Next Article in Journal
Room Temperature Surfactant-Free Synthesis of Gold Nanoparticles in Alkaline Ethylene Glycol
Next Article in Special Issue
Probing Low-Temperature OCM Performance over a Dual-Domain Catalyst Bed
Previous Article in Journal
Combining Two Types of Boron in One Molecule (To the 60th Anniversary of the First Synthesis of Carborane)
Previous Article in Special Issue
Production of Alkyl Levulinates from Carbohydrate-Derived Chemical Intermediates Using Phosphotungstic Acid Supported on Humin-Derived Activated Carbon (PTA/HAC) as a Recyclable Heterogeneous Acid Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hydrogen Evolution upon Ammonia Borane Solvolysis: Comparison between the Hydrolysis and Methanolysis Reactions

1
ISM, UMR CNRS N°5255, University of Bordeaux, 33405 Talence CEDEX, France
2
Institute of Circular Economy, Faculty of Materials and Manufacturing, Beijing University of Technology, Beijing 100124, China
*
Authors to whom correspondence should be addressed.
Chemistry 2023, 5(2), 886-899; https://doi.org/10.3390/chemistry5020060
Submission received: 5 February 2023 / Revised: 20 March 2023 / Accepted: 7 April 2023 / Published: 13 April 2023

Abstract

:
Hydrogen (H2) production is a key challenge for green carbon-free sustainable energy. Among the H2 evolution methods from H-rich materials, ammonia borane (AB) solvolysis stands as a privileged source under ambient and sub-ambient conditions given its stability, non-toxicity, and solubility in protic solvents, provided suitable and optimized nanocatalysts are used. In this paper dedicated to Prof. Avelino Corma, we comparatively review AB hydrolysis and alcoholysis (mostly methanolysis) in terms of nanocatalyst performances and discuss the advantages and inconveniences of these two AB solvolysis methods including AB regeneration.

Graphical Abstract

1. Introduction

Hydrogen (H2) generation is one of the main focuses of current research, particularly because it is a green energy source toward the replacement of harmful fossil-based fuel [1,2]. H2 energy has appeared to be essential, because it is carbon-free, without pollution, and it involves an excellent energy storage density [1]. Besides water (hydrogen evolution reaction, HER [3]), liquid-phase hydrogen generation systems include a variety of metal borohydrides, ammonia and its borane derivatives (ammonia borane, amine boranes, hydrazine borane, etc.), formic acid, and several hydrocarbons [1]. Among all these possibilities, since the seminal work by Chandra and Xu [4], ammonia borane (AB) has appeared to be by far the most studied H2 source. During the first decade of the century, AB thermolysis was the subject of various studies [5,6], but the high temperatures required (only 15% H2 is obtained at 200 °C [7]), long induction times, and ill-defined mixture of BxNy derivatives obtained made nanocatalyzed AB solvolysis a privileged route [2,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23]. Therefore, we are briefly reviewing here the two major AB solvolysis routes, i.e., hydrolysis and methanolysis, with the goal of comparing them, because, despite similarities, there are important differences in terms of environmental concern, rates, efficiencies, facilities, and side reactions. Publications generally focus on either of them without comparison with the alternative solvolysis method. In conclusion, researchers should be able to consider the parameters and state of the art toward selecting their choice. Finally, in the conclusion, prospects are formulated toward improving the hydrogen generation process upon AB solvolysis.

2. Ammonia Borane (AB)

Ammonia borane, currently abbreviated as AB, is a white solid, due to the intermolecular H bonding (NH—HB distance about 2 Å [24]). Its molecular weight Mw is 30.8 g·mol−1 and its density at 25 °C is ρ = 0.78 g·mL−1. It is stable in water at neutral pH, whereas NaBH4, another H2 source, slowly reacts with water. AB is stabilized, formally by a dative bond from N to B with a net electron transfer from N to B (Figure 1, left) of approximately 0.5 electrons and with the formation of a compound that is 65% ionic; the ionic structure is represented on the right of Figure 1 [25]. Thus, the N-H bonds are acidic, whereas the B-H bonds are hydridic.
AB is soluble in ether, 1,2-dimethoxyethane (DME), water, alcohols, and many ionic liquids; slightly soluble in benzene, dichloromethane, and chloroform; insoluble in alkanes.
AB was first synthesized by Shore and Parry and reported in 1955 (Equation (1)) [26].
LiBH 4 + NH 4 Cl NH 3 BH 3 + LiCl + H 2
Since then, many variations of this commonly utilized metathesis-type synthesis between metal borohydrides MBH4 and ammonium salts NH4X have appeared, summarized six decades later in Shore’s review [27].
The synthesis of AB was improved by the use of methyl borate, B(OMe)3 [28]. Methyl borate had indeed long been available from boric acid, B(OH)3, and methanol, a reaction reported in 1953 by Schlesinger et al. (Equation (2)) [29].
H 3 BO 3 + 4 CH 3 OH B ( OMe ) 3 + MeOH + 3 H 2 O
In particular, Ramachandran’s group developed this especially practical synthesis from B(OMe)3 using LiAlH4 in THF added to a mixture of B(OMe)3 and NH4Cl in THF at 0 °C over 1 h, followed by stirring 3 h at 0 °C (90% isolated yield after extraction with ether) (Equation (3)) [30].
B ( OMe ) 3 + NH 4 Cl THF ,   0   ° C ,   3   h LiAlH 4 BH 3 NH 3 + Al ( OMe ) 3 + LiCl + H 2
In 2015, Ramachandran’s group even reported an astute extension of the metathesis reaction of Equation (1) to the syntheses of a variety of amine borane derivatives in high yields and purity by reactions of alkylammonium sulfate derivatives with NaBH4 [29].

3. H2 Generation upon AB Hydrolysis

Besides properties such as a mild reducing agent and other properties that are reviewed elsewhere [2,5,6], AB is an excellent source of H2, which is due to its low molecular weight, high hydrogen content (19.6 wt%), high solubility in water (35 g of AB for 100 g of water), stability, and lack of toxicity.

3.1. Use of an Acidic Medium

Although AB is stable in neutral or slightly basic aqueous solution, H2 is formed in acidic solutions owing to the hydricity [31] of the BH groups of AB. Chandra and Xu reported that 3 mol H2 was released from AB in aqueous solution in the presence of solid acids such as Dorex, Amberlyst, and Nafion-H, in a few minutes under ambient conditions, according to Equation (4) [31]. CO2 also reacts as an acid, upon dissolution into water, releasing H2 together with the formation of boric acid [32].
NH 3 BH 3 + 2 H 2 O H + NH 4 + , BO 2 + 3 H 2

