Next Article in Journal
A Numerical Investigation on the Collision Behavior of Unequal-Sized Micro-Nano Droplets
Previous Article in Journal
Sustainable Nanotechnologies for Curative and Preventive Wood Deacidification Treatments: An Eco-Friendly and Innovative Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Realizing Hydrogen De/Absorption Under Low Temperature for MgH2 by Doping Mn-Based Catalysts

College of Energy and Power, Jiangsu University of Science and Technology, Zhenjiang 212003, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(9), 1745; https://doi.org/10.3390/nano10091745
Submission received: 6 August 2020 / Revised: 31 August 2020 / Accepted: 1 September 2020 / Published: 3 September 2020
(This article belongs to the Section Energy and Catalysis)

Abstract

:
Magnesium hydride (MgH2) has been considered as a potential material for storing hydrogen, but its practical application is still hindered by the kinetic and thermodynamic obstacles. Herein, Mn-based catalysts (MnCl2 and Mn) are adopted and doped into MgH2 to improve its hydrogen storage performance. The onset dehydrogenation temperatures of MnCl2 and submicron-Mn-doped MgH2 are reduced to 225 °C and 183 °C, while the un-doped MgH2 starts to release hydrogen at 315 °C. Further study reveals that 10 wt% of Mn is the better doping amount and the MgH2 + 10 wt% submicron-Mn composite can quickly release 6.6 wt% hydrogen in 8 min at 300 °C. For hydrogenation, the completely dehydrogenated composite starts to absorb hydrogen even at room temperature and almost 3.0 wt% H2 can be rehydrogenated in 30 min under 3 MPa hydrogen at 100 °C. Additionally, the activation energy of hydrogenation reaction for the modified MgH2 composite significantly decreases to 17.3 ± 0.4 kJ/mol, which is much lower than that of the primitive MgH2. Furthermore, the submicron-Mn-doped sample presents favorable cycling stability in 20 cycles, providing a good reference for designing and constructing efficient solid-state hydrogen storage systems for future application.

1. Introduction

Clean and sustainable energy is attracting tremendous attention worldwide because of the continuous shortage of fossil fuels and the worsening of environmental pollution. Hydrogen, which occupies higher energy density (142 MJ/kg) [1,2] than traditional fossil fuels and produces only clean and nontoxic water during combustion, is regarded as one of the most promising renewable energy resources [3,4]. Unfortunately, the utilization of hydrogen economy still faces many technical difficulties, especially for hydrogen storage [5,6,7]. Compared with the liquid and gaseous hydrogen, hydrogen stored in solid-state materials not only has the advantage of high hydrogen storage density, but also keeps safety during application [8,9,10]. Magnesium hydride (MgH2) with large mass hydrogen storage capacity (7.76 wt%), natural abundance, and excellent reversibility, ignites hope for meeting the demands of practical application of high-capacity hydrogen storage [11,12,13]. Nevertheless, its high thermodynamic stability and poor kinetic properties still lie in the way of practical application [14,15,16]. To conquer the above challenges, diverse technics like nanoconfinement [17,18,19,20,21,22], alloying [23,24,25,26,27], and catalyst doping [28,29,30,31,32,33,34,35] have been conducted over the past decades.
Transition metal halides were easy to be obtained and doped to MgH2 to improve its hydrogen storage properties [36,37,38,39,40,41]. Jangir et al. [42] observed that the initial desorption temperature of MgH2 was decreased by about 100 °C by doping TiF4 and the activation energy was lower by about 96 kJ/mol. Zhang et al. [43] successfully prepared a MgH2-NiCl2 composite and SEM tests exhibited that the addition of NiCl2 was conductive to decreasing the size of MgH2 grains and particles. Ismail et al. [44] doped FeCl3 into MgH2 to find that the desorption temperature of the MgH2-10 wt% FeCl3 composite was 90 °C lower than that of as-milled MgH2 and the activation energy for hydrogen desorption was also decreased from 166 kJ/mol to 130 kJ/mol. Mao et al. [45] revealed that the MgH2/NiCl2 sample could release 5.17 wt% H2 in 60 s at 300 °C and the dehydrogenation activation energy was decreased to 121.3 kJ/mol and 102.6 kJ/mol for MgH2/CoCl2 and MgH2/NiCl2 sample, respectively.
According to the above references, it can be concluded that doping transition metal halides into MgH2 could greatly enhance the de/hydrogenation properties. As far as we know, studies about Mn-based catalysts have rarely been researched, thus, it is urgent and interesting to explore the catalytic effect of MnCl2 for the reversible hydrogen storage performance of MgH2. However, there are still some shortcomings about the doping of transition metal halides. On one hand, the really doping amount of transition metal atoms are restricted because of the heavy halogen atoms. On the other hand, there would be a deadweight that Mg may react with halogen elements to form MgCl2 or MgF2 which could affect hydrogen capacity and absorption/ desorption rates [44].
In this work, the catalytic effect of MnCl2 was investigated and based on the microstructure evidence, submicron-Mn was successfully synthesized via a simple wet chemical method and doped directly to MgH2 to further enhance the hydrogen storage properties of MgH2. Moreover, its catalytic mechanism was explored and discussed in detail.

