Next Article in Journal
Fed-Batch Cultivation and Adding Supplements to Increase Yield of β-1,3-1,4-Glucanase by Genetically Engineered Escherichia coli
Next Article in Special Issue
CpCo(III) Precatalysts for [2+2+2] Cycloadditions
Previous Article in Journal
Regioselective Hydroxylation of Oleanolic Acid Catalyzed by Human CYP3A4 to Produce Hederagenenin, a Chiral Metabolite
Previous Article in Special Issue
Cobalt-Based Metal Organic Frameworks as Solids Catalysts for Oxidation Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Applications of Co-B Catalysts in NaBH4-Based Portable Hydrogen Generators

Boreskov Institute of Catalysis SB RAS, Lavrentieva Av. 5, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(2), 268; https://doi.org/10.3390/catal11020268
Submission received: 29 January 2021 / Revised: 15 February 2021 / Accepted: 15 February 2021 / Published: 17 February 2021
(This article belongs to the Special Issue Cobalt Catalysis: Recent Progress and Developments)

Abstract

:
This review highlights the opportunities of catalytic hydrolysis of NaBH4 with the use of inexpensive and active Co-B catalysts among the other systems of hydrogen storage and generation based on water reactive materials. This process is important for the creation of H2 generators required for the operation of portable compact power devices based on low-temperature proton exchange membrane fuel cells (LT PEM FC). Special attention is paid to the influence of the reaction medium on the formation of active state of Co-B catalysts and the problem of their deactivation in NaBH4 solution stabilized by alkali. The novelty of this review consists in the discussion of basic designs of hydrogen generators based on NaBH4 hydrolysis using cobalt catalysts and the challenges of their integration with LT PEM FC. The potential of using batch reactors in which there is no need to use aggressive alkaline NaBH4 solutions is discussed. Solid-phase compositions or pellets based on NaBH4 and cobalt-containing catalytic additives are proposed, the hydrogen generation from which starts immediately after the addition of water. The review made it possible to formulate the most acute problems, which require new sci-tech solutions.

Graphical Abstract

1. Introduction

The accumulation of electric energy in the hydrogen cycle is one of the main scenarios for the transition of the world power industry to carbon-free energy carriers. Electrochemical energy converters represented by fuel cells with a high efficiency (up to 70%) are proposed for the efficient involvement of hydrogen in the process of electric energy generation. Many efforts have been made for the development of power devices based on low-temperature solid-polymer fuel cells with proton exchange membrane (LT PEM FC). The rapid oxidation of H2 and the formation of water as a sole by-product are advantageous for the creation of small-sized and low-weight devices [1]. However, there is a problem that hinders the commercialization of fuel cells in the power source market, namely, the lack of suitable methods for compact, safe, and efficient hydrogen storage [2,3,4,5,6,7,8]. This problem is most acute in the development of small-sized energy devices, which require compact gas generators providing a high yield of hydrogen per unit weight or volume without additional heating and LT PEM FC poisoning. The indicated requirements restrain the choice of hydrogen-generating materials as the potential energy carriers.
Conventional methods of hydrogen storage in compressed (15–70 MPa) or liquefied (−253 °C and 0.2 MPa) form are unacceptable for use in compact power sources due to the low hydrogen content in a unit volume (0.042 g·cm−3 at 70 MPa for compressed H2 and 0.07 g·cm−3 for liquefied H2), large sizes of cryogenic plants, high energy consumption, explosion hazard, and evaporation losses up to 3 vol% per day. It is known that hydrogen readily diffuses through metals, thus destroying them [9].
Among various hydrogen-containing compounds, a high gravimetric and volumetric content of hydrogen is typical of hydrides, particularly borohydrides [10,11,12,13]. First, they exceed other compounds in the content of hydrogen. For example, hydrogen density in KBH4 is 0.083 g·cm−3; in NaBH4, 0.112 g·cm−3; and in LiBH4, 0.123 g cm−3; these values are higher than the density of liquid hydrogen. Second, in the case of their interaction with water, the yield of hydrogen increases due to the involvement of water in the chemical process. However, the hydrolysis of hydrides is accompanied by the release of a great amount of heat per mole of the produced hydrogen [14,15]. The hydrolysis of NaBH4 is characterized by a low thermal effect (53–80 kJ per mole of H2 [15,16]); hence, the entire process is safe and controllable. This hydride has an acceptable price, and its hydrolysis products are environmentally benign [17,18]. In this connection, NaBH4 has been thoroughly studied in the last 20 years as a compact form of hydrogen storage. The choice of acid accelerators and catalysts for the hydrolysis of NaBH4 and the organization of the stable and controllable generation of H2 in this process are the widely discussed subjects. A model where hydrogen is produced through the electrolysis of water, taking advantage of the electrical energy produced by a renewable generator (photovoltaic panels), and chemically stored by the synthesis of NaBH4, is discussed in [19]. It was shown that due to the safety and high volumetric density of NaBH4 it will be a promising technology for transportation to remote sites where hydrogen is released from NaBH4 hydrolysis and used for energy production.
This review considers the main classes of water-reactive materials for hydrogen production. The focus is on hydrogen production via the NaBH4 hydrolysis. Achievements in acid hydrolysis and catalytic hydrolysis with inexpensive and active cobalt catalysts are discussed, with the main emphasis on their transformations in the reaction medium. The new data on the preparation and investigation of solid-phase and pelletized compositions containing NaBH4 and a cobalt catalyst are reported. The design of hydrogen generators based on NaBH4 hydrolysis with cobalt catalysts and the problems of their integration with LT PEM FC are reviewed. The data obtained are generalized to outline main directions of further studies that are promising for the development of autonomous compact power devices based on LT PEM FC.