3.2. Use of a Transition-Metal Catalyst

In neutral aqueous medium, the presence of a transition-metal-based catalyst is necessary to achieve AB hydrolysis with the formation of 3 mol H2 and hydrated ammonium borate (Equation (5)).
NH 3 BH 3 + 4 H 2 O metal-based catalyst NH 4 + , B ( OH ) 4 + 3 H 2
A large variety of transition-metal-based catalysts including homogeneous and heterogeneous catalysts, transition-metal complexes, oxides, transition-metal nanoparticles (NPs), and heterobinuclear catalysts have been reported and reviewed by Xu’s group and others [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23]. Ramachandran’s group reported that upon catalysis with RuCl3, NH3 was also produced along with H2 evolution during AB hydrolysis, and the amount of NH3 produced increased upon increasing the AB concentration, which would be a problem for the realization of fuel cells. The setup for H2 production from AB and measurement is shown in Figure 2 [33].
These authors also found that the borate produced in the reaction of Equation (5) was ammonium tetraborate {[(NH)4]2[B4O5(OH)4]·1.41H2O} with the nature of the borate anions depending on the pH, and polyborate anions being formed at higher pH [32]. In early studies, transition-metal complexes were shown to be efficient for AB hydrolysis with H2 evolution. Mechanistic studies privileged both B-H and N-H cleavage in the rate-determining step (RDS) [33,34]. The intermediacy of organometallic or inorganic radicals during the catalyst process might play a significant role [35]. Other studies with molybdenum complexes showed that the mechanism depended on the oxidation state of the catalyst [36]. Finally, with Co-Ni and Ni NPs, O-H activation by the catalyst in the RDS was proposed [37,38].
Here, the focus is on nanocatalysts that have attracted most of the attention during the last two decades, for which the NP support has played an essential role [39,40,41,42,43,44,45,46,47,48,49]. In particular, porous materials including metal–organic frameworks (MOFs) have been a privileged type of NP support, because the NP catalyst can be formed and encapsulated inside MOFs for efficient nanocatalysis [50,51,52,53,54,55,56,57]. Therefore, hydrophilic late transition-metal salts are introduced in water, for instance, into the cavity of the zeolitic imidazolate framework (ZIF) [58], for which ZIF-8 [59,60,61] has been very successful in our laboratory using reduction with NaBH4 [62] of the transition-metal cation to NP inside the ZIF-8 cavity [18,63,64,65,66,67] (Scheme 1).
Under these conditions, Ni was shown to be the most catalytically effective first-row transition metal with an Ni NP size of 2.7 nm inside ZIF-8. In the presence of 0.3 N that appears to favor the catalysis, the turnover frequency (TOF) at 25 °C was of 85.7 molH2·molcat.−1·min−1 [63]. The reaction is first-order in catalyst concentration and zero-order in AB concentration, an indication that the role of AB in the transition state is not significant. This is confirmed by a kinetic isotope effect (KIE) with D2O (KIE = kH2O/kD2O) of 2.49, suggesting that water O-H bond cleavage is the rate-determining step (RDS). The favorable influence of OH on the reaction rate was taken into account by OH coordination onto the Ni NP surface, increasing the electron density on Ni, which is favorable for water O-H oxidative addition onto surface Ni atoms. Meanwhile, on–off control of the H2 evolution was demonstrated upon successive addition of OH that activated H2 evolution and H+ that stopped the reaction (Figure 3) [63].

3.3. Use of a Transition-Metal-Alloyed Catalyst

Heterobimetallic catalysts often result in positive synergistic effects [68,69], and Hou et al. synthesized graphene-oxide-supported bimetallic Ni-Co catalysts reaching a TOF of 154 molH2·molcat.−1·min−1 at ambient temperature [68]. A remarkable result was obtained in our group with Pt-Ni NPs in ZIF-8 (Ni2Pt@ZIF-8) for which the TOF value reached 669.3 molH2·molcat.−1·min−1. The synthesis of NiPt NP@ZIF-8 is shown in Scheme 2 [64].
This high TOF value was 87 times higher than the TOF of Pt@ZIF-8 under similar conditions, as a result of a positive volcano-type type synergy between Ni and Pt. The suggested mechanism was related to that with Ni@ZIF-8, and the primary KIE with D2O was high (4.67), showing that water O-H cleavage was the RDS [63]. A considerable volcano-type positive synergy in catalytic efficiency was also obtained with Co-Pt NPs embedded in “click” dendrimers. In these “click” dendrimers that contain crucial triazole ligands, the nanocatalysts were encapsulated inside the dendrimer by coordination onto these ligand groups bridging the dendrimer core to the dendrimer periphery (Figure 4). In the presence of 0.3 M NaOH, the TOF value was 476.2 molH2·molcatal−1·min−1 (952.4 molH2·molPt−1·min−1). In both the ZIF-8 (endoreceptor) and click dendrimers (exoreceptors), the N atoms of the support played an important role in the confinement of the catalyst together with the substrates, and similar mechanistic trends were observed with both supports (Scheme 3) [70].
Plasmon excitation by visible light is useful in nanocatalysis [71,72,73,74,75,76,77,78,79,80], including for H2 evolution from AB [76,77,78]. When the bimetallic nanocatalyst included Au as one of the two metals, excitation of the Au plasmon by illumination with visible light boosted H2 evolution, with the most efficient light effects being observed upon combining Au with M = Co, Ni, Ru, Rh, or Pt. The evolution of 3 mol H2 was observed in down to 1.3 min at 25 °C, 3 times faster than in the dark. Co was found to be the best first-raw transition metal alloyed with Au [78]. This boosting effect was taken into account by the plasmon-induced “hot” electron transfer, from Au to the other late transition metal M in the alloy, enriching the electron density on M, which enhances activation of the substrates [78,79]. Indeed, in the reaction mechanism, it is the transition metal atom M at the surface that is responsible for water O-H bond cleavage by oxidative addition of this bond in the rate-determining step (RDS). The increase in electron density at this metal center, such as the one provided from the plasmonic metal by “hot” electron transfer, largely favors this oxidative addition (vide infra). Primary KIEs observed using D2O under dark conditions and found to be even larger under visible-light irradiation confirmed that O-H bond cleavage was largely involved in the RDS. In the initial hydride transfer from AB to the nanocatalyst, it appears that water is not significantly involved (contrary to NaBH4 as the H2 source, because the negatively charged BH4 anion is more significantly hydrogen-bonded to water than in neutral AB). On the other hand, after such a hydride transfer (or oxidative addition of the B-H bond immediately followed by B to NP electron transfer), the nanocatalyst becomes negatively charged, which increases the ability of water activation by oxidative addition onto the active late transition metal M (but not Au alone, which is a very poor nanocatalyst of the reaction). The increase in electron density at the active metal site is still considerably enhanced by the plasmonic hot electron transfer. Thus, the anionic [NP-H] intermediate is likely to form a hydrogen bond with water, [NP-H]–H–OH, removing some electron density from the water O–H bond, which further facilitates oxidative addition of this bond onto the activated transition-metal surface. This first reaction phase ends by reductive elimination of the two hydride ligands forming H2 and the OH and NH3BH2 ligands providing NH3BH2OH that, in turn, undergoes H2 formation according to a similar process (Scheme 1) [78].