2. Materials and Methods

2.1. Sample Preparation

Powders of manganese chloride (MnCl2) was purchased from Sinopharm Chemical Reagent and Mn powders was commercially purchased from Aladdin Industrial Corporation. Submicron-Mn particles were prepared by a wet-chemical ball milling method. At first, 4 g Mn powders (99.95%, Aladdin Industrial Corporation, Shanghai, China), 12 mL heptane (98.5%, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), 0.6 mL oleic acid (90%, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), 0.2 mL oleylamine (98%, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), and 240 g balls were mingled in a home-made stainless steel jar under 0.1 MPa of Ar. The mixture was milled at a speed of 400 rpm for 60 h in the planetary ball mill (QM-3SP4, Nanjing, China). The treated slurry mixed with another 15 mL heptane, 1 mL oleic acid, and 1 mL oleic acid was then placed in a centrifuge tub. In addition, ethanol was used to centrifuge and wash the mixed solution eight times to remove larger particles and residual organic solvent. Finally, Mn submicron particles (submicron-Mn) can be acquired after vacuum-drying at room temperature for 12 h.
The MgH2 used was synthesized in our laboratory. First, Mg powder (99%, 100–200 mesh, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) was placed at 380 °C under the hydrogen pressure of 6.5 MPa to absorb hydrogen for 2 h. The second step was to ball-mill the processed samples at 450 rpm for 5 h. After repeating the above hydrogenation heat treatment, MgH2 can be finally acquired.
The MgH2 + x wt% submicron-Mn (x = 5, 10 and 15) and MgH2- MnCl2 composites were prepared by mechanical ball milling at 450 rpm for 2 h under 0.1 MPa of Ar (the ball to material ratio is 40:1). In order to avoid oxidation and contamination, all samples were handled and transferred in an Ar-filled glove box (Mikrouna, Shanghai, China) where the oxygen /water concentration was kept less than 0.1 ppm.

2.2. Sample Characterization

X-ray diffraction (XRD) analyses of all samples were performed on an X’Pert Pro X-ray diffractometer (PAN alytical, Royal Dutch Philips Electronics Ltd, Amsterdam, Netherlands) with Cu K alpha radiation at 40 KV, 40 mA to detect the phase compositions. To avoid air and water contamination, a special container was adopted for transferring and scanning samples. A scanning electron microscopy (SEM, Hitachi SU-70, Tokyo, Japan) with an energy dispersive spectroscopy (EDS) was performed to further characterize morphologies and element distribution of the samples. The hydrogen absorption and desorption properties were tested in a Sieverts-type apparatus. During testing non-isothermal hydrogen desorption properties, about 75 mg sample was heated to 450 °C at a heating rate of 2 °C min−1 in a sealed stainless steel reactor. For hydrogenation, the samples were gradually heated from room temperature to 400 °C at an average rate of 1 °C min−1 under 3 MPa H2. For isothermal measurements, the samples were first heated up to the desired temperature and then keeping the temperature constant in the whole test. In order to get the exact values of hydrogenation capacity, the second dehydrogenation measurements were also conducted to verify the accuracy of the values. Moreover, controlling the hydrogen pressure for de/hydrogenation tests well is also important, the isothermal absorption tests were performed at various temperatures under 3 MPa while the isothermal desorption performance was tested at different temperatures under hydrogen pressure below 0.001 MPa.