2. NaBH4 as Water-Reactive Material for Hydrogen Storage and Production

The analysis of literature has revealed two main directions in the development of portable systems for hydrogen storage and generation: the physical storage of adsorbed hydrogen and the chemical storage of bound hydrogen. The chemical storage of hydrogen as stable chemical compounds is commonly characterized by high hydrogen density, minimum requirements to auxiliary infrastructure, and low energy consumption [20]. However, there are some problems: the sensitivity of materials to storage conditions, stability in iterative hydrogenation-dehydrogenation cycles, the regeneration of dehydrogenation products, the cost and accessibility of components of a hydrogen generation system, etc.
Current studies in the field are focused also on achieving the target values of volumetric (VHSC, g·cm−3) and gravimetric (GHSC, wt.%) hydrogen storage capacity, complying with the requirements to temperature and hydrogen release rate, and establishing a mechanism of the hydrogenation–dehydrogenation reactions, particularly the catalyst- assisted reactions. The materials under consideration can be classified as reversible or irreversible storage systems. In reversible systems, a material can be recharged with gaseous hydrogen directly “on board” [20,21,22,23,24,25,26,27,28,29], whereas irreversible systems are regenerated chemically, “off board”.
Table 1 lists the main hydrogen storage systems based on hydrolysis or the involvement of water in the hydrogen generation process, in which the spent fuels or byproducts should be removed from the reactor and regenerated “off board”. All these processes are exothermic; they often do not require an additional heat supply and can be performed at ambient temperature, which is an essential advantage. Here, it is difficult to perform the reaction with a high yield using a limited amount of water to provide high GHSC values [30]. Serious problems emerge with heat and mass transfer, which are related to intense heating of the reaction medium, the encapsulation of unreacted hydride with hydrolysis products, and the deposition of hydrolysis products on the reactor walls and catalyst surface, leading to catalyst deactivation. This is why the literature provides data mostly for the diluted solutions with low GHSC.
The analysis of data listed in Table 1 revealed that sodium borohydride (NaBH4) is the most promising inexpensive material for hydrogen storage in terms of the operating temperature, stability, and safety in air. In addition, its catalytic hydrolysis has been thoroughly studied with respect to hydrogen generation and the regeneration of hydrolysis products (Figure 1). Note that the involvement of water increases twofold the hydrogen yield upon hydrolysis as compared to thermolysis. Compact hydrogen generators based on sodium borohydride are being developed in many countries: USA [39,40,41,42], Japan [43], South Korea [44,45,46], China [43], France [47], Portugal [48,49], Italy [50,51], Taiwan [52], Russia [53,54], and others. A review of NaBH4-based generators with the use of Co-containing catalyst will be given in Section 7. It should be noted that a somewhat greater interest in the hydrolysis of ammonia borane (Figure 1) is caused by the novelty of this compound and its high stability in air [55]; nevertheless, the high cost of ammonia borane and the presence of ammonia admixture in the evolved hydrogen limit its practical application. Thus, the review further considers studies on the systems for hydrogen generation and storage that are based on the hydrolysis of NaBH4.
Sodium borohydride is known to be highly stable in the solid state in the absence of moisture and carbon dioxide [18]. Upon dissolution, it slowly reacts with water, releasing 4 moles of hydrogen per 1 mole of NaBH4:
NaBH4 + 4H2O → Na+ + B(OH)4 + 4H2↑.
According to [16,39,56], a thermal effect of reaction (1) is 210–300 kJ/mol NaBH4. In the case of insufficient heat removal and high concentration of NaBH4, the temperature may increase, which may lead to a vigorous spontaneous release of H2. The reaction product is sodium borate, the form of which depends on the pH of the solution: at pH < 9, B(OH)3 prevails, whereas at pH > 9, the main species is B(OH)4 [57]. If the reaction product is dried at a temperature below 110 °C, it is present as NaBO2·2H2O and NaBO2·4H2O [58].
The rate of hydrogen evolution in the spontaneous hydrolysis of sodium borohydride decreases with time, and the complete conversion of the hydride is not achieved. This is related to an increase in the pH of the reaction medium due to the accumulation of sodium borate. This property formed a basis for the stabilization of NaBH4 aqueous solutions by an alkali (NaOH) [59,60], thus providing their long-term storage without considerable losses in the hydrogen capacity. A complete and rapid release of hydrogen from NaBH4 solutions (both aqueous and stabilized by NaOH) is possible only in the presence of acids and catalysts.