4. Catalyzed H2 Generation upon AB Methanolysis

It is striking that AB methanolysis (Equation (6)) has been so much less studied than AB hydrolysis.
NH 3 BH 3 + 4 CH 3 OH catalyst NH 4 B ( OCH 3 ) 4 + 3 H 2
The seminal AB methanolysis publication is that of Ramachandran and Gagare who reported in 2007 both optimized ammonia borane preparation and its methanolysis [33]. These authors utilized, for AB methanolysis, late transition-metal catalysts, both of the first-row metals and noble metals (CoCl2, Raney Ni, Pd/C, RuCl3 and RhCl3), with RuCl3 being found as the best catalyst. With this catalyst, they obtained a TOF value of 150 molH2·molcatal−1·min−1, which already was a very good performance compared to all those obtained afterward (vide infra) and up to now. Jagirdar’s group reported Co, Ni, and Cu-based nanocatalysts (including oxides and bimetallic Ni-Cu alloys), avoiding the use of noble metals [81,82]. Among all the AB methanolysis reports up to now (most of the reported TOF values are below 100 molH2·molcatal−1·min−1), a remarkable performance in terms of TOF value for this reaction was obtained by Sun’s group with 366.4 molH2·molcatal−1·min−1 using a Ag-Pd nanocatalyst, Ag30Pd70, synthesized by the coreduction of AgNO3 and PdCl2 in the presence of oleylamine, supported on carbon [83]. Another very efficient catalyst reported by Lu’s group contained ultrafine Rh NPs (1.4–2.6 nm) confined in a covalent organic framework (COF) synthesized from piperazine and cyanuric chloride, and it produced a TOF of 505 molH2·molcatal−1·min−1 at 298 K. Given the temperature effect decreasing the TOF value upon decreasing the temperature, such high TOF values ensure applicability for low-temperature devices [84].
The same group recently reported an even better catalyst, a porphyrin-derived N-doped porous carbon-confined Ru NP for AB methanolysis with a very high TOF value of 727 molH2·molcatal−1·min−1 at 25 °C [85]. Luo et al. reported L-proline (PRO)-capped Rh NPs as a catalyst with an AB methanolysis TOF of 1035 molH2·molcatal−1·min−1 and 0.3 M NaOH that is known to boost late transition-metal nanocatalysts of AB solvolysis (vide supra). The presence of a suitable O- and N-rich PRO template also contributes to the positive synergy. At 0 °C, this nanocatalyst was still effective in ethanol that is less toxic than methanol, with these trends being useful toward low-temperature portable applications [86]. The best AB methanolysis performance so far was obtained by Zahmakiran’s group using nanocatalyst MIL-101-decorated Pd NPs with TOF = 1080 molH2·molcatal−1·min−1 at room temperature [87].
The literature on AB methanolysis has recently been reviewed by Xu’s group [8], Li and Wang’s group [88], and Özkar [89], and reports on nanocatalysts have appeared based on first-row late transition-metals [8,90], and also on noble metals [90,91,92].
H2 evolution upon AB methanolysis presents several advantages over the method using AB hydrolysis toward applications [8,88,91]: (i) AB is highly soluble in methanol under ambient conditions (23 wt%) [33,92]; (ii) AB is very stable in methanol in the absence of a catalyst; (iii) H2 evolution upon methanolysis of AB is pure without ammonia (contrary to AB hydrolysis), which is an impurity poisoning fuel cells; (iv) AB methanolysis produces H2 at low temperature if the nanocatalyst is efficient enough (methanol freezes at (−97.8 °C), which could be useful toward portable devices at cold temperatures); (v) NH4B(OCH3)4 that is formed upon AB methanolysis readily reacts with LiAlH4 and NH4Cl at room temperature to regenerate AB, whereas the regeneration of NH4B(OH)4 formed upon AB hydrolysis is more complicated [33,89].
AB methanolysis is, in general, as for AB hydrolysis, first-order in catalyst concentration and zero-order in AB concentration, and the mechanism has been suggested to involve O-H cleavage in the RDS [81,82,83,84,85,86,87,88,89,90,91,92,93]. For instance, along this line, Lu’s group, in their work noted above on porphyrin-derived N-doped porous carbon-confined Ru NPs, indicated that AB methanolysis of both NH3BD3 and 15NH3BH3 in CH3OH showed secondary KIEs of 1.29 and 1.05, respectively, and methanolysis of NH3BH3 in CD3OD involved a primary KIE of 5.78. These results clearly showed the key role of methanol in the RDS step of AB methanolysis [85].
AB methanolysis was also recently catalyzed by Au-Pd@ ZIF-8, generated by reducing equimolar amounts of HAuCl4 and Na2PdCl4 by NaBH4 in methanol in the presence of ZIF-8, with the reaction being 3.7 times faster under visible-light illumination than in the dark. KIE values measured in CD3OD were 2.2 and 3.4, respectively, in agreement with O–H bond cleavage as the RDS. Synergy effects between Pd and Au were taken into account in DFT calculations by the d-orbital density of states (DOS) of different NPs, with the d-band center of Au (111) and Pd (111) being −3.2 and −2.4 eV, respectively, whereas, upon alloying, the d-band center of AuPd (111) was −1.5 eV, at a higher energy than those of both Au (111) and Pd (111). Thus, the d orbital of the metal in AuPd is closer to the energy level of the antibonding orbital of the adsorbate species than with the related Au and Pd nanocatalysts, signifying a higher AB adsorption energy on AuPd. Upon illumination, a hot plasmon Au electron is injected into the Pd-AB adsorbate, increasing the electron density on Pd toward the methanol O–H oxidative addition onto the Pd surface center of the alloy, as confirmed by the DFT calculations (Scheme 4) [93].
Alternatively, Yamashita’s group examined photocatalytic H2 generation over the Au/TiO2 nanocatalyst from AB decomposition. This reaction was shown to be initiated by the generation of plasmon-induced charge pairs. It is not known, however, how the RDS would change upon AB methanolysis compared to hydrolysis in this process [94].

5. Comparison between AB Hydrolysis and Methanolysis

AB solvolysis is a superb and the most studied efficient method of H2 generation, because AB is safe, stable in water and methanol under ambient conditions, and can now be obtained in >99% purity. Both AB hydrolysis and methanolysis produce 3 mol H2 per mol AB, but the H2 produced by hydrolysis contains all the more NH3 as the AB concentration is higher, which is damaging, because NH3 is poisoning fuel cells, whereas the H2 produced by AB methanolysis is pure from NH3 impurities. This is a very important advantage of methanolysis over hydrolysis, although nanocatalyzed AB hydrolysis proceeds faster than methanolysis. Another important parameter is the temperature, because hydrolysis is limited to reactions above 0 °C, the freezing point of water, whereas AB methanolysis can be operated at low to very low temperatures that are not limited by the extremely low freezing point of methanol as long as the catalyst is reactive enough at these low temperatures. Another essential parameter is the regeneration of AB from the AB solvolysis side product, NH4B(OH)4 in the case of hydrolysis, and NH4B(OCH3)4 in the case of AB methanolysis. This AB regeneration is much more complicated in the case of hydrolysis than in that of methanolysis, because NH4B(OCH3)4 is easily converted to AB upon reaction with LiAlH4 and NH4Cl, as reported by Ramachandran [33], whereas NH4B(OH)4 must first be converted to NH4B(OCH3)4 by reaction with methanol (Equation (7), Scheme 5). However, the recycling of LiAl(OMe)4 has not yet been conducted (Scheme 5) [33,88]
NH 4 B ( OCH 3 ) 4 + NH 4 Cl + LiAlH 4 NH 3 BH 3 + Al ( OCH 3 ) 3 + CH 3 OH + H 2 + LiCl + NH 3 81 %   yield   > 98 %   purity  
The energy-consuming step in AB regeneration is the reduction of the strong bonds B-O in B(OCH3)3 or NH4B(OH)4, produced by AB solvolysis, which must proceed using a hydride such as LiAlH4 Equation (7) [92].
A comparison of the mechanisms of nanocatalyzed AB dehydrogenative hydrolysis and methanolysis suggests that methanolysis is more boosted (3.7 times) by visible-light illumination compared to the reaction in the dark than hydrolysis (1.4 time). The KIE increased from 2.2 in the dark to 3.4 under illumination for AB methanolysis, but only from 4 to 4.3 for AB hydrolysis. This shows that visible light has much more influence on AB methanolysis than on AB hydrolysis. Upon examining the hydrogen bonding between the anionic hydridic species [Pd–H], formed by hydride transfer from AB to the NP, and the acidic solvent O–H bond, this bond is stronger upon AB hydrolysis than upon AB methanolysis, as water is more acidic than methanol. Consequently, solvent O–H bond weakening is more significant with water than with methanol, and solvent O–H bond oxidative addition on the nanocatalyst surface is easier with water than with methanol. This explains that the additional activation by visible light is necessary upon AB methanolysis compared to AB hydrolysis [66,78,93].

6. Conclusions

Among the H-rich H2 sources [1], AB stands as one of the best suppliers upon solvolysis, either hydrolysis or alcoholysis. This issue has been raised by the seminal report from Chandra and Xu in 2006 on AB hydrolysis [4]. Since then, considerable efforts have been devoted to improvements toward safe and efficient H2 production upon AB solvolysis. In particular, interest in these H2 generation methods has been boosted by the in-depth seminal work by the Ramachandran group, since 2007, who disclosed that the AB methanolysis reaction improved AB synthesis and proposed its regeneration from AB solvolysis reactions [28,30,33].
Given the AB stability in water and alcohols, catalysis, particularly nanocatalysis by late transition-metal NPs and alloys stabilized by optimized supports, plays the central role in H2 production from AB solvolysis. As in most catalytic reactions, noble metals (Ru, Rh, Pd, and Pt), despite their toxicity, high price, and scarcity, are more efficient than non-noble metals, i.e., those of the first-row transition metals (Co, Ni, Cu), but synergies among these metals and with the support can increase catalytic efficiencies, particularly doping with other elements such as B or P. Studies involving single-atom catalysts are also awaited [95,96]. Nanoporous supports are especially efficient and useful [97], because they favor the formation of ultrasmall and thus very efficient nanocatalysts and protects them from aggregation.
In comparing AB hydrolysis and methanolysis, it quickly appeared that AB hydrolysis encountered a major problem in that the AB hydrolysis product was all the more contaminated by NH3 as the AB concentration was higher, which was damaging due to the NH3 poisoning of fuel cells [98,99]. On the other hand, the H2 produced is pure upon methanolysis.
The other drawback of AB hydrolysis is the regeneration after hydrolysis that is longer and more complex than regeneration after AB methanolysis. Finally, the last drawback is that the temperature window is much more limited for hydrolysis, due to water freezing at 0 °C, compared to methanolysis.
Favorable points for hydrolysis in comparison with methanolysis are that water is cheaper and less toxic than methanol and that the hydrolysis reactions are faster than the methanolysis reactions. Yet, TOFs of the AB methanolysis reaction with sophisticated nanocatalysts have very recently approached or overtaken 1000 molH2·molcat.−1·min−1, allowing perspectives for the application to low-temperature devices [84,85,86]. Ethanol is less toxic than methanol, but nanocatalyzed reactions are slower in ethanol than in methanol.
In summary, the most critical issue for AB solvolysis is the high cost of AB and of its regeneration after solvolysis, even more so for hydrolysis than for methanolysis. More research is required for improvements in catalysts and supports, particularly with emphasis on non-noble metal nanocatalysts. Up-to-date investigation tools such as in situ XRD, electron microscopies, X-ray absorption, theoretical calculations, and Machine Learning [100] should help the improvements, scaling, and transfer to industry for device applications toward an “H2 economy”.

