Next Article in Journal
Comparative Analysis on the Performance and Exhaust Gas Emission of Cars with Spark-Ignition Engines
Next Article in Special Issue
Livestock’s Urine-Based Plant Microbial Fuel Cells Improve Plant Growth and Power Generation
Previous Article in Journal
Fractal Characteristics of Pore-Throats Structure and Quality Evaluation of Carbonate Reservoirs in Eastern Margin of Pre-Caspian Basin
Previous Article in Special Issue
Strategies for the Adoption of Hydrogen-Based Energy Storage Systems: An Exploratory Study in Australia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

New Liquid Chemical Hydrogen Storage Technology

1
Institute of Technology for Carbon Neutrality/Faculty of Materials Science and Energy Engineering, Shenzhen Institute of Advanced Technology (SIAT), Chinese Academy of Sciences (CAS), Shenzhen 518055, China
2
Shenzhen Key Laboratory of Energy Materials for Carbon Neutrality, Shenzhen 518055, China
3
Boreskov Institute of Catalysis, SB RAS, Novosibirsk 630090, Russia
4
Key Laboratory of Science and Technology on Wear and Protection of Materials, Lanzhou Institute of Chemical Physics (LICP), Chinese Academy of Sciences (CAS), Lanzhou 730000, China
5
Department of Chemical and Biomolecular Engineering, National University of Singapore, Singapore 117585, Singapore
*
Authors to whom correspondence should be addressed.
Submission received: 8 August 2022 / Revised: 27 August 2022 / Accepted: 27 August 2022 / Published: 31 August 2022
(This article belongs to the Special Issue Advances in Hydrogen Energy Ⅱ)

Abstract

:
The liquid chemical hydrogen storage technology has great potentials for high-density hydrogen storage and transportation at ambient temperature and pressure. However, its commercial applications highly rely on the high-performance heterogeneous dehydrogenation catalysts, owing to the dehydrogenation difficulty of chemical hydrogen storage materials. In recent years, the chemists and materials scientists found that the supported metal nanoparticles (MNPs) can exhibit high catalytic activity, selectivity, and stability for the dehydrogenation of chemical hydrogen storage materials, which will clear the way for the commercial application of liquid chemical hydrogen storage technology. This review has summarized the recent important research progress in the MNP-catalyzed liquid chemical hydrogen storage technology, including formic acid dehydrogenation, hydrazine hydrate dehydrogenation and ammonia borane dehydrogenation, discussed the urgent challenges in the key field, and pointed out the future research trends.

Graphical Abstract

1. Introduction

Hydrogen is a clean, efficient, and renewable energy carrier towards driving a carbon neutral future [1,2]. The World’s leading economies, including the United States, China, Japan, Russia, South Korea, and the European Union, have pumped a vast amount of money into commercializing the newly developed hydrogen energy technology. However, hydrogen is flammable, easy to diffuse, and difficult to store and transport, restricting its large-scale applications. Developing efficient, safe, and long-distance hydrogen storage systems is highly anticipated.
The current hydrogen storage systems mainly include [3,4] high-pressure gas cylinders, cryogenic liquid cylinders, metal hydrides, and porous materials (zeolite, activated carbon, metal-organic frameworks (MOFs), etc.). It is difficult for these systems to meet the requirements of the large-scale applications of hydrogen. For example, high-pressure gas/cryogenic liquid cylinders face problems such as high difficulty in cylinder manufacturing, high energy consumption in gas compression process, and hidden dangers during operation. Metal hydrides face problems such as high cost, high absorption/dehydrogenation temperature and low durability; while porous materials have low hydrogen storage density at room temperature.
In recent years, liquid chemical hydrogen storage technology has attracted much attention as a novel strategy for high-density hydrogen storage and transportation at ambient conditions [5,6]. Since liquid chemical compounds have similar features to the common fuels, such as diesel and gasoline, and their hydrogen is stored forming chemical bonds, high-density hydrogen can be transported, stored and refilled safely by using existing petroleum infrastructure. However, the dehydrogenation of liquid chemical hydrogen storage materials faces large challenges owing to their thermodynamic and kinetic factors. A list of chemical hydrogen storage materials includes different classes of compounds. Among them ammonia borane, formic acid, alcohols, hydrazine hydrate, aromatic compounds, and N-heteroaromatic compounds should be mentioned [6]. Some schemes of the mechanisms of the dehydrogenation reactions are shown in Figure S1 (Supplementary information file). They involve adsorption of the hydrogen carrier molecule on the catalyst, consecutive abstraction of hydrogen atoms from the molecule by the catalyst and desorption of molecular hydrogen. The catalysts facilitate some steps of the reactions due to the formation of surface intermediates.
Fortunately, the supported metal nanoparticles (MNPs) can exhibit an excellent catalytic performance in the hydrogen production from liquid chemical hydrogen storage materials, owing to their small size effect, synergistic effect, and metal–support interaction [7,8,9,10]. In addition, compared with homogeneous catalysts, the supported MNPs are easy to separate and recycle, providing a promising route to reducing the catalyst cost for commercial applications.
Herein, we focused on the recent important research progress in catalytic hydrogen production from chemical hydrogen storage materials by using the high-performance MNP catalysts, including formic acid dehydrogenation, hydrazine hydrate dehydrogenation and ammonia borane solution dehydrogenation, discussed the urgent challenges in this key field, and prospected the future research trends.
Physical properties of the considered compounds are presented in Table S1 (Supplementary information file). They contain a significant content of hydrogen corresponding to 4.4 wt.% in formic acid and about 8 wt.% in hydrazine hydrate or ammonia borane dissolved in water. The decomposition of these compounds also produces by-products. These by-products could be recycled and rehydrogenated. This provides efficient utilization of the hydrogen storage technology.

2. Formic Acid Dehydrogenation

Formic acid (HCOOH) is a promising liquid chemical hydrogen material owing to its advantages of low toxicity, high hydrogen content (4.4 wt.%), high stability, and easy availability through biomass processing and CO2 hydrogenation [11,12,13]. The stored hydrogen in formic acid can be released through two pathways, dehydrogenation (1) and dehydration (2) [14]:
HCOOH → H2 + CO2, ΔG298K = −48.8 kJ·mol−1
HCOOH → H2O + CO, ΔG298K = −28.5 kJ·mol−1
Given that dehydration (2) not only reduces hydrogen production but also generates carbon monoxide (which is toxic to dehydrogenation and fuel cell catalysts), the key to the practical application of formic acid dehydrogenation lies in the development of high-performance MNP catalysts. These catalysts can be used for gas-phase or liquid-phase dehydrogenation and can be easily separated from the reaction mixture as compared to dissolved metal complexes.

2.1. Monometallic Catalysts

Au bulk catalysts are inactive for the formic acid dehydrogenation. Surprisingly, Au NPs anchored on specific supports are highly active for this reaction, due to its surface electronic structure change resulting from the strong metal–support interaction. Xu et al. investigated the relationship between Au NPs and spherical SiO2, and found that Au NPs coated with amine-functionalized SiO2 spheres showed high dehydrogenation rates, releasing CO-free hydrogen at 363 K [15]. Au NPs coated with SiO2 spheres without amine modification possessed much poor catalytic activity. Cao et al. anchored ~1.8 nm Au NPs on the acid-resistant ZrO2 by regulating the deposition and activation processes [16]. The resultant Au/ZrO2 catalyst was highly active for the dehydrogenation of formic acid/triethylamine mixture, achieving an initial turnover frequency (TOF) value of 1590 h−1 at 323 K, without any traces of CO. The TOF value maintained at 1200 h−1 even after 100 h of continuous operation at 323 K. Zhang et al. loaded Au NPs on the Schiff base modified SiO2 surface (Figure 1), which showed fast dehydrogenation of high concentrations of formic acid in the absence of additives, owing to the excellent C-H bond activation performance of the interfaces between the protonated Schiff base and Au atoms [17]. The initial TOF value reached as high as 4368 h−1 at 323 K and the formic acid concentration of 10 mol L−1.
For the gas-phase reaction over catalysts with approximately the same mean size of the Au particles (2.4–3 nm) supported on different oxide supports, Zacharska et al. also demonstrated that the hydrogen production from formic acid strongly depends on the nature of the support [18]. An Au/Al2O3 catalyst showed a better performance in the reaction. The catalytic properties of this catalyst could be further improved by doping with potassium ions [19]. Recently, Bulushev et al. showed that the doping of an Au catalyst supported on carbon nanotubes with cucurbituril led to a significant decrease of the reaction temperature by 110 K due to the hydrogen abstraction from formic acid by carbonyl portals of cucurbituril [20].
Compared to Au catalysts, Pd NPs were explored extensively for the formic acid dehydrogenation. Yu et al. encapsulated ultrafine Pd NPs (~1.5 nm) inside the pores of Silicalite (MFI), which were highly active for the dehydrogenation of formic acid/sodium formate mixture, and the TOF value reached 3027 h−1 at 323 K with 100% hydrogen selectivity [21]. Yu et al. further found that the simultaneous encapsulation of Pd NPs and Ni(OH)2 within zeolite S-1 (Figure 2) boosted the catalytic activity of Pd NPs significantly [22]. The initial TOF value reached 5803 h−1 at 333 K, when the optimal molar ratio of Pd and Ni(OH)2 was 4/1.
In Xu’s group, a series of highly catalytically active Pd NPs were synthesized for the dehydrogenation of formic acid: (a) by using an NaOH-assisted NaBH4 reduction strategy, highly dispersed Pd NPs were immobilized on porous carbon MSC-30, which possessed a TOF value of 2623 h−1 at 323 K and 100% hydrogen selectivity [23]; (b) by using anhydrous methanol as a surface capping agent and reducing agent, ultrafine Pd NPs (~1.4 nm) were loaded on porous carbon Vulcan XC-72R, which showed an initial TOF value of 4452 h−1 at 323 K [24]; (c) by functionalizing graphene oxide with p-phenylenediamine (PDA), highly dispersed Pd NPs (~1.5 nm) were immobilized on the surface of graphene, which showed a TOF value of 3810 h−1 at 323 K; even if pure formic acid was used, this catalyst maintained high catalytic activity, achieving a TOF value of 1500 h−1 at the same temperature [25]; (d) by functionalizing porous carbon MSC-30 with abundant N sites through the pyrolysis of urea at low-temperature, ultrafine Pd NPs (~1.4 nm) were anchored on the surface of MSC-30, which exhibited a record TOF value of 8414 h−1 at 333 K and there was no change in activity and selectivity after 15 cycles [26]; and (e) highly dispersed Pd NPs were immobilized on Fe-N-C support, which produced a high TOF value of 7361 h−1 at 323 K without any traces of CO, owing to the strong interaction between Pd and Fe-N sites [27]. Cao et al. immobilized Au, Ir, Pt, Ru, Rh, Pd and other monometallic NPs on the surface of reduced graphite oxide (rGO) for the dehydrogenation of potassium formate solution and found that only Pd NPs showed extremely high initial TOF, up to 11,299 h−1 at 353 K—the other particles were inactive [28]. The lattice mismatch between the Pd nanoparticles and the surface of rGO was considered to be the main reason for the enhanced catalytic activity. Cao et al. also used the strong interaction between pyridine nitrogen sites and Pd to anchor Pd NPs on the surface of N-doped carbon [29]. The obtained Pd/N-doped carbon catalyst efficiently catalyzed the dehydrogenation of additive-free formic acid at room temperature, achieving a TOF of 5530 h−1 without any traces of CO.
For the gas-phase reaction, Bulushev et al. prepared supported Pd catalysts with single Pd atoms using N-doped carbon [30,31] and found that single-atom Pd sites showed a TOF value three-to-four times higher than that of Pd atoms in N-free carbon-supported Pd particles and support-free Pd catalyst. Jia et al. found that doping of dispersed Pd catalysts on different supports with potassium ions leads to an increase of the TOF values by a factor of up to 33 [32]. The TOF value reached 3600 h−1 at 353 K for the K-doped Pd/C catalyst with a mean particle size of 3.6 nm. Bulushev and Bulusheva recently discussed the role of supported single metal atoms in the reaction and showed that they can sometimes perform better than NPs [33]. This could also be valid for some Ru or Ir immobilized complexes [34].
Notably, the non-noble metal catalysts are more promising for commercial applications owing to their cost. However, the non-noble metal catalysts often exhibit poor catalytic activity in the dehydrogenation of formic acid, as they can be etched easily by formic acid during the dehydrogenation reaction. By pyrolysis of the cobalt acetate, phenanthroline and Vulcan XC-72R mixture at high temperature, Beller et al. synthesized carbon-supported Co active sites with good resistance to acid corrosion, which possessed excellent catalytic activity and stability for the dehydrogenation of formic acid at 371 K, achieving an average H2 production rate of 180 mL gcat−1 h−1 after continuous operation for 6 days [35], owing to the highly dispersed CoNx active centers. Furthermore, Beller et al. synthesized the N-doped carbon-supported Co single-atom catalysts (Co-N-C) through the pyrolysis of ZnCo-ZIFs at high temperature, which exhibited high activity in the catalytic hydrogen production from formic acid in propylene carbonate solvent [36], achieving an average H2 production rate of 319.2 mL gcat−1 h−1 (initial TOF=357 h-1) much higher than that of Co NPs prepared by the same process.

