Google Scholar] [CrossRef] [Green Version]
  15. Blaine, R.L.; Kissinger, H.E. Homer Kissinger and the Kissinger equation. Thermochim. Acta 2012, 540, 1–6. [Google Scholar] [CrossRef]
  16. Li, L.; Wang, Y.; Qiu, F.; Xu, Y.; An, C.; Jiao, L.; Yuan, H. Reversible hydrogen storage properties of NaAlH4 enhanced with TiN catalyst. J. Alloys Compd. 2013, 566, 137–141. [Google Scholar] [CrossRef]
  17. Shaw, L.L.; Ren, R.; Markmaitree, T.; Osborn, W. Effects of mechanical activation on dehydrogenation of the lithium amide and lithium hydride system. J. Alloys Compd. 2008, 448, 263–271. [Google Scholar] [CrossRef]
  18. Xiong, Z.; Hu, J.; Wu, G.; Chen, P.; Luo, W.; Gross, K.; Wang, J. Thermodynamic and kinetic investigations of the hydrogen storage in the Li–Mg–N–H system. J. Alloys Compd. 2005, 398, 235–239. [Google Scholar] [CrossRef]
  19. Ismail, M. Study on the hydrogen storage properties and reaction mechanism of NaAlH4–MgH2–LiBH4 ternary-hydride system. Int. J. Hydrogen Energy 2014, 39, 8340–8346. [Google Scholar] [CrossRef]
  20. Gu, J.; Gao, M.; Pan, H.; Liu, Y.-F.; Li, B.; Yang, Y.; Liang, C.; Fu, H.; Guo, Z.X. Improved hydrogen storage performance of Ca(BH4)2: A synergetic effect of porous morphology and in situ formed TiO2. Energy Environ. Sci. 2012, 6, 847–858. [Google Scholar] [CrossRef]
  21. Valentoni, A.; Mulas, G.; Enzo, S.; Garroni, S. Remarkable hydrogen storage properties of MgH2 doped with VNbO5. Phys. Chem. Chem. Phys. 2018, 20, 4100–4108. [Google Scholar] [CrossRef]
  22. Cui, J.; Wang, H.; Liu, J.; Ouyang, L.; Zhang, Q.; Sun, D.; Yao, X.; Zhu, M. Remarkable enhancement in dehydrogenation of MgH2 by a nano-coating of multi-valence Ti-based catalysts. J. Mater. Chem. A 2013, 1, 5603–5611. [Google Scholar] [CrossRef]
  23. Jangir, M.; Jain, A.; Yamaguchi, S.; Ichikawa, T.; Lal, C.; Jain, I. Catalytic effect of TiF4 in improving hydrogen storage properties of MgH2. Int. J. Hydrogen Energy 2016, 41, 14178–14183. [Google Scholar] [CrossRef]
  24. Ali, N.; Ismail, M. Modification of NaAlH4 properties using catalysts for solid-state hydrogen storage: A review. Int. J. Hydrogen Energy 2021, 46, 766–782. [Google Scholar] [CrossRef]
  25. Sandrock, G.; Gross, K.; Thomas, G. Effect of Ti-catalyst content on the reversible hydrogen storage properties of the sodium alanates. J. Alloys Compd. 2002, 339, 299–308. [Google Scholar] [CrossRef]
  26. Gross, A.F.; Vajo, J.J.; Van Atta, S.L.; Olson, G.L. Enhanced Hydrogen Storage Kinetics of LiBH4 in Nanoporous Carbon Scaffolds. J. Phys. Chem. C 2008, 112, 5651–5657. [Google Scholar] [CrossRef]
  27. Li, Y.; Zhou, G.; Fang, F.; Yu, X.Y.; A Zhang, Q.; Ouyang, L.Z.; Zhu, M.; Sun, D. De-/re-hydrogenation features of NaAlH4 confined exclusively in nanopores. Acta Mater. 2011, 59, 1829–1838. [Google Scholar] [CrossRef]
Figure 1. Hydrogen release curves of different forms of NaAlH4 (noted in the figure) measured at a constant temperature ramp rate, β, of 2 °C min−1.
Figure 1. Hydrogen release curves of different forms of NaAlH4 (noted in the figure) measured at a constant temperature ramp rate, β, of 2 °C min−1.
Reactions 02 00001 g001
Figure 2. Hydrogen release curve from bulk-deposited measured during temperature ramp rate, β, of 1 °C min−1 with the corresponding hydrogen release peak fits.
Figure 2. Hydrogen release curve from bulk-deposited measured during temperature ramp rate, β, of 1 °C min−1 with the corresponding hydrogen release peak fits.
Reactions 02 00001 g002
Figure 3. Kissinger equation plots of (a) bulk, (b) bulk-deposited, and (c) nanoconfined NaAlH4 and the corresponding fits to the Tmax of H2 release processes obtained from fitting the H2 release curves with gaussian and bigaussian peak functions.
Figure 3. Kissinger equation plots of (a) bulk, (b) bulk-deposited, and (c) nanoconfined NaAlH4 and the corresponding fits to the Tmax of H2 release processes obtained from fitting the H2 release curves with gaussian and bigaussian peak functions.
Reactions 02 00001 g003
Table 1. Hydrogen release peaks and Kissinger method fitting results.
Table 1. Hydrogen release peaks and Kissinger method fitting results.
MaterialMost likely H2 Release stepTemperature Range of Tmax/°CEa/kJ mol−1Additional Comments
BulkEquation (1)167–172Low amount, most likely only surface layer, no clear dependence of Tmax on β
BulkEquation (1)230–243270 ± 10Over a wide T range, from the whole phase
BulkEquation (2)240–264150 ± 60Surface layer, low amount
BulkEquation (2)245–279120 ± 20From the whole phase
BulkEquation (3)308–337150 ± 30From the whole phase
Bulk-depositedEquation (1)129–165Over a wide T range, no clear dependence of Tmax on β. Most likely decomposition of nano- and surface layer.
Bulk-depositedEquation (1)165–174400 ± 50H2 release over a very narrow T range
Bulk-depositedEquation (1)174–190170 ± 20From the whole phase
Bulk-depositedEquation (2)180–22780 ± 10From the whole phase
Bulk-depositedEquation (3)263–32590 ± 20From the whole phase
NanoconfinedEquation (1)51–5770 ± 40Only present at high β
NanoconfinedEquations (1) and (2)41–9430 ± 5Overlap strongly, both over a wide T range
NanoconfinedEquations (1) and (2)102–18045 ± 5
NanoconfinedEquation (3)407–452210 ± 30Low amount
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tuul, K.; Palm, R. Influence of Nanoconfinement on the Hydrogen Release Processes from Sodium Alanate. Reactions 2021, 2, 1-9. https://doi.org/10.3390/reactions2010001

AMA Style

Tuul K, Palm R. Influence of Nanoconfinement on the Hydrogen Release Processes from Sodium Alanate. Reactions. 2021; 2(1):1-9. https://doi.org/10.3390/reactions2010001

Chicago/Turabian Style

Tuul, Kenneth, and Rasmus Palm. 2021. "Influence of Nanoconfinement on the Hydrogen Release Processes from Sodium Alanate" Reactions 2, no. 1: 1-9. https://doi.org/10.3390/reactions2010001

Article Metrics

Back to TopTop