Author Contributions

Conceptualization, N.K., C.W. and D.A.; methodology, N.K., C.W. and D.A.; software, N.K., C.W. and D.A.; validation, C.W. and D.A.; formal analysis, N.K., C.W. and D.A.; investigation, N.K., C.W. and D.A.; resources, C.W. and D.A.; data curation, N.K., C.W. and D.A.; writing—original draft preparation, NK.; writing—review and editing, C.W. and D.A.; visualization, N.K., C.W. and D.A.; supervision, C.W. and D.A.; project administration, D.A.; funding acquisition, D.A. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support from the Chinese Scholarship Council (CSC grant to N.K.), the National Natural Science Foundation of China (No. 21805166), the start-up funding by Beijing University of Technology (C.W.), the University of Bordeaux, and the Centre National de la Recherche Scientifique (CNRS) is gratefully acknowledged.

Data Availability Statement

Not applicable.

Acknowledgments

Outstanding contributions and fruitful discussions concerning original works with colleagues and former students cited in references, in particular, Fangyu Fu and Wang Qi in the University of Bordeaux, are gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. He, T.; Pachfule, P.; Wu, H.; Xu, Q.; Chen, P. Hydrogen Carriers. Nat. Rev. Mater. 2016, 1, 16059. [Google Scholar] [CrossRef]
  2. Lebrouhi, B.E.; Djoupo, J.J.; Lamrani, B.; Benabdelaziz, K.; Kousksou, T. Global Hydrogen Development—A Technical and Geopolitical Overview. Int. J. Hydrogen Energy 2022, 47, 7016–7048. [Google Scholar] [CrossRef]
  3. Zhu, J.; Hu, L.; Zhao, P.; Lee, L.Y.S.; Wong, K.Y. Recent Advances in Electrocatalyic Hydrogen Evolution. Chem. Rev. 2020, 120, 851–918. [Google Scholar] [CrossRef] [PubMed]
  4. Chandra, M.; Xu, Q.J. A High-Performance Hydrogen Generation System: Transition Metal-Catalyzed Dissociation and Hdrolysis of Ammonia-Borane. Power Sources 2006, 156, 190–194. [Google Scholar] [CrossRef]
  5. Hamilton, C.W.; Baker, R.T.; Staubitz, A.; Manners, I. B–N compounds for chemical hydrogenstorage. Chem. Soc. Rev. 2009, 38, 279–293. [Google Scholar] [CrossRef]
  6. Staubitz, A.; Robertson, A.P.M.; Manners, I. B-N Compounds for Chemical Hydrogen Storage. Chem. Rev. 2010, 110, 4079–4124. [Google Scholar] [CrossRef]
  7. Wang, P. Molecular Systems for Ligh-Driven Hydrogen Production. Dalton Trans. 2012, 41, 4296e302. [Google Scholar]
  8. Zhan, W.-W.; Zhu, Q.-L.; Xu, Q. Dehydrogenation of Ammonia Borane by Metal Nanoparticle Catalysts. ACS Catal. 2016, 6, 6892–6905. [Google Scholar] [CrossRef]
  9. Li, Z.; Xu, Q. Metal-Nanocatalyzed Hydrogen Generation from Formic Acid. Acc. Chem. Res. 2017, 50, 1449–1458. [Google Scholar] [CrossRef]
  10. Yu, X.; Tang, Z.; Sun, D.; Ouyang, L.; Zhu, M. Recent Advances and Remaining Challenges of Nanosctuctured Materials for Hydrogen Storage Applications. Prog. Mater. Sci. 2017, 88, 1–48. [Google Scholar] [CrossRef]
  11. Demici, U.B. Ammonia borane, a material with exceptional properties for ammonia borane storage. Int. J. Hydrogen Energy 2017, 42, 9978–10013. [Google Scholar] [CrossRef]
  12. Zhu, B.; Zou, R.; Xu, Q. Metal-Organic Framework Based Catalysts for Hydrogen Evolution. Adv. Energy Mater. 2018, 8, 1801193. [Google Scholar] [CrossRef]
  13. Akbayrak, S.; Özkar, S. Ammonia borane as hydrogen storage materials. Int. J. Hydrogen Energy 2018, 43, 18592–18606. [Google Scholar] [CrossRef]
  14. Sun, Q.M.; Wang, N.; Xu, Q.; Yu, J. Nanopore-Supported Metal Nanocatalysts for Efficient Hydrogen Generation from Liquid Phase Chemical Hydrogen Storage. Adv. Mater. 2020, 32, 2001818. [Google Scholar] [CrossRef] [PubMed]
  15. Huang, Z.G.; Wang, S.N.; Dewhurst, R.D.; Ignat’ev, N.V.; Finze, M.; Braunschweig, H. Boron: Its Role in Energy-Related Processes and Applications. Angew. Chem. Int. Ed. 2020, 59, 8800–8816. [Google Scholar] [CrossRef] [PubMed]
  16. Axet, M.R.; Philippot, K. Catalysis with Colloidal Ruthenium Nanoparticles. Chem. Rev. 2020, 120, 1085–1145. [Google Scholar] [CrossRef] [PubMed]
  17. Yao, Q.L.; Ding, Y.; Lu, Z.-H. Nobel-metal free catalysts for hydrogen generation from boron- and nitrogen-based hydrides. Inorg. Chem. Front. 2020, 7, 3837–3874. [Google Scholar] [CrossRef]
  18. Wang, C.; Wang, Q.; Fu, F.; Astruc, D. Hydrogen Generation from Nanocatalyzed Hydrolysis of Hydrogen-Rich Boron Derivatives: Recent Developments. Acc. Chem. Res. 2020, 53, 2483–2493. [Google Scholar] [CrossRef]
  19. Wang, C.; Astruc, D. Recent Developments of Liquid-Phase Hydrogen Generation. Chem. Soc. Rev. 2021, 50, 3437–3484. [Google Scholar] [CrossRef]
  20. Yue, M.; Lambert, H.; Pahon, E.; Roche, R.; Jemei, S.; Hissel, D. Hydrogen Energy Systems: A Critical Review of Technologies, Applications, Trends and Challenges. Renew. Sustain. Energy Rev. 2021, 146, 111180. [Google Scholar] [CrossRef]
  21. Lau, S.; Gasperini, D.; Webster, R.L. Ammonia Boranes as Transfer Hydrogenation and Hydrogenation Reagents. Angew. Chem. Int. Ed. 2021, 6, 14272–14294. [Google Scholar] [CrossRef] [PubMed]
  22. Mboyi, C.D.; Poinsot, D.; Roger, J.; Fajerwerg, K.; Kahn, M.L.; Hierso, J.-C. The Hydrogen-Storage Challenge: Nanoparticles for Metal-Catalyzed Ammonia Borane Dehydrgenation. Small 2021, 17, 2102759. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, X.; Zhang, X.; Li, D.-S.; Zhang, S.; Zhang, Q.J. Recent Advances in the “On-Off” Approaches for the on-Demand Liquid-Phase Hydrogen Evolution. Mater. Chem. A 2021, 9, 18164–18174. [Google Scholar] [CrossRef]
  24. Klooster, W.T.; Koetzle, T.F.; Siegbahn, P.E.M.; Richardson, T.B.; Crabtree, R.H. Study of the NH...H-B Dihydrogen Bond Including the Crystal Structure of BH3NH3 by Neutron Diffraction. J. Am. Chem. Soc. 1999, 121, 6337–6343. [Google Scholar] [CrossRef]
  25. Plumley, J.A.; Evanseck, J.D. Covalent and Ionic Nature of the Dative Bond and Account of Accurate Ammonia-Borane Binding Enthalpies. J. Phys. Chem. A 2007, 111, 13472–13483. [Google Scholar] [CrossRef]
  26. Shore, S.G.; Parry, R.W. The Cystalline Compound Ammonia Borane, H3NBH3. J. Am. Chem. Soc. 1955, 77, 6084–6085. [Google Scholar] [CrossRef]
  27. Li, H.; Yang, Q.; Shore, S.G. Ammonia Borane, Past as Prolog. J. Organomet. Chem. 2014, 751, 60–66. [Google Scholar] [CrossRef]