2.2. Bimetallic and Multimetallic Catalysts

Given that the intrinsic catalytic capability of metals can be enhanced through the synergistic effect of different metal elements, the bimetallic or multimetallic catalysts are a promising choice for the highly efficient dehydrogenation of formic acid. As early as in 2008, Xing et al. prepared AgPd/C and AuPd/C for the dehydrogenation of a formic acid/sodium formate mixture, which showed much higher catalytic activities than those of Ag/C, Au/C and Pd/C [37]. By introducing CeO2 as a co-catalyst, the TOF of AuPd/C was further improved, reaching 832 h−1 at 373 K. This group further synthesized the AuPd@Cu/C core-shell catalyst based on the weak CO adsorption characteristic of the Au surface, which showed an average H2 production rate of 90 mL g−1 min−1 at 365 K, with a small trace of CO [38]. Tsang et al. synthesized various core-shell catalysts by depositing Pd on the surface of MNPs, such as Ag, Rh, Au, Ru and Pt, respectively [39]. Compared with the monometallic catalysts, the catalytic capabilities of the core-shell catalysts were enhanced significantly, and Ag@Pd showed a TOF value of 252 h−1 at 323 K, the highest activity among all the core-shell catalysts. The enhanced catalytic capability was attributed to the strong electronic interaction between core metal and shell metal. Chen et al. synthesized AuPd NPs on porous carbon by using a Mg2+ ion-assisted low-temperature reduction strategy, which possessed an initial TOF value of 1120 h−1 for the dehydrogenation of formic acid without a trace of CO [40].
The support can control the size, distribution and metal–support interaction of MNPs, and then boost their catalytic capabilities. In an early study, Xu et al. encapsulated highly dispersed AuPd NPs inside the ethylene diamine-modified ED-MIL-101, which showed enhanced catalytic activity for the dehydrogenation of formic acid with good resistance to CO poisoning, in comparison with AuPd NPs inside MIL-101 [41]. Since then, large numbers of bimetallic NPs were reported. For example, Yan et al. synthesized AuPd and MnOx NPs on a composite support of ZIF-8/graphene, which showed a high activity for the catalytic hydrogen production from pure formic acid (Figure 3), achieving an initial TOF value of 328 h−1 at room temperature [42]. This group further immobilized AuPd NPs (~2.4 nm) on amine-modified N-doped graphene, which were highly active for the dehydrogenation of formic acid, achieving an initial TOF value as high as 4445.6 h−1 at room temperature [43]. The enhanced catalytic activity was attributed to the low adsorption capacity of Pd surface after alloying. Xu et al. immobilized AgPd NPs (~2.5 nm) on a ZrO2/C/rGO support, which showed a high TOF value of 4500 h−1 at 323 K for the dehydrogenation of the formic acid/sodium formate mixture [44].
New strategies have also been developed to control the size, distribution and metal-support interaction of bimetallic NPs, and then improved their catalytic performance. Xu et al. developed a non-noble metal sacrificial approach, based on the different solubility between noble (Ag/Au/Pd) and non-noble (Co) metals in acid solution (H3PO4), to immobilize highly dispersed AgPd and AuPd NPs on reduced graphene oxide (rGO), respectively [45,46]. The obtained AgPd/rGO and AuPd/rGO exhibited high activity and selectivity for the dehydrogenation of the formic acid/sodium formate mixture, achieving TOF values of 2739 h−1 and 4840 h−1 at 323 K without the releasing of CO, respectively. In addition, Xu et al. developed a solid-ligand-assisted approach, based on the strong binding ability of N sites to Au3+ and Rh3+ ions, to immobilize the bulk immiscible AuRh NPs (~1.6 nm) on the surface of porous carbon with abundant N and O groups [47]. The pure Au or Rh NPs obtained by using the similar approach were inactive in the dehydrogenation of the formic acid/sodium formate mixture. Surprisingly, the AuRh NPs showed outstanding activity and the TOF value reached 2297 h−1 at 333 K (Figure 4).
In addition to bimetallic NPs, multimetallic NPs have exhibited much high catalytic activity and selectivity for the dehydrogenation of formic acid. Jiang et al. immobilized CoPdAu NPs and NiPdAu NPs on porous carbon Vulcan XC-72R by using a wet-chemical reduction method, respectively [48,49]. Since the second and third components modified the electronic structure of Pd, the CoPdAu/C and NiPdAu/C catalysts showed good catalytic activities, and the TOF values reached 80 and 12.4 h−1 at room temperature, respectively. Zahmakiran et al. prepared a PdNiAg/C catalyst with a particle size of ~5.6 nm, which exhibited a TOF value of 85 h−1 at 323 K in the dehydrogenation of formic acid/sodium formate mixture [50]. Luo et al. synthesized CoAgPd NPs (~2.8 nm) by oleic acid as a reducing agent and stabilizer [51]. Such ternary NPs achieved a TOF value of 110 h−1 for the dehydrogenation of the formic acid/sodium formate mixture.
A summary Table S2, including the TOF values for the hydrogen production from formic acid and mean metal particle sizes of the most active catalysts, is presented in Supplementary information file. It is seen that to provide high activity, the mean particle size should correspond to about 2 nm. Evidently, not all the mentioned catalysts may provide stable activity in continuous flow, which could be desirable for large-scale applications.

3. Hydrazine Hydrate Dehydrogenation

Hydrazine hydrate (N2H4·H2O) is often used as a reducing agent, anti-oxidant, or fuel. Given its high hydrogen content (8.0 wt.%), outstanding thermal stability and liquid state at room temperature, hydrazine hydrate can be considered as a promising hydrogen storage material [52,53]. At the same time, this compound is toxic and it is difficult to produce hydrazine hydrate back from the by-product N2.
If we use a catalyst, the hydrogen stored in hydrazine hydrate will release through two pathways: complete decomposition (3) or incomplete decomposition (4):
N2H4 → N2 + 2H2
N2H4 → 1/3N2 + 4/3NH3
Complete decomposition (3) is an ideal route for producing hydrogen. Incomplete decomposition (4) produces ammonia, which not only reduces the yield of hydrogen but also poisons the fuel cell catalysts and even the environment. In order to achieve a commercial application of hydrazine hydrate as a chemical hydrogen material, the high-performance catalysts for its complete decomposition are highly desirable.