3. Results and Discussion

To investigate the catalytic effect of MnCl2 on the hydrogen storage properties of MgH2, 5 wt% MnCl2 was ball milled with MgH2 to prepare the MgH2 + 5wt% MnCl2 composite, and temperature-programmed desorption (TPD) tests were conducted from room temperature to 450 °C, the results are shown in Figure 1a. The un-doped MgH2 began to release hydrogen from 315 °C and about 7.45 wt% hydrogen could be desorbed after the non-isothermal test. It is clear that the dehydrogenation temperature shifts to lower temperature after doping MnCl2 powders, which started to release hydrogen at 230 °C and about 6.8 wt% hydrogen could be attained when heating up to 350 °C. XRD measurements were also used to shed light on the microstructure evolution in the desorption process. Figure 1b shows that MgH2 still dominates the diffraction peaks for MgH2-MnCl2 sample and no apparent new phases appeared during the ball-milling process, indicating that it was only a physical mixture after the ball milling process. After dehydrogenation, it is interesting that the signal of Mn phase emerged at 43° besides Mg phase. It is much likely that MnCl2 reacted with MgH2 during the dehydrogenation process, just as reported in other transition metal halides-modified MgH2 systems [42,44]. Therefore, Mn may be the key to enhance the dehydrogenation performance of MgH2. In order to confirm this conjecture, submicron-Mn particles were further synthesized and a series of tests were performed.
The as-prepared Mn particles were analyzed by XRD and SEM measurements, as shown in Figure 2. It can be seen from Figure 2a that the diffraction peaks in the XRD patterns of as-prepared Mn correspond well to that of the standard peaks of Mn phase (PDF#0637). Moreover, the diffraction peaks for as-prepared Mn become wider and weaker, suggesting a decrease in crystallite size under the effect of ball-milling. The SEM images from Figure 2b–d shows that the purchased Mn powders present the appearance of micrometer range particles while the particle size of as-prepared submicron-Mn ranged from 500 nm to 800 nm, which could be defined as submicron particles. Combing XRD results with SEM pictures, it can be concluded that submicron-Mn particles were successfully synthesized and a great catalytic effect was expected on the hydrogen storage properties of MgH2.
To witness the modification impact of the as-synthesized Mn submicron particles on MgH2, isothermal and non-isothermal dehydrogenation tests were operated. As a comparison, the purchased Mn was also doped into MgH2. The non-isothermal dehydrogenation curves in Figure 3a depicted that the onset temperatures of MgH2 + 5wt% Mn composite was 225 °C, 5 °C, and 90 °C lower than that of MgH2 + 5wt% MnCl2 composite and additive-free MgH2, respectively. Just as expected, after submicron-Mn was doped to MgH2, MgH2 + 5wt% submicron-Mn composite began to release hydrogen at 183 °C, superior to purchased Mn and MnCl2. In order to figure out the best doping amount, different amounts of submicron-Mn were ball-milled with MgH2 and further isothermal and non-isothermal measurements were conducted on the MgH2 + submicron-Mn composites. It could be clearly seen from Figure 3b that the volumetric release curves of submicron-Mn modified samples shifted toward lower temperatures with the increasing doping amount. The MgH2 + 5 wt% submicron-Mn composite possessed onset dehydrogenation temperatures of 183 °C, about 132 °C lower than that of prepared MgH2. As for the MgH2 + 10 wt% submicron-Mn and MgH2 + 15 wt% submicron-Mn composites, the onset desorption temperatures further decreased to 175 °C and 165 °C, respectively. When the temperature rose to 350 °C, about 6.8 wt%, 6.5 wt%, and 6.1 wt% H2 could be obtained for the MgH2 + 5 wt% submicron-Mn, MgH2 + 10 wt% submicron-Mn, and MgH2 + 15 wt% submicron-Mn samples, respectively. Further isothermal dehydrogenation measurements of the above three samples are performed at 275 °C. Figure 3c shows that only 4.7 wt% of H2 was desorbed in the first 10 min for the MgH2 + 5 wt% submicron-Mn composite. For the MgH2 + 10 wt% submicron-Mn and the MgH2 + 15 wt% submicron-Mn samples, the values increased to 6.1 wt% and 6.0 wt%. On the contrary, the pristine MgH2 sample could hardly release hydrogen under the same condition.
According to the results of TPD curves, it could be concluded that the addition of submicron-Mn could remarkably improve the hydrogen desorption kinetics of MgH2. Moreover, the initial dehydrogenation temperature did not decrease obviously after increasing the doping amount of catalyst. From a comprehensive perspective of the dehydrogenation temperature and capacity, the MgH2 + 10 wt% submicron-Mn was chosen for further study. Figure 3d presented the isothermal dehydrogenation curves of MgH2 + 10 wt% submicron-Mn composite at 250 °C, 275 °C, and 300 °C, respectively. The MgH2 + 10 wt% submicron-Mn composite could quickly release 6.6 wt% hydrogen in 8 min at 300 °C (almost 96.5% of theoretical hydrogen storage capacity). At 275 °C, this sample could desorb the same amount hydrogen with 20 min. Furthermore, about 6.0 wt% H2 could be acquired at 250 °C.