  28. Ramachandran, P.V.; Raju, B.C.; Gagare, P.D. One-Pot Synthesis of Ammonia-Borane and Trialkylamine-Borane from Trimethyl Borate. Org. Lett. 2012, 14, 6119–6121. [Google Scholar] [CrossRef]
  29. Schlesinger, H.I.; Brown, H.C.; Mayfield, D.L.; Gilbreath, J.R. The Preparation of Other Borohydrides by Metathetical Reactions Utilizing the Metal Alkali Borohydrides. J. Am. Chem. Soc. 1953, 75, 213–215. [Google Scholar] [CrossRef]
  30. Ramachandran, P.V.; Kulkarni, A.S. Open Flask Synthesis of Amine Boranes via Tandem Amine-Ammonium Salt Equilibration-Metathesis. Inorg. Chem. 2015, 54, 5618–5620. [Google Scholar] [CrossRef]
  31. Wiedner, E.S.; Chambers, M.B.; Pitman, C.L.; Bullock, R.M.; Miller, A.J.M.; Appel, A.M. Thermodynamic Hydricity of Transition Metal Hydrides. Chem. Rev. 2016, 116, 8655–8692. [Google Scholar] [CrossRef] [PubMed]
  32. Chandra, M.; Xu, Q.J. Dissociation and hydrolysis of ammonia-borane with solid acids and carbon dioxide: An efficient hydrogen generation system. Power Sources 2006, 159, 855–860. [Google Scholar] [CrossRef]
  33. Ramachandran, P.V.; Gagare, P.D. Preparation of Ammonia Borane in High Yield and Purity, Methanolysis, and Regeneration. Inorg. Chem. 2007, 46, 7810–7817. [Google Scholar] [CrossRef] [PubMed]
  34. Bhattacharya, P.; Krause, J.A.; Guan, H. Mechanistic Studies of Ammonia Borane Dehydrogenation Catalyzed by Iron Pincer Complexes. J. Am. Chem. Soc. 2014, 136, 11153–11161. [Google Scholar] [CrossRef] [PubMed]
  35. Astruc, D. Transition-metal radicals: Chameleon structure and catalytic function. Acc. Chem. Res. 1991, 24, 36–42. [Google Scholar] [CrossRef]
  36. Buss, J.A.; Edouard, G.A.; Cheng, C.; Shi, J.; Agapie, T. Molybdenum Catalyzed Ammonia Borane Dehydrogenation. J. Am. Chem. Soc. 2014, 136, 11272–11275. [Google Scholar] [CrossRef] [Green Version]
  37. Chen, W.; Li, D.; Wang, Z.; Qian, G.; Sui, Z.; Duan, X.; Zhou, X.; Yeboah, I.; Chen, D. Reaction mechanism and kinetics for hydrolytic dehydrogenation of ammonia borane on a Pt/CNT catalyst. AIChE J. 2017, 63, 60–65. [Google Scholar] [CrossRef]
  38. Li, Z.; He, T.; Liu, L.; Chen, W.; Zhang, M.; Wu, G.; Chen, P. Covalent triazine framework supported non-noble metal nanoparticles with superior activity for catalytic hydrolysis of ammonia borane: From mechanistic study to catalyst design. Chem. Sci. 2017, 8, 781–788. [Google Scholar] [CrossRef] [Green Version]
  39. Corma, A.; Garcia, H. Supported gold nanoparticles as catalysts for organic reactions. Chem. Soc. Rev. 2008, 37, 2096–2126. [Google Scholar] [CrossRef]
  40. Statakis, M.; Garcia, H. Catalysis by Supported Gold Nanoparticles: Beyond Aerobic Oxidative Processes. Chem. Rev. 2012, 112, 4469–4506. [Google Scholar] [CrossRef]
  41. Polshettiwar, V.; Cha, D.; Zhang, X.X.; Basset, J.M. High-Surface-Area Silica Nanospheres (KCC-1) with a Fibrous Morphology. Angew. Chem. Int. Ed. 2010, 49, 9652–9656. [Google Scholar] [CrossRef] [PubMed]
  42. Figri, A.; Bouhrara, M.; Nekoueishahraki, B.; Basset, J.M.; Polshettiwar, V. Nanocatalysts for Suzuki cross-coupling reactions. Chem. Soc. Rev. 2011, 40, 5181–5203. [Google Scholar]
  43. Narayanan, R.; El-Sayed, B.J. Catalysis with Transition Metal Nanoparticles in Colloidal Solution:  Nanoparticle Shape Dependence and Stability. Phys. Chem. B 2005, 109, 12663–12676. [Google Scholar] [CrossRef] [PubMed]
  44. Shifrina, Z.B.; Matveeva, V.G.; Bronstein, M. Role of Polymer Structures in Catalysis by Transition Metal and Metal Oxide Nanoparticle Composites. Chem. Rev. 2020, 120, 1350–1396. [Google Scholar] [CrossRef]
  45. Li, G.; Jin, R.C. Atomically Precise Gold Nanoclusters as New Model Catalysts. Acc. Chem. Res. 2013, 46, 1749–1758. [Google Scholar] [CrossRef]
  46. Niu, Z.Q.; Li, Y.D. Removal and Utilization of Capping Agents in Nanocatalysis. Chem. Mater. 2014, 26, 72–83. [Google Scholar] [CrossRef]
  47. Gawande, M.B.; Shelke, S.N.; Sboril, R.; Varma, R.S. Microwave-Assisted Chemistry: Synthetic Applications for Rapid Assembly of Nanomaterials and Organics. Acc. Chem. Res. 2014, 47, 1338–1348. [Google Scholar] [CrossRef]
  48. Astruc, D.; Heuze, K.; Gatard, S.; Méry, D.; Nlate, S.; Plault, L. Metallodendritic Catalysis for Redox and Carbon—Carbon Bond Formation Reactions: A Step towards Green Chemistry. Adv. Syn. Catal. 2005, 347, 329–338. [Google Scholar] [CrossRef]
  49. Astruc, D. Introduction: Nanoparticles in Catalysis. Chem. Rev. 2020, 120, 461–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Corma, A.; Garcia, H.; Xamena, F.X.L. Engineering Metal Organic Frameworks for Heterogeneous Catalysis. Chem. Rev. 2010, 110, 4606–4655. [Google Scholar] [CrossRef]
  51. Viciano-Chumillas, M.; Mon, M.; Ferrando-Soria, J.; Corma, A.; Leyva-Perez, A.; Armentano, D.; Pardo, E. Metal–Organic Frameworks as Chemical Nanoreactors: Synthesis and Stabilization of Catalytically Active Metal Species in Confined Spaces. Acc. Chem. Res. 2020, 53, 520–531. [Google Scholar] [CrossRef] [PubMed]
  52. Choi, K.M.; Na, K.; Somorjai, G.A.; Yaghi, O.M. Chemical Environment Control and Enhanced Catalytic Performance of Platinum Nanoparticles Embedded in Nanocrystalline Metal–Organic Frameworks. J. Am. Chem. Soc. 2015, 137, 7810–7816. [Google Scholar] [CrossRef] [PubMed]
  53. Dhakshinamoorthy, A.; Li, Z.H.; Garcia, H. Catalysis and photocatalysis by metal organic frameworks. Chem. Soc. Rev. 2018, 47, 8134–8172. [Google Scholar] [CrossRef] [PubMed]
  54. Dhakshinamoorthy, A.; Asiri, M.; Garcia, H. Metal–Organic Frameworks as Multifunctional Solid Catalysts. Trends Chem. 2020, 2, 454–466. [Google Scholar] [CrossRef]
  55. Furukawa, H.; Cordova, K.E.; Yaghi, O.M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 974–986. [Google Scholar] [CrossRef] [Green Version]
  56. Lee, J.; Farha, O.K.; Roberts, J.; Scheidt, K.A.; Nguyen, S.T.; Hupp, J.T. Metal–organic framework materials as catalysts. Chem. Soc. Rev. 2009, 38, 1450–1459. [Google Scholar] [CrossRef]
  57. Wang, Q.; Astruc, D. State of the Art and Prospects in Metal–Organic Framework (MOF)-Based and MOF-Derived Nanocatalysis. Chem. Rev. 2020, 120, 1438–1511. [Google Scholar] [CrossRef]