3.1. Monometallic Catalysts

In an earlier study, Xu et al. prepared a series of noble (Rh, Ru, Ir, Pt, Pd) and non-noble (Co, Cu, Ni, Fe) MNPs by using cetyltrimethylammonium bromide (CTAB) as a capping agent for the decomposition of hydrazine hydrate [54]. The experimental results indicated that Rh NPs have the highest hydrogen selectivity, up to 44%; Ru, Ir, Co NPs have a certain activity with the selectivity of <7%; other NPs, such as Pt, Pd, Cu, Ni, and Fe, were inactive. The authors proposed that Ru, Ir, and Co NPs tend to break N-N bonds rather than N-H bonds, and thus promote the occurrence of incomplete decomposition. Zhang et al. prepared an Ni/Al2O3 nanocatalyst by using co-precipitation method combined with high-temperature reduction and Ni-Al hydrotalcite as a raw material [55]. The Ni/Al2O3 catalyst exhibited a selectivity of 93% for the complete decomposition of hydrazine hydrate at 303 K, and this selectivity was increased to 96% through the introduction of KOH. This group further used amorphous CeO2 as a support to anchor Ni NPs (Ni/CeO2) [56]. When the molar ratio of Ni to CeO2 was 1/0.08, the Ni/CeO2 catalyst exhibited good catalytic activity (TOF = 56.1 h−1, 303 K) and selectivity (99%). Similarly, Varma et al. prepared a Ni/CeO2 nanocatalyst by a solution combustion method. When the loading of Ni reached 6 wt.%, Ni/CeO2 exhibited good catalytic activity (TOF = 34.0 h−1, 323 K) and H2 selectivity (100%) [57].
Notably, although monometallic NPs often show poor catalytic performance for the complete decomposition of hydrazine hydrate, their catalytic performance can be improved by doping, increasing the number of basic sites and strengthening the metal–support interactions [58,59]. Wang et al. encapsulated Ni NPs inside titanate nanotubes, which exhibited a TOF value of 96.0 h−1 at 333 K (1 M NaOH) and 100% H2 selectivity, owing to the high dispersion of Ni NPs and the enhanced interaction between Ni and titanate [58]. Lu et al. confined Ni-CeO2 nanowires on SiO2 by the reverse micelle method, which possessed high catalytic activity (TOF = 219.5 h−1, 343 K) and H2 selectivity (100%), due to the enhanced electronic synergistic effect between Ni NPs and the abundant oxygen vacancies of CeO2 [59].

3.2. Bimetallic Catalysts

Compared to the studies with monometallic NPs, the studies of Ni-based alloy NPs, especially RhNi and PtNi catalysts, have contributed too much to the area of decomposition of hydrazine hydrate. In 2009, Xu et al. synthesized ~3 nm RhNi NPs by using the surface capping agent CTAB, which showed 100% H2 selectivity when the molar ratio of Rh to Ni reached 4/1 [60]. In addition, their H2 selectivity relied significantly on the reaction temperature: the H2 selectivity was only 28% at room temperature, however, it increased to 100% when the temperature reached 323 K [61]. In 2010, this group found that PtNi alloy NPs are highly active for the complete decomposition of hydrazine hydrate, though the Ni and Pt NPs are often inactive [62]. Since then, highly dispersed PtNi NPs (~2.4 nm) were synthesized by NaOH-assisted NaBH4 reduction process, which exhibited a TOF value of 150 h−1 [63]; ultrafine PtNi NPs were immobilized on 3-dimensional (3D) N-doped graphene frameworks, which achieved 100% H2 selectivity and a recorded TOF value of 943 h−1 at 303 K, and such 3D N-doped graphene frameworks were synthesized by the self-assembly of graphene layers with simultaneous nitrogen doping via crosslinking of graphene oxide with melamine formaldehyde resin under hydrothermal conditions followed by carbonization [64]; PtNi NPs were anchored on ZrO2/porous carbon/graphene, which showed 100% H2 selectivity and a high TOF value of 1920 h−1 at 323 K (Figure 5) [65].
The strong metal–support interaction can significantly enhance the catalytic activity and selectivity of bimetallic NPs for the complete decomposition of hydrazine hydrate. In one example, Lu et al. encapsulated ultrafine PtNi NPs inside MIL-101 and found that the confinement of pores can significantly enhance the interaction between PtNi and MIL-101, thus boosting their catalytic activity (TOF = 140 h−1, 303 K), selectivity (100%) and stability (not less than 20 cycles) [66]. In another example, this group immobilized PtNi NPs on the surface of TiO2-modified Ti3C2Tx nanosheets, which exhibited a high TOF value of 1220 h−1 at 323 K, due to the enhanced synergistic effect between PtNi atoms and the abundant functional groups on Ti3C2Tx nanosheet [67]. In addition, Wang et al. prepared N-doped carbon spheres by co-pyrolysis of SiO2 spheres and dopamine for immobilizing PtNi NPs [68]. The obtained PtNi/N-doped carbon showed a high TOF value of 1602 h−1 at 323 K. N-doped graphene hydrogels with abundant N sites were also used for anchoring PtNi-CeOx NPs [69]. The PtNi-CeOx/N-doped graphene hydrogels achieved a high TOF value of 3064 h−1 at 323 K for the dehydrogenation of hydrazine hydrate.
PdNi and IrNi alloys are another choice for the efficient catalytic hydrogen production of hydrazine hydrate. Xu et al. found that Pd0.4Ni0.6 NPs exhibited the best catalytic activity and selectivity in all their metal ratios, achieving a hydrogen selectivity of 82% at 323 K [70]. Lu et al. anchored IrNi NPs on the surface of lanthanum oxycarbonate, which showed a high TOF value of 1250 h−1 at 323 K, as the abundant basic sites on the surface of lanthanum oxycarbonate facilitated the cleavage of N-H bonds [71].
In addition to Ni-based alloy, Co-based alloys also showed high catalytic performance. Yan et al. prepared Co0.65Pt0.30(CeOx)0.05 nanocatalysts by a wet-chemical method, which showed H2 selectivity of 72.1% at room temperature and a TOF value of 194.8 h−1 [72]. Lu et al. prepared PtCo/La(OH)3 catalyst, based on the outstanding synergistic effect between PtCo and the abundant basic sites on La(OH)3, which showed 100% H2 selectivity and a high TOF value of 2400 h−1 at 323 K [73].
Heteroatom doping can change the electronic structure of active metals and then improve their catalytic performance. Zhang et al. introduced B element into RhNi to form a RhNiB catalyst, which possessed an eight times higher catalytic activity than that of RhNi at room temperature and alkaline conditions, and 100% H2 selectivity [74]. Luo et al. synthesized graphene-supported RhNiP NPs, which showed a five times higher catalytic activity than that of graphene-supported RhNi NPs, achieving a TOF value of 471 h−1 at 323 K and alkaline conditions [75]. Liu et al. prepared graphene-supported PtNiP NPs by using sodium hypophosphite as the P source (Figure 6) [76]. Due to the synergistic effect resulting from P doping and the strong metal–graphene interaction, the Pt0.3Ni0.7P/rGO catalyst showed high catalytic activity for the dehydrogenation of hydrazine hydrate, demonstrating a TOF value of 742 h−1 at 323 K.
In order to obtain low-cost catalysts, the researchers have made great efforts to design and prepare alloy catalysts containing two non-noble elements. Lu et al. prepared Cr2O3-modified NiFe NPs for the dehydrogenation of hydrazine hydrate to produce hydrogen [77]. Compared to monometallic Fe and Ni catalysts, NiFe-Cr2O3 showed higher catalytic activity, and the TOF value and H2 selectivity reached 893.5 h−1 and 100% at 343 K, respectively. The presence of Cr2O3 not only restricted the aggregation of NiFe NPs but also enhanced synergistic effect between NiFe and Cr2O3. Varma et al. prepared a NiCu/CeO2 catalyst by one-step solution combustion method for the dehydrogenation of hydrazine hydrate [78]. When the alkali concentration was 6 M and the temperature was 323 K, the TOF value reached 1450 h−1—much higher than that of Ni/CeO2. Wen et al. immobilized highly dispersed NiCo NPs (~4 nm) on the surface of ultrathin layered NiO-CoOx for the catalytic hydrogen production from hydrazine hydrate [79]. The NiCo/NiO-CoOx exhibited the TOF of 5.49 h−1 and 100% H2 selectivity. The interaction between metals and the high density of basic sites was considered as the main reason for its enhanced catalytic performance.
A summary Table S3, including the TOF values for the hydrogen production from hydrazine hydrate and mean metal particle sizes of the most active catalysts, is presented in Supplementary information file. It is seen that to provide high activity, the mean particle size should normally correspond to about 2.5 nm. There are some exceptions for the catalysts with non-noble metal catalysts.

4. Ammonia Borane Dehydrogenation

Ammonia borane (NH3BH3) is another promising chemical hydrogen storage material owing to its high hydrogen content (19.6 wt.%), low molecular weight (30.9 g mol−1), and good solubility in water, and other solvents. Ammonia borane can release hydrogen through thermal decomposition, hydrolysis (5) or methanolysis (6) with the use of an appropriate catalyst. In the present review, we do not consider thermal decomposition. The key to promoting the practical application of ammonia borane as hydrogen source lies in the development of efficient and stable catalysts.
NH3BH3 + 2H2O → NH4BO2 + 3H2
NH3BH3 + 4CH3OH → NH4B(OCH3)4 + 3H2
For practical utilization of water solutions of ammonia borane for hydrogen production, it should be considered that it is difficult to convert the by-product NH4BO2 back to ammonia borane due to the problems of converting the B-O bonds into the B-H bonds.