Besides the significantly improved dehydrogenation properties, hydrogen absorption kinetics of MgH2+ submicron-Mn composites were also investigated. As shown in Figure 4, the isothermal and non-isothermal hydrogenation tests were performed. From Figure 4a, it can be seen that the dehydrogenated MgH2 + 10 wt% submicron-Mn sample could absorb H2 even at room temperature and about 5.5 wt% hydrogen could be re-absorbed when heating up to 250 °C. However, the dehydrogenated MgH2 sample sluggishly took up hydrogen from 186 °C. The hydrogen absorption curves of MgH2 + 10 wt% submicron-Mn at relatively low temperature were also performed, shown in Figure 4b. Even at a low temperature of 50 °C, the dehydrogenated MgH2 + 10 wt% submicron-Mn sample still absorbed 1.8 wt% hydrogen within 40 min. When the temperature went up to 75 °C, the hydrogen uptake of the submicron-Mn-doped sample amounted to 2.3 wt% under the same condition. After hydrogenation at 100 °C for 40 min, the fully dehydrogenated composite could absorb 3.2 wt% hydrogen. As a comparison, non-isothermal hydrogenation measurement of MgH2 sample were also conducted (Figure 4c). For the dehydrogenated MgH2 sample, only 2.6 wt% H2 was absorbed even at 210 °C within 30 min.
In addition, the Ea values of the hydrogenation reaction were calculated to further explore the improved kinetics of hydrogenation for MgH2 + 10 wt% submicron-Mn sample. Some kinetic models such as Johnson-Mehl-Avrami-Kolmogorov (JMAK) model for the gas-solid reaction were adopted to simulate the evolution of kinetics [46,47]. Figure 4d depicts the JMAK model through fitting the absorption curves and the Ea values for the hydrogenation reactions were finally calculated according to Arrhenius equation [48].
ln[−ln(1 − α)] = nlnk + nlnt
k = Aexp(−Ea/RT)
where α is the fraction of Mg converted to MgH2 with time, k is an effective kinetic parameter, and n is the Avrami exponent. The numerical values of n and nlnk carried out by fitting the JMAK plots are shown in Figure S1.
In accordance with the curves shown in Figure 4d, the calculated Ea value of the hydrogenation process for the dehydrogenated MgH2 was calculated to be 72.5 ± 2.7 kJ/mol, while the value was reduced to 17.3 ± 0.4 kJ/mol for the dehydrogenated MgH2 + 10 wt% submicron-Mn sample. The greatly reduced activation energy also indicates that the energy barrier for hydrogenation was distinctly decreased after the addition of submicron-Mn, which well explains the evidently improved hydrogenation kinetics of the MgH2 + 10 wt% submicron-Mn sample.
Although doping submicron-Mn into MgH2 has shown great improvement in the reversible hydrogen storage properties, the catalytic mechanism of submicron-Mn in modifying MgH2 remained unknown. To further elucidate the hydrogen de/absorption mechanism, XRD tests of the MgH2 + 10 wt% submicron-Mn sample in ball-milled, dehydrogenated, and rehydrogenated state were performed. In the ball-milled state (Figure 5a), MgH2 phases still dominated the XRD pattern while the diffraction peaks of doped submicron-Mn were also found at around 43°. Interestingly, the XRD results demonstrated that the submicron-Mn was stable and persistently acted as an active substance during the process of de/hydrogenation. After dehydrogenation (Figure 5b) and 20th rehydrogenation (Figure 5c), the primary phase transformation during cycling is the transformation between Mg and MgH2. However, the diffraction peak of Mn which could be found at 43 ° was stable after 20 cycles and no other phases of Mn-related composites occurred, illuminating that Mn was the active catalyst to enhance the hydrogen storage properties.
In order to realize the practical application of hydrogen energy, preserving long-term kinetics is considered as one of the important indexes to evaluate the practicability for hydrogen storage materials. In this case, as shown in Figure 6, cycling tests of the MgH2 + 10 wt% submicron-Mn composite were further operated under the conditions of isothermal dehydrogenation and hydrogenation at 275 °C. As revealed in this pattern, the MgH2 + 10 wt% submicron-Mn sample could acquire a hydrogen capacity of 6.46 wt% in the first desorption. When exposed to hydrogen atmosphere of 3 MPa, the dehydrogenated sample quickly absorbed 5.94 wt% hydrogen. After 20 cycles, this composite could also release 5.72 wt% H2 (almost 89% of the original capacity). Compared with our previous study [29], the cycling property was better than that of MgH2 + nano-Fe samples, which had an evident decrease in the first 20 cycles. In general, the degenerating cycling properties caused by that MgH2 particles tend to grow and aggregate during the process of thermolysis [9,49]. Thus, although the addition of submicron-Mn can significantly enhance the de/hydrogenation performance of MgH2; other technics should still be explored to improve the cycling properties.