  58. Park, K.S.; Ni, Z.; Cote, A.P.; Choi, J.Y.; Huang, R.D.; Uribe-Romo, F.J.; Chae, H.K.; O’Keeffe, M.; Yaghi, O.M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. USA 2006, 103, 10186–10191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Wang, H.; Pei, X.; Kalmutzki, M.J.; Yang, J.; Yaghi, O.M. Large Cages of Zeolitic Imidazolate Frameworks. Acc. Chem. Res. 2022, 55, 707–721. [Google Scholar] [CrossRef]
  60. Na, K.; Choi, M.; Yaghi, O.M.; Somorjai, G.A. Metal Nanocrystals Embedded in Single Nanocrystals of MOFs Give Unusual Selectivity as Heterogeneous Catalysts. Nano Lett. 2014, 14, 5979–5983. [Google Scholar] [CrossRef]
  61. Rungtaweevoranit, B.; Baek, J.; Araujo, J.R.; Archanjo, B.S.; Choi, K.M.; Yaghi, O.M.; Somorjai, G.A. Copper Nanocrystals Encapsulated in Zr-based Metal–Organic Frameworks for Highly Selective CO2 Hydrogenation to Methanol. Nano Lett. 2016, 16, 7645–7649. [Google Scholar] [CrossRef] [PubMed]
  62. Michaud, P.; Astruc, D.; Ammeter, J.H. Electron-transfer pathways in the reduction of d6 and d7 organoiron cations by lithium tetrahydroaluminate and sodium tetrahydroborate. J. Am. Chem. Soc. 1982, 104, 3755–3757. [Google Scholar] [CrossRef]
  63. Wang, C.; Tuminetti, J.; Wang, Z.; Zhang, C.; Ciganda, R.; Moya, S.; Ruiz, J.; Astruc, D. Hydrolysis of Ammonia-Borane over Ni/ZIF-8 Nanocatalyst: High Efficiency, Mechanism, and Controlled Hydrogen Release. J. Am. Chem. Soc. 2017, 139, 11610–11615. [Google Scholar] [CrossRef] [Green Version]
  64. 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]
  65. Wang, C.; Liu, X.; Wu, Y.; Astruc, D. PtNi@ZIF-8 nanocatalyzed high efficiency and complete hydrogen generation from hydrazine borane: Origin and mechanistic insight. J. Mater. Chem. A 2022, 10, 17614–17623. [Google Scholar] [CrossRef]
  66. Kang, N.; Wei, X.; Shen, R.; Li, B.; Guisasola, E.; Cal, E.; Moya, S.; Salmon, L.; Wang, C.; Coy, E.; et al. Fast Au-Ni@ZIF-8-catalyzed ammonia borane hydrolysis boosted by dramatic volcano-type synergy and plasmonic acceleration. Appl. Catal. B 2023, 320, 121957. [Google Scholar] [CrossRef]
  67. Zhao, Q.; Liu, X.; Astruc, D. Tetrahydroxydiboron (bis-boric acid), a versatile reagent for borylation, hydrogenation, catalysis, radical reactions and H2 generation. Eur. J. Inorg. Chem. 2023. [Google Scholar] [CrossRef]
  68. Hou, C.-C.; Li, Q.; Wang, C.-J.; Peng, C.-Y.; Chen, Q.-Q.; Ye, H.-F.; Fu, W.-F.; Che, C.-M.; López, N.; Chen, Y. Ternary Ni–Co–P nanoparticles as noble-metal-free catalysts to boost the hydrolytic dehydrogenation of ammonia-borane. Energy Environ. Sci. 2017, 10, 1770–1776. [Google Scholar] [CrossRef] [Green Version]
  69. Sun, D.; Hao, Y.; Wang, C.; Zhang, X.; Yu, X.; Yang, X.; Li, L.; Lu, Z.; Shang, W. TiO2–CdS supported CuNi nanoparticles as a highly efficient catalyst for hydrolysis of ammonia borane under visible-light irradiation. Int. J. Hydrogen Energy 2020, 45, 4390–4402. [Google Scholar] [CrossRef]
  70. Wang, Q.; Fu, F.; Yang, S.; Martinez Moro, M.; Ramirez, M.; Moya, S.; Salmon, L.; Ruiz, J.; Astruc, D. Dramatic Synergy in CoPt Nanocatalysts Stabilized by “Click” Dendrimers for Evolution of Hydrogen from Hydrolysis of Ammonia Borane. ACS Catal. 2019, 9, 1110–1119. [Google Scholar] [CrossRef]
  71. Linic, S.; Aslam, U.; Boerigter, C.; Morabito, M. Photochemical transformations on plasmonic metal nanoparticles. Nat. Mater. 2015, 14, 567–576. [Google Scholar] [CrossRef] [PubMed]
  72. Wang, C.; Astruc, D. Nanogold plasmonic photocatalysis for organic synthesis and clean energy conversion. Chem. Soc. Rev. 2014, 43, 7188–7216. [Google Scholar] [CrossRef] [Green Version]
  73. Gellé, A.; Jin, T.; de la Garza, L.; Price, G.D.; Besteiro, L.V.; Moores, A. Applications of Plasmon-Enhanced Nanocatalysis to Organic Transformations. Chem. Rev. 2020, 120, 986–1041. [Google Scholar] [CrossRef] [PubMed]
  74. Zada, A.; Muhammad, P.; Ahmad, W.; Hussain, Z.; Ali, S.; Khan, M.; Khan, Q.; Maqbool, M. Surface Plasmonic-Assisted Photocatalysis and Optoelectronic Devices with Noble Metal Nanocrystals: Design, Synthesis, and Applications. Adv. Funct. Mater. 2019, 30, 1906744. [Google Scholar] [CrossRef]
  75. Wen, M.; Kuwahara, Y.; Mori, K.; Yamashita, H. Enhancement of Catalytic Activity Over AuPd Nanoparticles Loaded Metal Organic Framework Under Visible Light Irradiation. Top. Catal. 2016, 59, 1765–1771. [Google Scholar] [CrossRef]
  76. Rej, S.; Hsia, C.-F.; Chen, T.-Y.; Lin, F.-C.; Huang, J.-S.; Huang, M.H. Facet-Dependent and Light-Assisted Efficient Hydrogen Evolution from Ammonia Borane Using Gold–Palladium Core–Shell Nanocatalysts. Angew. Chem. Int. Ed. 2016, 55, 7222–7226. [Google Scholar] [CrossRef]
  77. Verma, P.; Yuan, K.; Kuwahara, Y.; Mori, K.; Yamashita, H. Enhancement of plasmonic activity by Pt/Ag bimetallic nanocatalyst supported on mesoporous silica in the hydrogen production from hydrogen storage material. Appl. Catal. B 2018, 223, 10–15. [Google Scholar] [CrossRef]
  78. Kang, N.; Wang, Q.; Djeda, R.; Wang, W.; Fu, F.; Martinez Moro, M.; Ramirez, M.A.S.; Moya, E.; Coy, L.; Salmon, J.-L.; et al. Visible-Light Acceleration of H2 Evolution from Aqueous Solutions of Inorganic Hydrides Catalyzed by Gold-Transition-Metal Nanoalloys. ACS Appl. Mater. Interfaces 2020, 12, 53816–53826. [Google Scholar] [CrossRef]
  79. Zhang, Y.C.; He, S.; Guo, W.X.; Hu, Y.; Huang, J.W.; Mulcahy, J.R.; Wei, D. Surface-Plasmon-Driven Hot Electron Photochemistry. Chem. Rev. 2018, 118, 2927–2954. [Google Scholar] [CrossRef]
  80. Zhang, P.; Wang, T.; Gong, J.L. Mechanistic Understanding of the Plasmonic Enhancement for Solar Water Splitting. Adv. Mater. 2015, 27, 5328–5342. [Google Scholar] [CrossRef]
  81. Kalidindi, S.B.; Sanyal, U.; Jagirdar, B.R. Nanostructured Cu and Cu@Cu2O core shell catalysts for hydrogen generation from ammonia–borane. Phys. Chem. Chem. Phys. 2008, 10, 5870–5874. [Google Scholar] [CrossRef] [Green Version]
  82. Kalidindi, S.B.; Vernekar, A.A.; Jagirdar, B.R. Co–Co2B, Ni–Ni3B and Co–Ni–B nanocomposites catalyzed ammonia–borane methanolysis for hydrogen generation. Phys. Chem. Chem. Phys. 2009, 11, 770–775. [Google Scholar] [CrossRef] [PubMed]