4.1. Hydrolysis of Ammonia Borane

4.1.1. Monometallic Catalysts

In 2006, Xu et al. reported for the first time that Pt, Ru and Rh NPs possessed high catalytic activity for the hydrolysis of ammonia borane under ambient conditions [80], which attracted much attention in the scientific community on the application of ammonia borane as a practical chemical hydrogen storage material. Since then, various high-surface-area solid materials, such as porous carbon, mesoporous SiO2, metal oxides, and MOFs, have been explored as supports to anchor highly dispersed MNPs for the catalytic hydrogen production from ammonia borane. For example, a double-solvents approach was developed to encapsulate small and highly dispersed Pt NPs (~1.8 nm) inside the pores of MIL-101, which possessed high catalytic activity for the ammonia borane hydrolysis, achieving a TOF value of 414 min−1 [81]. Ultrasmall and uniformly dispersed Ru NPs were immobilized on the graphene, which showed a high initial TOF value (600 min−1) at room temperature [82]. Pt and Rh NPs were immobilized on the surface of activated carbon nanotubes, which efficiently catalyzed the hydrolysis of ammonia borane to produce hydrogen, achieving high TOF values of 567 min−1 and 760 min−1, respectively [83,84]. Highly dispersed Pt NPs were anchored on the surface of graphene by a non-noble metal sacrificial approach [85], which showed a high TOF value of 284 min−1 at 298 K (Figure 7). The MOF-derived Ru/Cu@C catalysts exhibited a TOF value of 97 min−1 at room temperature for ammonia borane hydrolysis [86]. Ultrasmall Pt clusters were immobilized within the charged ionic organic molecular cage I-cage-Cl based on the electrostatic repulsion effect, and the obtained Pt@I-cage-Cl catalyst showed a unique on–off effect on the hydrolysis of ammonia borane with a TOF value of 115 min−1 at 300 K [87].
Notably, non-noble metal NPs can exhibit good catalytic activity in the hydrolysis of ammonia borane, which is of great importance for practical applications. Lu et al. found that the catalytic activity of Co NPs on polyethyleneimine-modified graphene oxide was promoted by the synergistic effect of Co and amino groups, showing a TOF value of 39.9 min−1 at room temperature [88]. Shaabani et al. used 3D graphene framework-supported Ni NPs to perform hydrolytic ammonia borane decomposition, and achieved a TOF value of 41.7 min−1 at room temperature [89]. However, non-noble metals often show unstable catalytic activities, as they are easily oxidized in air. In this regard, the researchers synthesized CoB [90], CoP [91], and Ni2P [92] catalysts, which showed excellent anti-oxidation ability and catalytic activity in the catalytic hydrogen production from ammonia borane.

4.1.2. Bimetallic Catalysts

Compared to monometallic catalysts, bimetallic catalysts often show better catalytic activity and selectivity in the hydrolysis of ammonia borane [93,94,95,96,97]. For example, Rakap et al. reported that the polyvinylpyrrolidone (PVP)-stablized PdRh catalyst possessed a TOF value of 1333 min−1 for the hydrolysis of ammonia borane at room temperature, which was much higher than that of monometallic Pd or Rh [94]. Cho et al. synthesized octahedral Au@Pt core-shell NPs by using polydiallyldimethylammonium (PDDA) as stabilizer through polyol reduction method [95]. When Pt/Au molar ratio was optimized to be 1/1, the Au@Pt nanocatalyst exhibited the best catalytic activity with an activation energy of 44.28 kJ mol−1. Yu et al. encapsulated ultrafine Rh0.8Ru0.2 NPs inside Si-OH-modified MFI zeolites (Figure 8), which showed high catalytic activity for the hydrolysis of ammonia borane, and the TOF value reached 1006 min−1 at room temperature [96].
The introduction of a non-noble metal into a noble metal catalyst can enhance the catalytic performance, while reducing the catalyst cost. Li et al. immobilized RuCu NPs (~5.4 nm) on the surface of TiO2@C-N for the hydrolysis of ammonia borane, which showed a high TOF value of 626 min−1 at room temperature (based on Ru metal), when the molar ratio of Ru/Cu was 3/2 [97]. Zhang et al. immobilized Pt0.4Ni0.6 NPs (~2.6 nm) on the graphene for the hydrolysis of potassium borohydride/ammonia borane mixture, which showed a TOF value of 408 min−1 at 308 K [98].
By controlling the size, distribution and metal–support interaction, the catalytic performance of the bimetallic catalysts can be improved significantly. Ren et al. immobilized AuCo NPs on carbon nitride (CN) by impregnation for the catalytic hydrolysis of ammonia borane [99]. With light radiation, the AuCo/CN catalysts showed excellent catalytic activity (TOF = 2897 h−1) at room temperature. The synergistic effect between Au and Co, and the Motty–Schottky effect at the metal-semiconductor interface were responsible for the enhanced catalytic performance. By using the double-solvents approach combined with a high-concentration NaBH4 reduction strategy, Xu et al. encapsulated AuNi and AuCo NPs inside the pores of MIL-101, respectively [100,101]. Owing to the small size effect and high dispersion, AuNi@MIL-101 and AuCo@MIL-101 showed catalytic activities for the hydrolysis of ammonia borane. The optimal TOF values reached 66.2 and 23.5 min−1 at room temperature, respectively. Astruc et al. encapsulated ultrafine Ni2Pt NPs inside the pores of ZIF-8 by a deposition-precipitation process, which exhibited excellent catalytic activity for the hydrolysis of ammonia borane, owing to the enhanced small size effect, distribution, and metal–support interaction [102]. The optimal TOF value reached 2222 min−1. Ma et al. immobilized highly dispersed CoNi NPs on the surface of α-MoC through a strong metal–support interaction, which possessed a high catalytic activity for the hydrolysis of ammonia borane [103]. The TOF value reached 321.1 min−1 at room temperature; one of the best TOF value showed by non-noble metals.

4.2. Methanolysis of Ammonia Borane

In recent years, researchers have explored a series of MNPs for the highly efficient methanolysis of ammonia borane, which provides a new method for the practical application of chemical hydrogen storage. In one case, Wang et al. loaded Ru NPs (~5 nm) on the surface of montmorillonite (MMT) by an ion-exchange method with hydrogen reduction, which showed an outstanding H2 release rate up to 29 L min−1 g−1 for the methanolysis of ammonia borane [104]. Such a catalyst showed good durability, as Ru NPs were firmly stabilized inside the montmorillonite interlayers. Xu et al. used the imine molecular cage CC3R to immobilize Rh NPs, which showed excellent catalytic activity and selectivity for the methanolysis of ammonia borane, achieving a high TOF value of 215.3 min−1 at room temperature [105]. The enhanced catalytic activity of Rh NPs was attributed to the good solubility of CC3 cages, which facilitated the homogenization of the heterogeneous Rh NPs with high dispersibility in reaction solvents. By N2 flowing combined with water or methanol washing, the catalysts were easy to recycle after use.
Given the hydrophobicity of organic molecular cages, Xu et al. developed a reverse double-solvents approach to encapsulate ultrafine metal clusters inside the organic molecular cage RCC3 [106]. This approach involved three steps: (a) the hydrophobic RCC3 cages were dispersed into H2O solution firstly, and then a small amount of hydrophobic solvent (CH2Cl2) containing metal precursors was slowly dropped; (b) the RCC3 cages adsorbed the CH2Cl2 containing metal precursors into their hydrophobic cavities; and (c) by reducing with the high-concentration NaBH4, the metal precursors inside RCC3 were converted to metal clusters. The Pd@RCC3 catalyst showed a high TOF value of 176 min−1 at 303 K for the methanolysis of ammonia borane (Figure 9); one of the best activities among Pd catalysts. This group further encapsulated ultrastable Pd clusters inside the reticular Schiff base nanospaces, which showed a similar TOF value of 177 min−1 at 303 K for the methanolysis of ammonia borane [107]. The clean surface of metal clusters and the synergistic effect between Schiff base sites and metal atoms were responsible for the enhanced catalytic performance. In addition, Sun et al. encapsulated metal clusters inside ionic molecular cages and racemic molecular cages, which also showed excellent catalytic activities and selectivity for the catalytic methanolysis of ammonia borane [108,109].
A summary Table S4 including the TOF values for the hydrogen production from ammonia borane and mean metal particle sizes of the most active catalysts is presented in Supplementary information file. It is seen that to provide high activity, the mean particle size should correspond to about 2 nm.

5. Conclusions and Outlook

Due to the great potential of utilizing existing infrastructure for high-density hydrogen storage and transportation at ambient conditions, chemical hydrogen storage technology has received extensive attention from academia and industry. In recent years, the chemists and materials scientists have designed and synthesized a series of MNP catalysts with excellent catalytic activities, selectivity and stability to solve the challenges in this field, such as the difficulty in the dehydrogenation of chemical hydrogen storage materials, deactivation of the catalysts provided by by-products adsorption on the catalyst, metal sintering and leaching, poor reversibility, necessity to sometimes clean the produced hydrogen and the high cost of catalysts. The main principles for catalyst design and preparation involve: (a) reducing the size of MNPs and increasing the number of surface atoms, forcing MNPs to have more active sites; (b) forming alloys, and the synergistic effect between different metal elements can promote the surface electronic structure of the active metals; (c) enhancing the metal–support interaction, tailoring the surface electronic structures of MNPs, and preventing MNPs from re-growth during catalytic reactions, (d) improving the distribution of MNPs to make them more accessible; and (d) selecting appropriate additives to provide a suitable catalytic environment.
Future research in this field should continue to focus on: (a) improving the catalytic activity, selectivity, and stability of MNP catalysts, and reducing their costs to meet the demands of practical applications; (b) developing new hydrogen storage chemicals containing high hydrogen content and the corresponding highly active dehydrogenation catalysts; and (c) developing highly efficient chemical hydrogen storage systems and energy conversion devices to improve the energy utilization efficiency. We surely believe that the scientists will make great breakthroughs in this field and will then construct a low-carbon and sustainable society.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/en15176360/s1, Figure S1: Schematic representation of the mechanisms of (a) formic acid, (b) hydrazine hydrate and (c) hydrolytic ammonia borane decomposition; Table S1: Properties of the considered hydrogen storage materials; Table S2: The most active catalysts in the formic acid dehydrogenation; Table S3: The most active catalysts in the hydrazine hydrate dehydrogenation; Table S4: The most active catalysts in the ammonia borane dehydrogenation in the presence of water or methanol.

Author Contributions

X.Y. proposed the topic and wrote the paper. D.A.B., J.Y. and Q.Z. revised the paper. All authors have read and agreed to the published version of the manuscript.