4. Conclusions

In summary, MnCl2 and Mn particles were doped as catalysts into MgH2 to improve its hydrogen storage properties and the submicron-Mn particles exhibited superior catalytic effect. The MgH2 + 10 wt% submicron-Mn composite started to release hydrogen at 175 °C and about 6.6 wt% hydrogen could be obtained within 8 min at 300 °C. For absorption performance, the dehydrogenated MgH2 + 10 wt% submicron-Mn sample began to absorb H2 at room temperature the completely dehydrogenated sample could assimilate 3.0 wt% H2 within 30 min under 100 °C while the dehydrogenated MgH2 needed a high temperature of 210 °C to absorb the same amount of H2. Besides, the Ea of hydrogen absorption of MgH2 was reduced to 17.3 ± 0.4 kJ/mol because of the addition of submicron-Mn. Moreover, the MgH2 + 10 wt% nano-Mn composite exhibited good cycling performance that 89% of initial hydrogen could still be maintained after 20 cycles.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/9/1745/s1, Figure S1: JMAK plots of MgH2 + 10 wt% submicron-Mn (a) composite and MgH2 (b).

Author Contributions

L.Z. and S.S. designed experiments; Z.S. carried out experiments; N.Y., T.B., and J.Z. analyzed experimental results; Z.S., Z.Y., and L.Z. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China [Grant No. 51801078], the National Science Foundation of Jiangsu Province [Grant No. BK20180986].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wan, C.; Zhou, L.; Sun, L.; Xu, L.; Cheng, D.-G.; Chen, F.; Zhan, X.; Yang, Y. Boosting visible-light-driven hydrogen evolution from formic acid over AgPd/2D g-C3N4 nanosheets Mott-Schottky photocatalyst. Chem. Eng. J. 2020, 396, 125229. [Google Scholar] [CrossRef]
  2. Jain, I.P. Hydrogen the fuel for 21st century. Int. J. Hydrogen Energy 2009, 34, 7368–7378. [Google Scholar] [CrossRef]
  3. Sharma, S.; Ghoshal, S.K. Hydrogen the future transportation fuel: From production to applications. Renew. Sustain. Energy Rev. 2015, 43, 1151–1158. [Google Scholar] [CrossRef]
  4. Reddy, S.N.; Nanda, S.; Dalai, A.K.; Kozinski, J.A. Supercritical water gasification of biomass for hydrogen production. Int. J. Hydrogen Energy 2014, 39, 6912–6926. [Google Scholar] [CrossRef]
  5. Zhang, J.; Zhu, Y.; Yao, L.; Xu, C.; Liu, Y.; Li, L. State of the art multi-strategy improvement of Mg-based hydrides for hydrogen storage. J. Alloys Compd. 2019, 782, 796–823. [Google Scholar] [CrossRef]
  6. Yu, X.; Tang, Z.; Sun, D.; Ouyang, L.; Zhu, M. Recent advances and remaining challenges of nanostructured materials for hydrogen storage applications. Prog. Mater. Sci. 2017, 88, 1–48. [Google Scholar] [CrossRef]
  7. Zhong, H.; Wang, H.; Liu, J.W.; Sun, D.L.; Fang, F.; Zhang, Q.A.; Ouyang, L.Z.; Zhu, M. Enhanced hydrolysis properties and energy efficiency of MgH2-base hydrides. J. Alloys Compd. 2016, 680, 419–426. [Google Scholar] [CrossRef]
  8. Ma, Z.; Liu, J.; Zhu, Y.; Zhao, Y.; Lin, H.; Zhang, Y.; Li, H.; Zhang, J.; Liu, Y.; Gao, W.; et al. Crystal-facet-dependent catalysis of anatase TiO2 on hydrogen storage of MgH2. J. Alloys Compd. 2020, 822, 153553. [Google Scholar] [CrossRef]
  9. Liu, M.; Xiao, X.; Zhao, S.; Saremi-Yarahmadi, S.; Chen, M.; Zheng, J.; Li, S.; Chen, L. ZIF-67 derived Co@CNTs nanoparticles: Remarkably improved hydrogen storage properties of MgH2 and synergetic catalysis mechanism. Int. J. Hydrogen Energy 2019, 44, 1059–1069. [Google Scholar] [CrossRef]
  10. Gao, S.; Wang, X.; Liu, H.; He, T.; Wang, Y.; Li, S.; Yan, M. Effects of nano-composites (FeB, FeB/CNTs) on hydrogen storage properties of MgH2. J. Power Source 2019, 438, 227006. [Google Scholar] [CrossRef]
  11. Zhou, C.; Bowman, R.C., Jr.; Fang, Z.Z.; Lu, J.; Xu, L.; Sun, P.; Liu, H.; Wu, H.; Liu, Y. Amorphous TiCu-Based Additives for Improving Hydrogen Storage Properties of Magnesium Hydride. ACS Appl. Mater. Interfaces 2019, 11, 38868–38879. [Google Scholar] [CrossRef] [PubMed]