  83. Sun, D.H.; Li, P.Y.; Xu, Y.; Huang, J.L.; Li, Q. Monodisperse AgPd alloy nanoparticles as a highly active catalyst towards the methanolysis of ammonia borane for hydrogen generation. RSC Adv. 2016, 6, 105940–105947. [Google Scholar] [CrossRef] [Green Version]
  84. Li, X.; Zhang, C.; Luo, M.; Yao, Q.; Lu, Z.-H. Ultrafine Rh nanoparticles confined by nitrogen-rich covalent organic frameworks for methanolysis of ammonia borane. Inorg. Chem. Front. 2020, 7, 1298–1306. [Google Scholar] [CrossRef]
  85. Li, X.; Yao, Q.; Li, Z.; Li, H.; Zhu, Q.-L.; Lu, Z.-H. Modulation of Cu and Rh single-atoms and nanoparticles for high-performance hydrogen evolution activity in acidic media. J. Mater. Chem. A 2021, 10, 326–336. [Google Scholar] [CrossRef]
  86. Luo, W.; Cheng, W.; Hu, M.; Wang, Q.; Cheng, X.; Zhang, Y.; Wang, Y.; Gao, D.; Bi, J.; Fan, G. Ultrahigh Catalytic Activity of l-Proline-Functionalized Rh Nanoparticles for Methanolysis of Ammonia Borane. ChemSusChem 2019, 12, 535–541. [Google Scholar] [CrossRef] [PubMed]
  87. Caner, N.; Yurderi, M.; Bulut, A.; Kanberoglu, G.S.; Kaya, M.; Zahmakiran, M. Chromium based metal-organic framework MIL-101 decorated palladium nanoparticles for the methanolysis of ammonia-borane. New J. Chem. 2020, 44, 12435–12439. [Google Scholar] [CrossRef]
  88. Li, H.; Yao, Z.; Wang, X.; Zhu, Y.; Chen, Y. Review on Hydrogen Production from Catalytic Ammonia Borane Methanolysis: Advances and Perspectives. Energy Fuels 2022, 36, 11745–11759. [Google Scholar] [CrossRef]
  89. Özkar, S. Transition metal nanoparticle catalysts in releasing hydrogen from the methanolysis of ammonia borane. Int. J. Hydrogen Energy 2020, 45, 7881–7891. [Google Scholar] [CrossRef]
  90. Yurderi, M.; Bulut, A.; Ertas, İ.E.; Zahmakiran, M.; Kaya, M. Supported copper–copper oxide nanoparticles as active, stable and low-cost catalyst in the methanolysis of ammonia–borane for chemical hydrogen storage. Appl. Catal. B 2015, 165, 169–175. [Google Scholar] [CrossRef]
  91. Tunc, N.; Rakap, M. Preparation and characterization of Ni-M (M: Ru, Rh, Pd) nanoclusters as efficient catalysts for hydrogen evolution from ammonia borane methanolysis. Renew. Energy 2020, 155, 1222–1230. [Google Scholar] [CrossRef]
  92. Zhang, S.; Li, M.; Li, L.; Dushimimana, F.; Zhao, J.; Wang, S.; Han, J.; Zhu, X.; Liu, X.; Ge, Q.F.; et al. Visible-Light-Driven Multichannel Regulation of Local Electron Density to Accelerate Activation of O-H and B-H Bonds for Ammonia Borane Hydrolysis. ACS Catal. 2020, 10, 14903–14915. [Google Scholar] [CrossRef]
  93. Kang, N.; Shen, R.; Li, B.; Fu, F.; Espuche, B.; Moya, S.; Salmon, L.; Pozzo, J.-L.; Astruc, D. Dramatic acceleration by visible light and mechanism of AuPd@ZIF-8-catalyzed ammonia borane methanolysis for efficient hydrogen production. J. Mater. Chem. A 2023, 11, 5245–5256. [Google Scholar] [CrossRef]
  94. Jo, S.; Verma, P.; Kuwahara, Y.; Mori, K.; Choi, W.; Yamashita, H. Enhanced hydrogen production from ammonia borane using controlled plasmonic performance of Au nanoparticles deposited on TiO2. J. Mater. Chem. A 2017, 5, 21883–21892. [Google Scholar] [CrossRef]
  95. Ashby, C.; Dobbs, F.R.; Harry Hopkins, P.J., Jr. Composition of lithium aluminum hydride, lithium borohydride, and their alkoxy derivatives in ether solvents as determined by molecular association and conductance studies. J. Am. Chem. Soc. 1975, 97, 3158–3162. [Google Scholar] [CrossRef]
  96. Yang, X.F.; Wang, A.Q.; Qiao, B.T.; Li, J.; Liu, J.Y.; Zhang, T. Single-Atom Catalysts: A New Frontier in Heterogeneous Catalysis. Acc. Chem. Res. 2013, 46, 1740–1748. [Google Scholar] [CrossRef]
  97. Liu, L.C.; Corma, A. Metal Catalysts for Heterogeneous Catalysis: From Single Atoms to Nanoclusters and Nanoparticles. Chem. Rev. 2018, 118, 4981–5079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Liang, L.; Qin, F.; Wang, S.; Wu, J.; Li, R.; Wang, Z.; Ren, M.; Liu, D.; Wang, D.; Astruc, D. Overview of the materials design and sensing strategies of nanopore devices. Coord. Chem. Rev. 2023, 478, 214998. [Google Scholar] [CrossRef]
  99. Ouyang, L.Z.; Ouyang, J.; Chen, K.; Zhu, M.; Liu, Z.W. Hydrogen Production via Hydrolysis and Alcoholysis of Light Metal-Based Materials: A Review. Nano-Micro Lett. 2021, 13, 134. [Google Scholar] [CrossRef]
  100. Wang, J.; Gao, Y.; Kong, H.; Kim, J.; Choi, S.; Ciucci, F.; Hao, Y.; Yang, S.H.; Shao, Z.P.; Lim, J. Non-precious-metal catalysts for alkaline water electrolysis: Operando characterizations, theoretical calculations, and recent advances. Chem. Soc. Rev. 2020, 49, 9154–9196. [Google Scholar] [CrossRef]
Figure 1. Bonding in NH3BH3.
Figure 1. Bonding in NH3BH3.
Chemistry 05 00060 g001
Figure 2. Setup for the measurement of evolved H2: (A) thermocouple, (B) bath temperature, (C) round-bottom flask, (D) stop cock, (E) gas burette, (F) leveling flask, (G) magnetic stirrer, (H) water inlet, (I) water outlet, (J) reflux condenser, and (K) water trap. Reproduced with permission from Ref. [32]. Copyright 2007 American Chemical Society.
Figure 2. Setup for the measurement of evolved H2: (A) thermocouple, (B) bath temperature, (C) round-bottom flask, (D) stop cock, (E) gas burette, (F) leveling flask, (G) magnetic stirrer, (H) water inlet, (I) water outlet, (J) reflux condenser, and (K) water trap. Reproduced with permission from Ref. [32]. Copyright 2007 American Chemical Society.
Chemistry 05 00060 g002
Scheme 1. Reproduced with permission from Ref. [63]. Copyright 2017 American Chemical Society.
Scheme 1. Reproduced with permission from Ref. [63]. Copyright 2017 American Chemical Society.
Chemistry 05 00060 sch001
Figure 3. On–off control of H2 evolution during Ni NP@ZIF-8-catalyzed AB hydrolysis upon successive addition of NaOH (on) and HCl (off) in water. Reproduced with permission from Ref. [63]. Copyright 2017 American Chemical Society.
Figure 3. On–off control of H2 evolution during Ni NP@ZIF-8-catalyzed AB hydrolysis upon successive addition of NaOH (on) and HCl (off) in water. Reproduced with permission from Ref. [63]. Copyright 2017 American Chemical Society.
Chemistry 05 00060 g003
Scheme 2. Reproduced with permission from Ref. [63]. Copyright 2017 American Chemical Society.
Scheme 2. Reproduced with permission from Ref. [63]. Copyright 2017 American Chemical Society.