Funding

X.Y. thanks the Joint Interdisciplinary Research Project of SIAT (E25427), and D.A.B. highly appreciates the financial support from the Russian Science Foundation (grant 17-73-30032).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Majumdar, A.; Deutch, J.M.; Prasher, R.S.; Griffin, T.P. A framework for a hydrogen economy. Joule 2021, 5, 1905–1908. [Google Scholar] [CrossRef]
  2. Song, F.; Zhang, T.; Zhou, D.; Sun, P.; Lu, Z.; Bian, H.; Dang, J.; Gao, H.; Qian, Y.; Li, W.; et al. Charge Transfer of Interfacial Catalysts for Hydrogen Energy. ACS Mater. Lett. 2022, 4, 967–977. [Google Scholar] [CrossRef]
  3. Wu, C.; Li, Y.; Li, Y. Hydrogen Storage and Transportation; Chemical Industry Press Co., Ltd.: Beijing, China, 2020. [Google Scholar]
  4. Chen, Z.; Kirlikovali, K.O.; Idrees, K.B.; Wasson, M.C.; Farha, O.K. Porous Materials for Hydrogen Storage. Chem 2022, 8, 693–716. [Google Scholar] [CrossRef]
  5. Onishi, N.; Laurenczy, G.; Beller, M.; Himeda, Y. Recent progress for reversible homogeneous catalytic hydrogen storage in formic acid and in methanol. Coord. Chem. Rev. 2018, 373, 317–332. [Google Scholar] [CrossRef]
  6. Lang, C.; Jia, Y.; Yao, X. Recent advances in liquid-phase chemical hydrogen storage. Energy Storage Mater. 2020, 26, 290–312. [Google Scholar] [CrossRef]
  7. Yang, X.; Xu, Q. Gold-containing metal nanoparticles for catalytic hydrogen generation from liquid chemical hydrides. Chin. J. Catal. 2016, 37, 1594–1599. [Google Scholar] [CrossRef]
  8. Onishi, N.; Iguchi, M.; Yang, X.; Kanega, R.; Kawanami, H.; Xu, Q.; Himeda, Y. Development of effective catalysts for hydrogen storage technology using formic acid. Adv. Energy Mater. 2019, 9, 1801275. [Google Scholar] [CrossRef]
  9. Yang, X.; Xu, Q. Encapsulating metal nanocatalysts within porous organic hosts. Trends Chem. 2020, 2, 214–226. [Google Scholar] [CrossRef]
  10. Wang, C.; Astruc, D. Recent developments of nanocatalyzed liquid-phase hydrogen generation. Chem. Soc. Rev. 2021, 50, 3437. [Google Scholar] [CrossRef]
  11. Zheng, J.; Zhou, H.; Wang, C.G.; Ye, E.; Xu, J.W.; Loh, X.J.; Li, Z. Current research progress and perspectives on liquid hydrogen rich molecules in sustainable hydrogen storage. Energy Storage Mater. 2021, 35, 695–722. [Google Scholar] [CrossRef]
  12. Xu, R.; Lu, W.; Toan, S.; Zhou, Z.; Russell, C.K.; Sun, Z.; Sun, Z. Thermocatalytic formic acid dehydrogenation: Recent advances and emerging trends. J. Mater. Chem. A 2021, 9, 24241–24260. [Google Scholar] [CrossRef]
  13. Bulushev, D.A.; Ross, J.R.H. Heterogeneous catalysts for hydrogenation of CO2 and bicarbonates to formic acid and formates. Catal. Rev. 2018, 60, 566–593. [Google Scholar] [CrossRef]
  14. Wang, T.; Xue, W.; Wang, Y.J. Review on heterogeneous catalysts for hydrogen generation via liquid-phase formic acid decomposition. J. Chem. Eng. Chin. Univ. 2019, 33, 1–9. [Google Scholar]
  15. Yadav, M.; Akita, T.; Tsumori, N.; Xu, Q. Strong metal–molecular support interaction (SMMSI): Amine-functionalized gold nanoparticles encapsulated in silica nanospheres highly active for catalytic decomposition of formic acid. J. Mater. Chem. 2012, 22, 12582–12586. [Google Scholar] [CrossRef]
  16. Bi, Q.Y.; Du, X.L.; Liu, Y.M.; Cao, Y.; He, H.Y.; Fan, K.N. Efficient subnanometric gold-catalyzed hydrogen generation via formic acid decomposition under ambient conditions. J. Am. Chem. Soc. 2012, 134, 8926–8933. [Google Scholar] [CrossRef]
  17. Liu, Q.Q.; Yang, X.F.; Huang, Y.Q.; Xu, S.T.; Su, X.L.; Pan, X.; Xu, J.M.; Wang, A.Q.; Liang, C.H.; Wang, X.K.; et al. A Schiff base modified gold catalyst for green and efficient H2 production from formic acid. Energy Environ. Sci. 2015, 8, 3204. [Google Scholar] [CrossRef]
  18. Zacharska, M.; Chuvilin, A.L.; Kriventsov, V.V.; Beloshapkin, S.; Estrada, M.; Simakov, A.; Bulushev, D.A. Support effect for nanosized Au catalysts in hydrogen production from formic acid decomposition. Catal. Sci. Technol. 2016, 6, 6853–6860. [Google Scholar] [CrossRef]
  19. Bulushev, D.A.; Zacharska, M.; Guo, Y.; Beloshapkin, S.; Simakov, A. CO-free hydrogen production from decomposition of formic acid over Au/Al2O3 catalysts doped with potassium ions. Catal. Commun. 2017, 92, 86–89. [Google Scholar] [CrossRef]
  20. Bulushev, D.A.; Chekhova, G.N.; Sobolev, V.I.; Chuvilin, A.L.; Fedoseeva, Y.V.; Gerasko, O.A.; Okotrub, A.V.; Bulusheva, L.G. Cucurbit [6] uril as a co-catalyst for hydrogen production from formic acid. Mater. Today Energy 2022, 26, 100998. [Google Scholar] [CrossRef]
  21. Wang, N.; Sun, Q.; Bai, R.; Li, X.; Guo, G.; Yu, J. In situ confinement of ultrasmall Pd clusters within nanosized silicalite-1 zeolite for highly efficient catalysis of hydrogen generation. J. Am. Chem. Soc. 2016, 138, 7484–7487. [Google Scholar] [CrossRef]
  22. Sun, Q.; Wang, N.; Bing, Q.; Si, R.; Liu, J.; Bai, R.; Zhang, P.; Jia, M.; Yu, J. Subnanometric hybrid Pd-M(OH)2, M = Ni, Co, clusters in zeolites as highly efficient nanocatalysts for hydrogen generation. Chem 2017, 3, 477–493. [Google Scholar] [CrossRef] [Green Version]
  23. Zhu, Q.L.; Tsumori, N.; Xu, Q. Sodium hydroxide-assisted growth of uniform Pd nanoparticles on nanoporous carbon MSC-30 for efficient and complete dehydrogenation of formic acid under ambient conditions. Chem. Sci. 2014, 5, 195–199. [Google Scholar] [CrossRef]
  24. Zhu, Q.L.; Tsumori, N.; Xu, Q. Immobilizing extremely catalytically active palladium nanoparticles to carbon nanospheres: A weakly-capping growth approach. J. Am. Chem. Soc. 2015, 137, 11743–11748. [Google Scholar] [CrossRef] [PubMed]
  25. Song, F.Z.; Zhu, Q.L.; Tsumori, N.; Xu, Q. Diamine-alkalized reduced graphene oxide: Immobilization of sub-2 nm palladium nanoparticles and optimization of catalytic activity for dehydrogenation of formic acid. ACS Catal. 2015, 5, 5141–5144. [Google Scholar] [CrossRef]
  26. Li, Z.; Yang, X.; Tsumori, N.; Liu, Z.; Himeda, Y.; Autrey, T.; Xu, Q. Tandem nitrogen functionalization of porous carbon: Toward immobilizing highly active palladium nanoclusters for dehydrogenation of formic acid. ACS Catal. 2017, 7, 2720–2724. [Google Scholar] [CrossRef]
  27. Zhong, S.; Yang, X.; Chen, L.; Tsumori, N.; Taguchi, N.; Xu, Q. Interfacing with Fe–N–C sites boosts the formic acid dehydrogenation of palladium nanoparticles. ACS Appl. Mater. Interfaces 2021, 13, 46749–46755. [Google Scholar] [CrossRef]
  28. Bi, Q.Y.; Lin, J.D.; Liu, Y.M.; Du, X.L.; Wang, J.Q.; He, H.Y.; Cao, Y. An aqueous rechargeable formate-based hydrogen battery driven by heterogenous Pd catalysis. Angew. Chem. Int. Ed. 2014, 53, 13583–13587. [Google Scholar] [CrossRef]
  29. Bi, Q.Y.; Lin, J.D.; Liu, Y.M.; He, H.Y.; Huang, F.Q.; Cao, Y. Dehydrogenation of formic acid at room temperature: Boosting palladium nanoparticle efficiency by coupling with pyridinic-nitrogen-doped carbon. Angew. Chem. Int. Ed. 2016, 55, 11849–11853. [Google Scholar] [CrossRef]
  30. Bulushev, D.A.; Zacharska, M.; Shlyakhova, E.V.; Chuvilin, A.L.; Guo, Y.; Beloshapkin, S.; Okotrub, A.V.; Bulusheva, L.G. Single isolated Pd2+ cations supported on N-doped carbon as active sites for hydrogen production from formic acid decomposition. ACS Catal. 2016, 6, 681–691. [Google Scholar] [CrossRef]
  31. Bulushev, D.A.; Zacharska, M.; Lisitsyn, A.S.; Podyacheva, O.Y.; Hage, F.S.; Ramasse, Q.M.; Bangert, U.; Bulusheva, L.G. Single atoms of Pt-group metals stabilized by N-doped carbon nanofibers for efficient hydrogen production from formic acid. ACS Catal. 2016, 6, 3442–3451. [Google Scholar] [CrossRef]
  32. Jia, L.; Bulushev, D.A.; Ross, J.R.H. Formic acid decomposition over palladium-based catalysts doped by potassium carbonate. Catal. Today 2016, 259, 453–459. [Google Scholar] [CrossRef] [Green Version]
  33. Bulushev, D.A.; Bulusheva, L.G. Catalysts with single metal atoms for the hydrogen production from formic acid. Catal. Rev. 2021; in press. [Google Scholar] [CrossRef]