  12. Jain, I.P.; Lal, C.; Jain, A. Hydrogen storage in Mg: A most promising material. Int. J. Hydrogen Energy 2010, 35, 5133–5144. [Google Scholar] [CrossRef]
  13. Jefferson, M. Sustainable energy development: Performance and prospects. Renew. Energy 2006, 31, 571–582. [Google Scholar] [CrossRef]
  14. Sun, Z.; Lu, X.; Nyahuma, F.M.; Yan, N.; Xiao, J.; Su, S.; Zhang, L. Enhancing Hydrogen Storage Properties of MgH2 by Transition Metals and Carbon Materials: A Brief Review. Front. Chem. 2020, 8, 552. [Google Scholar] [CrossRef] [PubMed]
  15. Xu, G.; Shen, N.; Chen, L.; Chen, Y.; Zhang, W. Effect of BiVO4 additive on the hydrogen storage properties of MgH2. Mater. Res. Bull. 2017, 89, 197–203. [Google Scholar] [CrossRef]
  16. Ouyang, L.Z.; Yang, X.S.; Zhu, M.; Liu, J.W.; Dong, H.W.; Sun, D.L.; Zou, J.; Yao, X.D. Enhanced Hydrogen Storage Kinetics and Stability by Synergistic Effects of in Situ Formed CeH2.73 and Ni in CeH2.73-MgH2-Ni Nanocomposites. J. Phys. Chem. C 2014, 118, 7808–7820. [Google Scholar] [CrossRef]
  17. Zhang, X.L.; Liu, Y.F.; Zhang, X.; Hu, J.J.; Gao, M.X.; Pan, H.G. Empowering hydrogen storage performance of MgH2 by nanoengineering and nanocatalysis. Mater. Today Nano 2020, 9, 100064. [Google Scholar] [CrossRef]
  18. Zhang, L.; Sun, Z.; Yao, Z.; Yang, L.; Yan, N.; Lu, X.; Xiao, B.; Zhu, X.; Chen, L. Excellent catalysis of Mn3O4 nanoparticles on the hydrogen storage properties of MgH2: An experimental and theoretical study. Nanoscale Adv. 2020, 2, 1666–1675. [Google Scholar] [CrossRef] [Green Version]
  19. Gao, S.C.; Wang, H.; Wang, X.H.; Liu, H.Z.; He, T.; Wang, Y.Y.; Wu, C.; Li, S.Q.; Yan, M. MoSe2 hollow nanospheres decorated with FeNi3 nanoparticles for enhancing the hydrogen storage properties of MgH2. J. Alloys Compd. 2020, 830, 154631. [Google Scholar] [CrossRef]
  20. Chen, G.; Zhang, Y.; Chen, J.; Guo, X.; Zhu, Y.; Li, L. Enhancing hydrogen storage performances of MgH2 by Ni nano-particles over mesoporous carbon CMK-3. Nanotechnology 2018, 29, 265705. [Google Scholar] [CrossRef]
  21. He, D.; Wang, Y.; Wu, C.; Li, Q.; Ding, W.; Sun, C. Enhanced hydrogen desorption properties of magnesium hydride by coupling non-metal doping and nano-confinement. Appl. Phys. Lett. 2015, 107, 243907. [Google Scholar] [CrossRef]
  22. Yu, H.; Bennici, S.; Auroux, A. Hydrogen storage and release: Kinetic and thermodynamic studies of MgH2 activated by transition metal nanoparticles. Int. J. Hydrogen Energy 2014, 39, 11633–11641. [Google Scholar] [CrossRef]
  23. Lu, C.; Ma, Y.; Li, F.; Zhu, H.; Zeng, X.; Ding, W.; Deng, T.; Wu, J.; Zou, J. Visualization of fast “hydrogen pump” in core-shell nanostructured Mg@Pt through hydrogen-stabilized Mg3Pt. J. Mater. Chem. A 2019, 7, 14629–14637. [Google Scholar] [CrossRef]
  24. Chen, M.; Xiao, X.; Zhang, M.; Liu, M.; Huang, X.; Zheng, J.; Zhang, Y.; Jiang, L.; Chen, L. Excellent synergistic catalytic mechanism of in-situ formed nanosized Mg2Ni and multiple valence titanium for improved hydrogen desorption properties of magnesium hydride. Int. J. Hydrogen Energy 2019, 44, 1750–1759. [Google Scholar] [CrossRef]
  25. Zhu, M.; Wang, H.; Ouyang, L.; Zeng, M. Composite structure and hydrogen storage properties in Mg-base alloys. Int. J. Hydrogen Energy 2006, 31, 251–257. [Google Scholar] [CrossRef]
  26. Liao, B.; Lei, Y.Q.; Chen, L.X.; Lu, G.L.; Pan, H.G.; Wang, Q.D. Effect of the La/Mg ratio on the structure and electrochemical properties of LaxMg3-xNi9 (x = 1.6–2.2) hydrogen storage electrode alloys for nickel–metal hydride batteries. J. Power Source 2004, 129, 358–367. [Google Scholar] [CrossRef]
  27. Liang, G. Synthesis and hydrogen storage properties of Mg-based alloys. J. Alloys Compd. 2004, 370, 123–128. [Google Scholar] [CrossRef]
  28. Zhang, M.; Xiao, X.; Luo, B.; Liu, M.; Chen, M.; Chen, L. Superior de/hydrogenation performances of MgH2 catalyzed by 3D flower-like TiO2@C nanostructures. J. Energy Chem. 2020, 46, 191–198. [Google Scholar] [CrossRef]