3. Acid Hydrolysis of NaBH4

This method was considered by Schlesinger et al. in the well-known paper published in 1953, which was devoted to the hydrolysis of sodium borohydride [61]. The development of hydrogen economy gave an impetus to investigation of such systems. Since the GHSC is among the key parameters [62], the most popular are such designs of hydrogen generators where a limited volume of acid solution with the optimized concentration is fed to solid NaBH4 at a controllable rate [63,64]. The problems to be solved are the necessity of increasing the hydrogen yield under water deficit conditions and the formation of solid products hindering the contact between hydride and acid accelerator [65].
The application of strong inorganic acids (HCl, H2SO4) instead of weaker organic acids ensures a higher yield of hydrogen [63,65] and makes it possible to decrease the amounts of acid and water, thus providing high GHSC values. According to [66], the acid hydrolysis of NaBH4 in the case of hydrochloric acid can be represented as follows (2):
NaBH4(s) + xHCl + (x + y + 2)H2O → xNaCl + xH3BO3 + (1 − x)NaBO2 + yH2O + 4H2.
The use of 2.74 M HCl at x = 0.19 and y = 4.74 gave GHSC = 5.1%, taking into account the weight of all the chemical reagents in the system. The optimization of conditions of this process was described in detail in [67].
A high aggressiveness of strong acids imposes substantial limitations on the materials used in the construction of hydrogen generators, so the application potential of weaker acids in this process is still being investigated. Thus, the design of a hydrogen generator proposed in [68] implies the interaction of a stoichiometric amount of water with a mixture of solid reagents NaBH4 + H3BO3, which provides GHSC = 3.88%. Among the tested solutions of organic acids, high activity was observed for 12 N formic acid [65], acetic acid at a molar ratio CH3COOH/NaBH4 = 2 [63], as well as formic and malic acids at the concentration above 10% [64].
As shown in the literature, the process of acid hydrolysis of NaBH4 can be arranged in different ways. For example, heterogeneous acid catalysts can be employed [69,70], thermal interaction of NaBH4 with H2C2O4·2H2O is investigated [71], and the hydrolysis is performed using the ZnCl2 + NaBH4 composition, where hydrochloric acid is released in the reaction medium upon the hydrolysis of ZnCl2 [72]. A combination of low-concentrated acids (0.01, 0.1, or 0.25 M H3PO4; 0.25 M HCl; 0.25, 0.1 M CH3COOH) with catalysts (Ni, Cu) is used to hydrolyze the NaBH4 solution stabilized by alkali [73,74,75].
The purity of evolved hydrogen is very important for the acid hydrolysis of NaBH4. Conventional designs of the reactors include gas–liquid separators, water adsorbent, and acid vapor adsorbents [66,76]. Particular attention is paid to purification of hydrogen from alkali impurities that are carried by water in the gas flow [64]; however, a possible release of diborane (B2H6) is not mentioned. On the other hand, B2H6 can form via the interaction of sodium borohydride with inorganic acids [77]:
2NaBH4 + 2HCl → 2NaCl + B2H6 + 2H2.
Certainly, in diluted solutions, diborane is instantaneously hydrolyzed with a release of hydrogen. Nevertheless, when acid hydrolysis is carried out in the highly concentrated solutions of sulfuric and hydrochloric acids, not only acid impurities but also diborane are released to the hydrogen-containing gas [78]. This problem of the acid hydrolysis of NaBH4 deserves special attention.