  34. Bulushev, D.A. Progress in catalytic hydrogen production from formic acid over supported metal complexes. Energies 2021, 14, 1334. [Google Scholar] [CrossRef]
  35. Tang, C.; Surkus, E.; Chen, F.; Pohl, M.M.; Agostini, G.; Schneider, M.; Junge, H.; Beller, M. A stable nanocobalt catalyst with highly dispersed CoNx active sites for the selective dehydrogenation of formic acid. Angew. Chem. Int. Ed. 2017, 56, 16616–16620. [Google Scholar] [CrossRef]
  36. Li, X.; Surkus, A.E.; Rabeah, J.; Anwar, M.; Dastigir, S.; Junge, H.; Brückner, A.; Beller, M. Cobalt single-atom catalysts with high stability for selective dehydrogenation of formic acid. Angew. Chem. Int. Ed. 2020, 59, 15849–15854. [Google Scholar] [CrossRef]
  37. Zhou, X.C.; Huang, Y.J.; Xing, W.; Liu, C.P.; Liao, J.H.; Lu, T.H. High-quality hydrogen from the catalyzed decomposition of formic acid by Pd–Au/C and Pd–Ag/C. Chem. Commun. 2008, 30, 3540–3542. [Google Scholar] [CrossRef]
  38. Huang, Y.J.; Zhou, X.C.; Yin, M.; Liu, C.P.; Xing, W. Novel PdAu@Au/C core−shell catalyst: Superior activity and selectivity in formic acid decomposition for hydrogen generation. Chem. Mater. 2010, 22, 5122–5128. [Google Scholar] [CrossRef]
  39. Tedsree, K.; Li, T.; Jones, S.; Chan, C.W.A.; Yu, K.M.K.; Bagot, P.A.J.; Marquis, E.A.; Smith, G.D.; Tsang, S.C.E. Hydrogen production from formic acid decomposition at room temperature using a Ag–Pd core–shell nanocatalyst. Nat. Nanotechnol. 2011, 6, 302–307. [Google Scholar] [CrossRef]
  40. Wu, S.; Yang, F.; Wang, H.; Chen, R.; Sun, P.C.; Chen, T.H. Mg2+-assisted low temperature reduction of alloyed AuPd/C: An efficient catalyst for hydrogen generation from formic acid at room temperature. Chem. Commun. 2015, 51, 10887–10890. [Google Scholar] [CrossRef]
  41. Gu, X.J.; Lu, Z.H.; Jiang, H.L.; Akita, T.; Xu, Q. Synergistic catalysis of metal–organic framework-immobilized Au–Pd nanoparticles in dehydrogenation of formic acid for chemical hydrogen storage. J. Am. Chem. Soc. 2011, 133, 11822–11825. [Google Scholar] [CrossRef]
  42. Yan, J.M.; Wang, Z.L.; Gu, L.; Li, S.J.; Wang, H.L.; Zheng, W.T.; Jiang, Q. AuPd–MnOx/MOF–graphene: An efficient catalyst for hydrogen production from formic acid at room temperature. Adv. Energy Mater. 2015, 5, 1500107. [Google Scholar] [CrossRef]
  43. Li, S.J.; Zhou, Y.T.; Kang, X.; Liu, D.X.; Gu, L.; Zhang, Q.H.; Yan, J.M.; Jiang, Q. A simple and effective principle for a rational design of heterogeneous catalysts for dehydrogenation of formic acid. Adv. Mater. 2019, 31, 1806781. [Google Scholar] [CrossRef] [PubMed]
  44. Song, F.Z.; Zhu, Q.L.; Yang, X.; Zhan, W.W.; Pachfule, P.; Tsumori, N.; Xu, Q. Metal-organic framework templated porous carbon-metal oxide/reduced graphene oxide as superior support of bimetallic nanoparticles for efficient hydrogen generation from formic acid. Adv. Energy Mater. 2018, 8, 1701416. [Google Scholar] [CrossRef]
  45. Chen, Y.; Zhu, Q.L.; Tsumori, N.; Xu, Q. Immobilizing highly catalytically active noble metal nanoparticles on reduced graphene oxide: A non-noble metal sacrificial approach. J. Am. Chem. Soc. 2015, 137, 106–109. [Google Scholar] [CrossRef]
  46. Yang, X.; Pachfule, P.; Chen, Y.; Tsumori, N.; Xu, Q. Highly efficient hydrogen generation from formic acid using a reduced graphene oxide-supported AuPd nanoparticle catalyst. Chem. Commun. 2016, 52, 4171–4174. [Google Scholar] [CrossRef]
  47. Yang, X.; Li, Z.; Kitta, M.; Tsumori, N.; Guo, W.; Zhang, Z.; Zhang, J.; Zou, R.; Xu, Q. Solid-solution alloy nanoclusters of the immiscible gold-rhodium system achieved by a solid ligand-assisted approach for highly efficient catalysis. Nano Res. 2020, 13, 105–111. [Google Scholar] [CrossRef]
  48. Wang, Z.L.; Yan, J.M.; Ping, Y.; Wang, H.L.; Zheng, W.T.; Jiang, Q. An efficient CoAuPd/C catalyst for hydrogen generation from formic acid at room temperature. Angew. Chem. Int. Ed. 2013, 52, 4406–4409. [Google Scholar] [CrossRef]
  49. Wang, Z.L.; Ping, Y.; Yan, J.M.; Wang, H.L.; Jiang, Q. Hydrogen generation from formic acid decomposition at room temperature using a NiAuPd alloy nanocatalyst. Int. J. Hydrog. Energy 2014, 39, 4850–4856. [Google Scholar] [CrossRef]
  50. Yurderi, M.; Bulut, A.; Zahmakiran, M.; Kaya, M. Carbon supported trimetallic PdNiAg nanoparticles as highly active, selective and reusable catalyst in the formic acid decomposition. Appl. Catal. B Environ. 2014, 160–161, 514–524. [Google Scholar] [CrossRef]
  51. Yang, L.; Luo, W.; Cheng, G. Monodisperse CoAgPd nanoparticles assembled on graphene for efficient hydrogen generation from formic acid at room temperature. Int. J. Hydrog. Energy 2016, 41, 439–446. [Google Scholar] [CrossRef]
  52. He, L.; Liang, B.; Huang, Y.; Zhang, T. Design strategies of highly selective nickel catalysts for H2 production via hydrous hydrazine decomposition: A review. Natl. Sci. Rev. 2018, 5, 356–364. [Google Scholar] [CrossRef] [Green Version]
  53. Zhang, A.; Yao, Q.; Lu, Z.H. Recent progress on catalysts for hydrogen evolution from decomposition of hydrous hydrazine. Acta Chim. Sin. 2021, 79, 885–902. [Google Scholar] [CrossRef]
  54. Singh, S.K.; Zhang, X.B.; Xu, Q. Room-temperature hydrogen generation from hydrous hydrazine for chemical hydrogen storage. J. Am. Chem. Soc. 2009, 131, 9894–9895. [Google Scholar] [CrossRef] [PubMed]
  55. He, L.; Huang, Y.; Wang, A.; Wang, X.; Chen, X.; Delgado, J.J.; Zhang, T. A noble-metal-free catalyst derived from Ni-Al hydrotalcite for hydrogen generation from N2H4·H2O decomposition. Angew. Chem. Int. Ed. 2012, 51, 6191–6194. [Google Scholar] [CrossRef] [PubMed]
  56. He, L.; Liang, B.; Li, L.; Yang, X.; Huang, Y.; Wang, A.; Wang, X.; Zhang, T. Cerium-oxide-modified nickel as a non-noble metal catalyst for selective decomposition of hydrous hydrazine to hydrogen. ACS Catal. 2015, 5, 1623–1628. [Google Scholar] [CrossRef]
  57. Kang, W.; Varma, A. Hydrogen generation from hydrous hydrazine over Ni/CeO2 catalysts prepared by solution combustion synthesis. Appl. Catal. B Environ. 2018, 220, 409–416. [Google Scholar] [CrossRef]
  58. Wang, H.; Wu, L.; Jia, A.; Li, X.; Shi, Z.; Duan, M.; Wang, Y. Ni nanoparticles encapsulated in the channel of titanate nanotubes: Efficient noble-metal-free catalysts for selective hydrogen generation from hydrous hydrazine. Chem. Eng. J. 2018, 332, 637–646. [Google Scholar] [CrossRef]
  59. Huang, M.; Yao, Q.; Feng, G.; Zou, H.; Lu, Z.H. Nickel-ceria nanowires embedded in microporous silica: Controllable synthesis, formation mechanism, and catalytic applications. Inorg. Chem. 2020, 59, 5781–5790. [Google Scholar] [CrossRef]
  60. Singh, S.K.; Xu, Q. Complete conversion of hydrous hydrazine to hydrogen at room temperature for chemical hydrogen storage. J. Am. Chem. Soc. 2009, 131, 18032–18033. [Google Scholar] [CrossRef]
  61. Singh, A.K.; Yadav, M.; Aranishi, K.; Xu, Q. Temperature-induced selectivity enhancement in hydrogen generation from Rh-Ni nanoparticle-catalyzed decomposition of hydrous hydrazine. Int. J. Hydrog. Energy 2012, 37, 18915–18919. [Google Scholar] [CrossRef]
  62. Singh, S.K.; Xu, Q. Bimetallic Ni-Pt nanocatalysts for selective decomposition of hydrazine in aqueous solution to hydrogen at room temperature for chemical hydrogen storage. Inorg. Chem. 2010, 49, 6148–6152. [Google Scholar] [CrossRef]
  63. Singh, A.K.; Xu, Q. Highly-dispersed surfactant-free bimetallic Ni–Pt nanoparticles as high-performance catalyst for hydrogen generation from hydrous hydrazine. Int. J. Hydrog. Energy 2014, 39, 9128–9134. [Google Scholar] [CrossRef]
  64. Kumar, A.; Yang, X.; Xu, Q. Ultrafine bimetallic Pt–Ni nanoparticles immobilized on 3-dimensional N-doped graphene networks: A highly efficient catalyst for dehydrogenation of hydrous hydrazine. J. Mater. Chem. A 2019, 7, 112–115. [Google Scholar] [CrossRef]
  65. Song, F.Z.; Yang, X.; Xu, Q. Ultrafine bimetallic Pt–Ni nanoparticles achieved by metal-organic framework templated zirconia/porous carbon/reduced graphene oxide: Remarkable catalytic activity in dehydrogenation of hydrous hydrazine. Small Methods 2020, 4, 1900707. [Google Scholar] [CrossRef]