  29. Zhang, L.; Ji, L.; Yao, Z.; Yan, N.; Sun, Z.; Yang, X.; Zhu, X.; Hu, S.; Chen, L. Facile synthesized Fe nanosheets as superior active catalyst for hydrogen storage in MgH2. Int. J. Hydrogen Energy 2019, 44, 21955–21964. [Google Scholar] [CrossRef]
  30. Zhang, L.; Cai, Z.; Yao, Z.; Ji, L.; Sun, Z.; Yan, N.; Zhang, B.; Xiao, B.; Du, J.; Zhu, X.; et al. A striking catalytic effect of facile synthesized ZrMn2 nanoparticles on the de/rehydrogenation properties of MgH2. J. Mater. Chem. A 2019, 7, 5626–5634. [Google Scholar] [CrossRef]
  31. Wang, Z.; Zhang, X.; Ren, Z.; Liu, Y.; Hu, J.; Li, H.; Gao, M.; Pan, H.; Liu, Y. In situ formed ultrafine NbTi nanocrystals from a NbTiC solid-solution MXene for hydrogen storage in MgH2. J. Mater. Chem. A 2019, 7, 14244–14252. [Google Scholar] [CrossRef]
  32. Cheng, H.; Chen, G.; Zhang, Y.; Zhu, Y.; Li, L. Boosting low-temperature de/re-hydrogenation performances of MgH2 with Pd-Ni bimetallic nanoparticles supported by mesoporous carbon. Int. J. Hydrogen Energy 2019, 44, 10777–10787. [Google Scholar] [CrossRef]
  33. Zhang, J.; Shi, R.; Zhu, Y.; Liu, Y.; Zhang, Y.; Li, S.; Li, L. Remarkable Synergistic Catalysis of Ni-Doped Ultrafine TiO2 on Hydrogen Sorption Kinetics of MgH2. ACS Appl. Mater. Interfaces 2018, 10, 24975–24980. [Google Scholar] [CrossRef] [PubMed]
  34. Li, L.; Jiang, G.; Tian, H.; Wang, Y. Effect of the hierarchical Co@C nanoflowers on the hydrogen storage properties of MgH2. Int. J. Hydrogen Energy 2017, 42, 28464–28472. [Google Scholar] [CrossRef]
  35. Su, W.; Zhu, Y.; Zhang, J.; Liu, Y.; Yang, Y.; Mao, Q.; Li, L. Effect of multi-wall carbon nanotubes supported nano-nickel and TiF3 addition on hydrogen storage properties of magnesium hydride. J. Alloys Compd. 2016, 669, 8–18. [Google Scholar] [CrossRef]
  36. Plerdsranoy, P.; Thiangviriya, S.; Dansirima, P.; Thongtan, P.; Kaewsuwan, D.; Chanlek, N.; Utke, R. Synergistic effects of transition metal halides and activated carbon nanofibers on kinetics and reversibility of MgH2. J. Phys. Chem. Solids 2019, 124, 81–88. [Google Scholar] [CrossRef]
  37. Ismail, M.; Mustafa, N.S.; Juahir, N.; Yap, F.A.H. Catalytic effect of CeCl3 on the hydrogen storage properties of MgH2. Mater. Chem. Phys. 2016, 170, 77–82. [Google Scholar] [CrossRef]
  38. Cui, J.; Liu, J.; Wang, H.; Ouyang, L.; Sun, D.; Zhu, M.; Yao, X. Mg-TM (TM: Ti, Nb, V, Co, Mo or Ni) core-shell like nanostructures: Synthesis, hydrogen storage performance and catalytic mechanism. J. Mater. Chem. A 2014, 2, 9645–9655. [Google Scholar] [CrossRef]
  39. Grzech, A.; Lafont, U.; Magusin, P.C.M.M.; Mulder, F.M. Microscopic Study of TiF3 as Hydrogen Storage Catalyst for MgH2. J. Phys. Chem. C 2012, 116, 26027–26035. [Google Scholar] [CrossRef]
  40. Ma, L.P.; Kang, X.D.; Dai, H.B.; Liang, Y.; Fang, Z.Z.; Wang, P.J.; Wang, P.; Cheng, H.M. Superior catalytic effect of TiF3 over TiCl3 in improving the hydrogen sorption kinetics of MgH2: Catalytic role of fluorine anion. Acta Mater. 2009, 57, 2250–2258. [Google Scholar] [CrossRef]
  41. Kim, J.W.; Ahn, J.P.; Jin, S.A.; Lee, S.H.; Chung, H.-S.; Shim, J.H.; Cho, Y.W.; Oh, K.H. Microstructural evolution of NbF5-doped MgH2 exhibiting fast hydrogen sorption kinetics. J. Power Source 2008, 178, 373–378. [Google Scholar] [CrossRef]
  42. Jangir, M.; Jain, A.; Yamaguchi, S.; Ichikawa, T.; Lal, C.; Jain, I.P. Catalytic effect of TiF4 in improving hydrogen storage properties of MgH2. Int. J. Hydrogen Energy 2016, 41, 14178–14183. [Google Scholar] [CrossRef]
  43. Zhang, J.; Qu, H.; Yan, S.; Yin, L.R.; Zhou, D.W. Dehydrogenation properties and mechanisms of MgH2-NiCl2 and MgH2-NiCl2-graphene hydrogen storage composites. Met. Mater. Int. 2017, 23, 831–837. [Google Scholar] [CrossRef]