Chemistry 05 00060 sch002
Scheme 3. Proposed mechanism for AB hydrolysis catalyzed by PtCo alloy@“click” dendrimer. A similar mechanism was suggested for catalysis by PtNi alloy@ZIF-8. In both cases, the N atom coordination of the support framework to the alloy surface increases the intermetallic synergy in the alloy and synergy with the framework. (ac) successively represent the formation of the first, second and third mol hydrogen, and (d) shows the formation of the final salt. Reproduced with permission from Ref. [70]. Copyright 2019 American Chemical Society.
Scheme 3. Proposed mechanism for AB hydrolysis catalyzed by PtCo alloy@“click” dendrimer. A similar mechanism was suggested for catalysis by PtNi alloy@ZIF-8. In both cases, the N atom coordination of the support framework to the alloy surface increases the intermetallic synergy in the alloy and synergy with the framework. (ac) successively represent the formation of the first, second and third mol hydrogen, and (d) shows the formation of the final salt. Reproduced with permission from Ref. [70]. Copyright 2019 American Chemical Society.
Chemistry 05 00060 sch003
Figure 4. “Click” dendrimers (i.e., dendrimer assembled by click chemistry, vide supra) and a “click” polymer reference allowing the investigation of the dendrimer effect in NP stabilization and efficiency (top). The green-colored cycles in all the dendrimers represent 1,2,3-triazoles formed upon Cu-alkyne-azide catalysis (CuAAC) by the “click” reaction between polyazido-terminated dendrimer cores and dendrons containing a terminal alkyne at their focal point. The incorporation of triazole-coordinated metal ions inside the “click” dendrimer followed by reduction using NaBH4 to dendrimer-encapsulated NPs (bottom). Reproduced with permission from Ref. [70]. Copyright 2020 American Chemical Society.
Figure 4. “Click” dendrimers (i.e., dendrimer assembled by click chemistry, vide supra) and a “click” polymer reference allowing the investigation of the dendrimer effect in NP stabilization and efficiency (top). The green-colored cycles in all the dendrimers represent 1,2,3-triazoles formed upon Cu-alkyne-azide catalysis (CuAAC) by the “click” reaction between polyazido-terminated dendrimer cores and dendrons containing a terminal alkyne at their focal point. The incorporation of triazole-coordinated metal ions inside the “click” dendrimer followed by reduction using NaBH4 to dendrimer-encapsulated NPs (bottom). Reproduced with permission from Ref. [70]. Copyright 2020 American Chemical Society.
Chemistry 05 00060 g004
Scheme 4. Mechanism of AuPd@ZIF-8 catalysis of AB methanolysis in the dark and under visible-light irradiation. The initial step involves hydride transfer from AB to the catalyst surface (or oxidative addition of the B-H bond followed by electron transfer from B to the NP), then H bonding [NP-H]–H–OCH3 between the nucleophilic negatively charged [NP-H] species and the electrophilic methanol proton (attracting CH3OH near the NP surface). This decreases the CH3O–H bond electron density, facilitating further oxidative addition of this O–H bond onto Pd, which is boosted by the plasmonic hot electron transfer under visible illumination. (a) Mechanistic aspects of the formation of the first mol H2. (b) formation of the second and and third mol H2. Reproduced with permission from Ref. [93]. Copyright 2023 Royal Society of Chemistry.
Scheme 4. Mechanism of AuPd@ZIF-8 catalysis of AB methanolysis in the dark and under visible-light irradiation. The initial step involves hydride transfer from AB to the catalyst surface (or oxidative addition of the B-H bond followed by electron transfer from B to the NP), then H bonding [NP-H]–H–OCH3 between the nucleophilic negatively charged [NP-H] species and the electrophilic methanol proton (attracting CH3OH near the NP surface). This decreases the CH3O–H bond electron density, facilitating further oxidative addition of this O–H bond onto Pd, which is boosted by the plasmonic hot electron transfer under visible illumination. (a) Mechanistic aspects of the formation of the first mol H2. (b) formation of the second and and third mol H2. Reproduced with permission from Ref. [93]. Copyright 2023 Royal Society of Chemistry.
Chemistry 05 00060 sch004
Scheme 5. AB hydrolysis and methanolysis and AB regeneration after solvolysis. The regeneration of AB from AB methanolysis needs a reaction with LiAlH4 and NH4Cl of the AB methanolysis product, NH4B(OCH3)4, in THF (0 °C to r.t., 90% yield), but the regeneration of the AB hydrolysis product NH4B(OH)4 is more complicated, because it first requires a reaction with methanol. LiAl(OMe)4 can be prepared by the disproportionation of LiAl(OMe)3H and is insoluble [95]. Inspired by Refs. [32,87]. Copyright 2007 and 2022. American Chemical Society.
Scheme 5. AB hydrolysis and methanolysis and AB regeneration after solvolysis. The regeneration of AB from AB methanolysis needs a reaction with LiAlH4 and NH4Cl of the AB methanolysis product, NH4B(OCH3)4, in THF (0 °C to r.t., 90% yield), but the regeneration of the AB hydrolysis product NH4B(OH)4 is more complicated, because it first requires a reaction with methanol. LiAl(OMe)4 can be prepared by the disproportionation of LiAl(OMe)3H and is insoluble [95]. Inspired by Refs. [32,87]. Copyright 2007 and 2022. American Chemical Society.
Chemistry 05 00060 sch005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kang, N.; Wang, C.; Astruc, D. Hydrogen Evolution upon Ammonia Borane Solvolysis: Comparison between the Hydrolysis and Methanolysis Reactions. Chemistry 2023, 5, 886-899. https://doi.org/10.3390/chemistry5020060

AMA Style

Kang N, Wang C, Astruc D. Hydrogen Evolution upon Ammonia Borane Solvolysis: Comparison between the Hydrolysis and Methanolysis Reactions. Chemistry. 2023; 5(2):886-899. https://doi.org/10.3390/chemistry5020060

Chicago/Turabian Style

Kang, Naixin, Changlong Wang, and Didier Astruc. 2023. "Hydrogen Evolution upon Ammonia Borane Solvolysis: Comparison between the Hydrolysis and Methanolysis Reactions" Chemistry 5, no. 2: 886-899. https://doi.org/10.3390/chemistry5020060

Article Metrics

Back to TopTop