  66. Zou, H.; Zhang, S.; Hong, X.; Yao, Q.; Luo, Y.; Lu, Z.H. Immobilization of Ni-Pt nanoparticles on MIL-101/rGO composite for hydrogen evolution from hydrous hydrazine and hydrazine borane. J. Alloys Compd. 2020, 835, 155426. [Google Scholar] [CrossRef]
  67. Guo, F.; Zou, H.; Yao, Q.; Huang, B.; Lu, Z.H. Monodispersed bimetallic nanoparticles anchored on TiO2-decorated titanium carbide MXene for efficient hydrogen production from hydrazine in aqueous solution. Renew. Energy 2020, 155, 1293–1301. [Google Scholar] [CrossRef]
  68. Qiu, Y.P.; Shi, Q.; Zhou, L.L.; Chen, M.H.; Chen, C.; Tang, P.P.; Walker, G.S.; Wang, P. NiPt nanoparticles anchored onto hierarchical nanoporous N-doped carbon as an efficient catalyst for hydrogen generation from hydrazine monohydrate. ACS Appl. Mater. Interfaces 2020, 12, 18617–18624. [Google Scholar] [CrossRef]
  69. Du, X.; Liu, C.; Du, C.; Cai, P.; Cheng, G.; Luo, W. Nitrogen-doped graphene hydrogel-supported NiPt-CeOx nanocomposites and their superior catalysis for hydrogen generation from hydrazine at room temperature. Nano Res. 2017, 10, 2856–2865. [Google Scholar] [CrossRef]
  70. Singh, S.K.; Iizuka, Y.; Xu, Q. Nickel-palladium nanoparticle catalyzed hydrogen generation from hydrous hydrazine for chemical hydrogen storage. Int. J. Hydrog. Energy 2011, 36, 11794–11801. [Google Scholar] [CrossRef]
  71. Hong, X.; Yao, Q.; Huang, M.; Du, H.; Lu, Z.H. Bimetallic NiIr nanoparticles supported on lanthanum oxy-carbonate as highly efficient catalysts for hydrogen evolution from hydrazine borane and hydrazine. Inorg. Chem. Front. 2019, 6, 2271–2278. [Google Scholar] [CrossRef]
  72. Song-Il, O.; Yan, J.M.; Wang, H.L.; Wang, Z.L.; Jiang, Q. High catalytic kinetic performance of amorphous CoPt NPs induced on CeOx for H2 generation from hydrous hydrazine. Int. J. Hydrog. Energy 2014, 39, 3755–3761. [Google Scholar] [CrossRef]
  73. Wang, K.; Yao, Q.; Qing, S.; Lu, Z.H. La(OH)3 nanosheet-supported CoPt nanoparticles: A highly efficient and magnetically recyclable catalyst for hydrogen production from hydrazine in aqueous solution. J. Mater. Chem. A 2019, 7, 9903–9911. [Google Scholar] [CrossRef]
  74. Wang, J.; Li, W.; Wen, Y.; Gu, L.; Zhang, Y. Rh-Ni-B Nanoparticles as Highly Efficient Catalysts for Hydrogen Generation from Hydrous Hydrazine. Adv. Energy Mater. 2015, 5, 1401879. [Google Scholar] [CrossRef]
  75. Du, X.; Tan, S.; Cai, P.; Luo, W.; Cheng, G. A RhNiP/rGO hybrid for efficient catalytic hydrogen generation from an alkaline solution of hydrazine. J. Mater. Chem. A 2016, 4, 14572–14576. [Google Scholar] [CrossRef]
  76. Liu, T.; Yu, J.; Bie, H.; Hao, Z. Highly efficient hydrogen generation from hydrous hydrazine using a reduced graphene oxide-supported NiPtP nanoparticle catalyst. J. Alloys Compd. 2017, 690, 783–790. [Google Scholar] [CrossRef]
  77. Chen, J.; Zou, H.; Yao, Q.; Luo, M.; Li, X.; Lu, Z.H. Cr2O3-modified NiFe nanoparticles as a noble-metal-free catalyst for complete dehydrogenation of hydrazine in aqueous solution. Appl. Surf. Sci. 2020, 501, 144247. [Google Scholar] [CrossRef]
  78. Kang, W.; Guo, H.; Varma, A. Noble-metal-free NiCu/CeO2 catalysts for H2 generation from hydrous hydrazine. Appl. Catal. B Environ. 2019, 249, 54–62. [Google Scholar] [CrossRef]
  79. Wu, D.; Wen, M.; Lin, X.; Wu, Q.; Gu, C.; Chen, H. A NiCo/NiO–CoOx ultrathin layered catalyst with strong basic sites for high-performance H2 generation from hydrous hydrazine. J. Mater. Chem. A 2016, 4, 6595–6602. [Google Scholar] [CrossRef]
  80. Chandra, M.; Xu, Q. A high-performance hydrogen generation system: Transition metal-catalyzed dissociation and hydrolysis of ammonia–borane. J. Power Sources 2006, 156, 190–194. [Google Scholar] [CrossRef]
  81. Aijaz, A.; Karkamkar, A.; Choi, Y.J.; Tsumori, N.; Roennebro, E.; Autrey, T.; Shioyama, H.; Xu, Q. Immobilizing highly catalytically active Pt nanoparticles inside the pores of metal–organic framework: A double solvents approach. J. Am. Chem. Soc. 2012, 134, 13926–13929. [Google Scholar]
  82. Du, C.; Ao, Q.; Cao, N.; Yang, L.; Luo, W.; Cheng, G. Facile synthesis of monodisperse ruthenium nanoparticles supported on graphene for hydrogen generation from hydrolysis of ammonia borane. Int. J. Hydrog. Energy 2015, 40, 6180–6187. [Google Scholar] [CrossRef]
  83. Chen, W.; Ji, J.; Duan, X.; Qian, G.; Li, P.; Zhou, X.; Chen, D.; Yuan, W. Unique reactivity in Pt/CNT catalyzed hydrolytic dehydrogenation of ammonia borane. Chem. Commun. 2014, 50, 2142–2144. [Google Scholar] [CrossRef]
  84. Yao, Q.; Lu, Z.H.; Jia, Y.; Chen, X.; Liu, X. In situ facile synthesis of Rh nanoparticles supported on carbon nanotubes as highly active catalysts for H2 generation from NH3BH3 hydrolysis. Int. J. Hydrog. Energy 2015, 40, 2207–2215. [Google Scholar] [CrossRef]
  85. Chen, Y.; Yang, X.; Kitta, M.; Xu, Q. Monodispersed Pt nanoparticles on reduced graphene oxide by a non-noble metal sacrificial approach for hydrolytic dehydrogenation of ammonia borane. Nano Res. 2017, 10, 3811–3816. [Google Scholar] [CrossRef]
  86. Pachfule, P.; Yang, X.; Zhu, Q.L.; Tsumori, N.; Uchida, T.; Xu, Q. From Ru nanoparticle-encapsulated metal–organic frameworks to highly catalytically active Cu/Ru nanoparticle-embedded porous carbon. J. Mater. Chem. A 2017, 5, 4835–4841. [Google Scholar] [CrossRef]
  87. Zhang, S.Y.; Kochovski, Z.; Lee, H.C.; Lu, Y.; Zhang, H.; Zhang, J.; Sun, J.K.; Yuan, J. Ionic organic cage-encapsulating phase-transferable metal clusters. Chem. Sci. 2019, 10, 1450–1456. [Google Scholar] [CrossRef]
  88. Hu, J.; Chen, Z.; Li, M.; Zhou, X.; Lu, H. Amine-capped Co nanoparticles for highly efficient dehydrogenation of ammonia borane. ACS Appl. Mater. Interfaces 2014, 6, 13191–13200. [Google Scholar] [CrossRef]
  89. Mahyari, M.; Shaabani, A. Nickel nanoparticles immobilized on three-dimensional nitrogen-doped graphene as a superb catalyst for the generation of hydrogen from the hydrolysis of ammonia borane. J. Mater. Chem. A 2014, 2, 16652–16659. [Google Scholar] [CrossRef]
  90. Tong, D.G.; Zeng, X.L.; Chu, W.; Wang, D.; Wu, P. Magnetically recyclable hollow Co–B nanospindles as catalysts for hydrogen generation from ammonia borane. J. Mater. Sci. 2010, 45, 2862–2867. [Google Scholar] [CrossRef]
  91. Rakap, M. Hydrogen generation from the hydrolytic dehydrogenation of ammonia borane using electrolessly deposited cobalt–phosphorus as reusable and cost-effective catalyst. J. Power Sources 2014, 265, 50–56. [Google Scholar] [CrossRef]
  92. Peng, C.Y.; Kang, L.; Cao, S.; Chen, Y.; Lin, Z.S.; Fu, W.F. Nanostructured Ni2P as a robust catalyst for the hydrolytic dehydrogenation of ammonia-borane. Angew. Chem. Int. Ed. 2015, 54, 15725–15729. [Google Scholar] [CrossRef]
  93. Gu, J.; Yang, Y.; Tang, Y. Effects of heat-treatment on the adhesive strength of diamond films coated on the cemented carbide substrates. J. Integr. Technol. 2017, 6, 80–88. [Google Scholar]
  94. Rakap, M. The highest catalytic activity in the hydrolysis of ammonia borane by poly(N-vinyl-2-pyrrolidone)-protected palladium–rhodium nanoparticles for hydrogen generation. Appl. Catal. B Environ. 2015, 163, 129–134. [Google Scholar] [CrossRef]
  95. Park, J.W.; Lai, S.W.; Cho, S.O. Catalytic hydrogen generation from hydrolysis of ammonia borane using octahedral Au@Pt nanoparticles. Int. J. Hydrog. Energy 2015, 40, 16316–16322. [Google Scholar] [CrossRef]
  96. Wang, N.; Sun, Q.; Zhang, T.; Mayoral, A.; Li, L.; Zhou, X.; Xu, J.; Zhang, P.; Yu, J. Impregnating subnanometer metallic nanocatalysts into self-pillared zeolite nanosheets. J. Am. Chem. Soc. 2021, 143, 6905–6914. [Google Scholar] [CrossRef]
  97. Huang, X.; Liu, Y.; Wen, H.; Shen, R.; Mehdi, S.; Wu, X.; Liang, E.; Guo, X.; Li, B. Ensemble-boosting effect of Ru-Cu alloy on catalytic activity towards hydrogen evolution in ammonia borane hydrolysis. Appl. Catal. B Environ. 2021, 287, 119960. [Google Scholar] [CrossRef]