  44. Ismail, M. Influence of different amounts of FeCl3 on decomposition and hydrogen sorption kinetics of MgH2. Int. J. Hydrogen Energy 2014, 39, 2567–2574. [Google Scholar] [CrossRef]
  45. Mao, J.; Guo, Z.; Yu, X.; Liu, H.; Wu, Z.; Ni, J. Enhanced hydrogen sorption properties of Ni and Co-catalyzed MgH2. Int. J. Hydrogen Energy 2010, 35, 4569–4575. [Google Scholar] [CrossRef]
  46. Yao, P.; Jiang, Y.; Liu, Y.; Wu, C.; Chou, K.C.; Lyu, T.; Li, Q. Catalytic effect of Ni@rGO on the hydrogen storage properties of MgH2. J. Magnes. Alloy. 2020, 8, 461–471. [Google Scholar] [CrossRef]
  47. Luo, Q.; Gu, Q.; Liu, B.; Zhang, T.F.; Liu, W.; Li, Q. Achieving superior cycling stability by in situ forming NdH2-Mg-Mg2Ni nanocomposites. J. Mater. Chem. A 2018, 6, 23308–23317. [Google Scholar] [CrossRef]
  48. Jensen, F. Activation energies and the arrhenius equation. Qual. Reliab. Eng. Int. 1985, 1, 13–17. [Google Scholar] [CrossRef]
  49. Xia, G.; Tan, Y.; Chen, X.; Sun, D.; Guo, Z.; Liu, H.; Ouyang, L.; Zhu, M.; Yu, X. Monodisperse magnesium hydride nanoparticles uniformly self-assembled on graphene. Adv. Mater. 2015, 27, 5981–5988. [Google Scholar] [CrossRef]
Figure 1. Volumetric release curves (a) MgH2, MgH2 + 5 wt% MnCl2 samples and XRD patterns of MgH2-MnCl2 composite in in ball-milling state and dehydrogenated state (b).
Figure 1. Volumetric release curves (a) MgH2, MgH2 + 5 wt% MnCl2 samples and XRD patterns of MgH2-MnCl2 composite in in ball-milling state and dehydrogenated state (b).
Nanomaterials 10 01745 g001
Figure 2. XRD patterns (a) and SEM images of purchased Mn (b) and as-prepared Mn particles (c,d).
Figure 2. XRD patterns (a) and SEM images of purchased Mn (b) and as-prepared Mn particles (c,d).
Nanomaterials 10 01745 g002
Figure 3. The volumetric release curves (a,b), isothermal dehydrogenation curves (c,d) of MgH2, MgH2 + Mn, and MgH2 + submicron-Mn samples.
Figure 3. The volumetric release curves (a,b), isothermal dehydrogenation curves (c,d) of MgH2, MgH2 + Mn, and MgH2 + submicron-Mn samples.
Nanomaterials 10 01745 g003
Figure 4. Non-isothermal hydrogenation curves (a), isothermal hydrogenation curves at different temperatures(b,c), and the relevant Arrhenius plots (d) of pure MgH2 and MgH2 + 10 wt% submicron-Mn sample.
Figure 4. Non-isothermal hydrogenation curves (a), isothermal hydrogenation curves at different temperatures(b,c), and the relevant Arrhenius plots (d) of pure MgH2 and MgH2 + 10 wt% submicron-Mn sample.
Nanomaterials 10 01745 g004
Figure 5. XRD patterns of MgH2 + 10 wt% submicron-Mn samples in three different states, ball-milled state (a), dehydrogenated state (b), and hydrogenated state (c).
Figure 5. XRD patterns of MgH2 + 10 wt% submicron-Mn samples in three different states, ball-milled state (a), dehydrogenated state (b), and hydrogenated state (c).
Nanomaterials 10 01745 g005
Figure 6. Isothermal dehydrogenation/hydrogenation cyclic kinetics curves of the MgH2 + 10 wt% submicron-Mn sample.
Figure 6. Isothermal dehydrogenation/hydrogenation cyclic kinetics curves of the MgH2 + 10 wt% submicron-Mn sample.
Nanomaterials 10 01745 g006

Share and Cite

MDPI and ACS Style

Sun, Z.; Zhang, L.; Yan, N.; Zheng, J.; Bian, T.; Yang, Z.; Su, S. Realizing Hydrogen De/Absorption Under Low Temperature for MgH2 by Doping Mn-Based Catalysts. Nanomaterials 2020, 10, 1745. https://doi.org/10.3390/nano10091745

AMA Style

Sun Z, Zhang L, Yan N, Zheng J, Bian T, Yang Z, Su S. Realizing Hydrogen De/Absorption Under Low Temperature for MgH2 by Doping Mn-Based Catalysts. Nanomaterials. 2020; 10(9):1745. https://doi.org/10.3390/nano10091745

Chicago/Turabian Style

Sun, Ze, Liuting Zhang, Nianhua Yan, Jiaguang Zheng, Ting Bian, Zongming Yang, and Shichuan Su. 2020. "Realizing Hydrogen De/Absorption Under Low Temperature for MgH2 by Doping Mn-Based Catalysts" Nanomaterials 10, no. 9: 1745. https://doi.org/10.3390/nano10091745

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