  98. Guo, J.; Li, X.; Duan, H.; Zhang, H.; Jia, Q.; Zhang, S. Graphene supported Pt–Ni bimetallic nanoparticles for efficient hydrogen generation from KBH4/NH3BH3 hydrolysis. Int. J. Hydrog. Energy 2022, 47, 11601–11610. [Google Scholar] [CrossRef]
  99. Guo, L.T.; Cai, Y.Y.; Ge, J.M.; Zhang, Y.N.; Gong, L.H.; Li, X.H.; Wang, K.X.; Ren, Q.Z.; Su, J.; Chen, J.S. Multifunctional Au–Co@CN nanocatalyst for highly efficient hydrolysis of ammonia borane. ACS Catal. 2015, 5, 388–392. [Google Scholar] [CrossRef]
  100. Zhu, Q.L.; Li, J.; Xu, Q. Immobilizing metal nanoparticles to metal–organic frameworks with size and location control for optimizing catalytic performance. J. Am. Chem. Soc. 2013, 135, 10210–10213. [Google Scholar] [CrossRef]
  101. Li, J.; Zhu, Q.L.; Xu, Q. Highly active AuCo alloy nanoparticles encapsulated in the pores of metal–organic frameworks for hydrolytic dehydrogenation of ammonia borane. Chem. Commun. 2014, 50, 5899–5901. [Google Scholar] [CrossRef]
  102. Fu, F.; Wang, C.; Wang, Q.; Martinez-Villacorta, A.M.; Escobar, A.; Chong, H.; Wang, X.; Moya, S.; Salmon, L.; Fouquet, E.; et al. Highly selective and sharp Volcano-type synergistic Ni2Pt@ZIF-8-catalyzed hydrogen evolution from ammonia borane hydrolysis. J. Am. Chem. Soc. 2018, 140, 10034–10042. [Google Scholar] [CrossRef] [PubMed]
  103. Ge, Y.; Qin, X.; Li, A.; Deng, Y.; Lin, L.; Zhang, M.; Yu, Q.; Li, S.; Peng, M.; Xu, Y.; et al. Maximizing the synergistic effect of CoNi catalyst on α-MoC for robust hydrogen production. J. Am. Chem. Soc. 2021, 143, 628–633. [Google Scholar] [CrossRef]
  104. Dai, H.B.; Kang, X.D.; Wang, P. Ruthenium nanoparticles immobilized in montmorillonite used as catalyst for methanolysis of ammonia borane. Int. J. Hydrog. Energy 2010, 35, 10317–10323. [Google Scholar] [CrossRef]
  105. Sun, J.K.; Zhan, W.W.; Akita, T.; Xu, Q. Toward homogenization of heterogeneous metal nanoparticle catalysts with enhanced catalytic performance: Soluble porous organic cage as a stabilizer and homogenizer. J. Am. Chem. Soc. 2015, 137, 7063–7066. [Google Scholar] [CrossRef]
  106. Yang, X.; Sun, J.K.; Kitta, M.; Pang, H.; Xu, Q. Encapsulating highly catalytically active metal nanoclusters inside porous organic cages. Nat. Catal. 2018, 1, 214–220. [Google Scholar] [CrossRef]
  107. Yang, X.; Chen, L.; Liu, H.; Kurihara, T.; Horike, S.; Xu, Q. Encapsulating ultrastable metal nanoparticles within reticular Schiff base nanospaces for enhanced catalytic performance. Cell Rep. Phys. Sci. 2021, 2, 100289. [Google Scholar] [CrossRef]
  108. Cao, W.; Zhou, J.; Kochovski, Z.; Miao, H.; Gao, Z.; Sun, J.; Yuan, J. Ionic organic cage-encapsulated metal clusters for switchable catalysis. Cell Rep. Phys. Sci. 2021, 2, 100546. [Google Scholar] [CrossRef]
  109. Zhu, L.; Zhang, S.; Yang, X.; Zhuang, Q.; Sun, J.K. Toward the controlled synthesis of highly dispersed metal clusters enabled by downsizing crystalline porous organic cage supports. Small Methods 2022, 6, 2200591. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of Au NPs loaded on the Schiff base modified SiO2 [17]. (Reprinted with permission from Ref. [17], 2015, Royal Society of Chemistry).
Figure 1. Schematic illustration of Au NPs loaded on the Schiff base modified SiO2 [17]. (Reprinted with permission from Ref. [17], 2015, Royal Society of Chemistry).
Energies 15 06360 g001
Figure 2. Cs-corrected STEM images of 0.8Pd0.2Ni(OH)2@S-1 and Pd@S-1 catalysts viewed along b axis orientation with low (a,d) and high magnification (b,e), and viewed along other orientation (c,f) [22]. (Reprinted with permission from Ref. [22], 2017, Elsevier).
Figure 2. Cs-corrected STEM images of 0.8Pd0.2Ni(OH)2@S-1 and Pd@S-1 catalysts viewed along b axis orientation with low (a,d) and high magnification (b,e), and viewed along other orientation (c,f) [22]. (Reprinted with permission from Ref. [22], 2017, Elsevier).
Energies 15 06360 g002
Figure 3. Schematic illustration of the process for the synthesis of the AuPd-MnOx/ZIF-8-rGO composite [42]. (Reprinted with permission from Ref. [42], 2015, John Wiley and Sons).
Figure 3. Schematic illustration of the process for the synthesis of the AuPd-MnOx/ZIF-8-rGO composite [42]. (Reprinted with permission from Ref. [42], 2015, John Wiley and Sons).
Energies 15 06360 g003
Figure 4. Volume of the generated gas vs. time for the dehydrogenation of formic acid over the Au/uC-548, Au0.5Rh0.5/uC-548 and Rh/uC-548 catalysts (n(Au+Rh)/nFA = 0.02, nSF/nFA = 2.5, 333 K), Inset: corresponding TOF values of the catalysts [47]. (Reprinted with permission from Ref. [47], 2020, Springer Nature).
Figure 4. Volume of the generated gas vs. time for the dehydrogenation of formic acid over the Au/uC-548, Au0.5Rh0.5/uC-548 and Rh/uC-548 catalysts (n(Au+Rh)/nFA = 0.02, nSF/nFA = 2.5, 333 K), Inset: corresponding TOF values of the catalysts [47]. (Reprinted with permission from Ref. [47], 2020, Springer Nature).
Energies 15 06360 g004
Figure 5. Selectivity and TOF values for the catalytic hydrogen production from hydrazine hydrate over PtxNi1−x@ZrO2/C/rGO (0 ≤ x ≤ 1.0) at 323 K [65]. (Reprinted with permission from Ref. [65], 2020, John Wiley and Sons).
Figure 5. Selectivity and TOF values for the catalytic hydrogen production from hydrazine hydrate over PtxNi1−x@ZrO2/C/rGO (0 ≤ x ≤ 1.0) at 323 K [65]. (Reprinted with permission from Ref. [65], 2020, John Wiley and Sons).
Energies 15 06360 g005
Figure 6. TEM (a), HRTEM (b) and high-angle annular dark-field STEM (HAADF, (c,d)) images of the Ni0.7Pt0.3P/rGO catalyst, as well as the corresponding element mapping of Pt (e), Ni (f) and P (g) [76]. (Reprinted with permission from Ref. [76], 2017, Elsevier).
Figure 6. TEM (a), HRTEM (b) and high-angle annular dark-field STEM (HAADF, (c,d)) images of the Ni0.7Pt0.3P/rGO catalyst, as well as the corresponding element mapping of Pt (e), Ni (f) and P (g) [76]. (Reprinted with permission from Ref. [76], 2017, Elsevier).
Energies 15 06360 g006
Figure 7. Volume of the generated H2 gas versus time for the hydrolytic dehydrogenation of ammonia borane over Pt/rGO at 298 K (nPt/nAB = 0.01) [85]. (Reprinted with permission from Ref. [85], 2017, Springer Nature).
Figure 7. Volume of the generated H2 gas versus time for the hydrolytic dehydrogenation of ammonia borane over Pt/rGO at 298 K (nPt/nAB = 0.01) [85]. (Reprinted with permission from Ref. [85], 2017, Springer Nature).
Energies 15 06360 g007
Figure 8. Bimetallic clusters inside self-pillared zeolite nanosheets for hydrolytic dehydrogenation of ammonia borane [96]. (Reprinted with permission from Ref. [96], 2021, ACS).
Figure 8. Bimetallic clusters inside self-pillared zeolite nanosheets for hydrolytic dehydrogenation of ammonia borane [96]. (Reprinted with permission from Ref. [96], 2021, ACS).
Energies 15 06360 g008
Figure 9. Time course plots for the methanolysis of ammonia borane over the Pd@RCC3 and Pd/RCC3 catalysts at 303 K [106]. (Reprinted with permission from Ref. [106], 2018, Springer Nature).
Figure 9. Time course plots for the methanolysis of ammonia borane over the Pd@RCC3 and Pd/RCC3 catalysts at 303 K [106]. (Reprinted with permission from Ref. [106], 2018, Springer Nature).
Energies 15 06360 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, X.; Bulushev, D.A.; Yang, J.; Zhang, Q. New Liquid Chemical Hydrogen Storage Technology. Energies 2022, 15, 6360. https://0-doi-org.brum.beds.ac.uk/10.3390/en15176360

AMA Style

Yang X, Bulushev DA, Yang J, Zhang Q. New Liquid Chemical Hydrogen Storage Technology. Energies. 2022; 15(17):6360. https://0-doi-org.brum.beds.ac.uk/10.3390/en15176360

Chicago/Turabian Style

Yang, Xinchun, Dmitri A. Bulushev, Jun Yang, and Quan Zhang. 2022. "New Liquid Chemical Hydrogen Storage Technology" Energies 15, no. 17: 6360. https://0-doi-org.brum.beds.ac.uk/10.3390/en15176360

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