4. Catalytic Hydrolysis of NaBH4

The most promising method used to obtain high-purity hydrogen from NaBH4 is its catalytic hydrolysis (1). Meanwhile, sodium borohydride is widely used as a reductant for obtaining highly dispersed metals and metalloids in a solution at ambient temperature [79], so the distinctive feature of catalysts for the hydrolysis of sodium borohydride is the possibility to form the active component from precursor compounds directly in the reaction medium under the action of the hydride (in situ).
Optimization of the catalyst makes it possible to control the hydrogen generation process [13,17,80,81,82] as well as to stop and launch it at the customer’s request [39]. In addition, the application of catalysts allows hydrogen generation to be performed in a wide temperature range from −40 to +60 °C. The possibility of generating hydrogen at negative temperatures is provided both by the low freezing point of NaBH4 solutions stabilized with an alkali [83] and by the heat that is rapidly released during the catalytic hydrolysis [16]. It should be noted that the produced hydrogen-containing gas does not contain diborane impurities, which were observed in the acid hydrolysis of NaBH4 [78].
The catalytic hydrolysis of NaBH4 is an intricate process, which is affected by the properties and concentration of catalyst, the concentration of NaBH4, the addition of NaOH, and the reaction temperature. At this stage of investigation, the mechanism and kinetics of this process remain debatable [82,84,85,86]. Way back in 1971, K. A. Holbrook and P. J. Twist studied the hydrolysis of an alkaline solution of NaBH4 by deuterated water in the presence of cobalt and nickel borides [87]. Mass-spectroscopic analysis of the evolved gas revealed that it contains one hydrogen atom from NaBH4 and one deuterium atom from D2O. The authors proposed the mechanism including dissociative chemisorption of the borohydride anion on the catalyst surface, charge transfer through the catalyst, and electrochemical decomposition of water as the key steps. This mechanism was verified by NMR spectroscopy for the Pd/C catalyst [88] and is used by researchers for interpreting various regularities that are observed in the hydrolysis of NaBH4 [86,89,90,91]. The kinetics of the NaBH4 catalytic hydrolysis is described using the Langmuir–Hinshelwood and Michaelis–Menten models, which take into account the step of BH4 adsorption on the catalyst surface and the effect of temperature [84,85,92].
The literature presents many studies on the catalytic hydrolysis of NaBH4 in the presence of catalysts containing noble metals [39,81,82,93,94,95]. According to independently performed studies [40,96,97], the highest activity in the NaBH4 hydrolysis was shown by rhodium, ruthenium, and platinum compounds. The catalytic activity decreases in the following order: Rh > Ru > Pt >> Pd [97]; this sequence is retained also when metals are deposited on different supports. The research studies on Rh catalysts are restricted by their high cost. Investigation of the catalytically active phase formed from compounds of ruthenium, rhodium, platinum, and palladium in NaBH4 medium has shown that it consists of agglomerated nanosized particles with a metal crystal lattice [98,99,100]. A residual amount of boron was presented in Ru metal [101]. The comparison of activity of noble and non-noble catalysts was discussed in [81,82]. It can be seen that the activity of non-noble Co-B-based catalysts (15,000–35,000 mLH2·min−1·g−1cat) reaches the activity of Ru or Pt catalysts. So, the high cost of platinum metals strongly limits their wide application in the hydrogen generation systems.
The catalytic hydrolysis of NaBH4 is considered as one of the main commercially accessible ways of hydrogen generation and storage for LT PEM FC. In this connection, particular attention is paid in the literature to a search for cost-effective methods of hydrogen regeneration from the hydrolysis products—sodium borates [102,103], for example, by means of thermochemical [104,105,106,107,108,109] or mechanical activation methods [103,110,111,112]. The latter ones allow the process to be performed at room temperature, thus enhancing its energy efficiency.

5. Co-B Catalysts for the NaBH4 Hydrolysis

5.1. The Formation of Active Component in the Reaction Medium (In Situ)

In 1953, Schlesinger et al. demonstrated [61] that the introduction of soluble chlorides of transition metals in a sodium borohydride solution results in the formation of a black precipitate, which is catalytically active toward the NaBH4 hydrolysis. The activity of chlorides in this case decreased in the series CoCl2 > NiCl2 > FeCl2 > CuCl2 > MnCl2. Later works on the catalytic hydrolysis of NaBH4 confirm that the highest gas release rate is observed in the presence of cobalt compounds, whereas nickel and iron compounds are less active [61,82,91]. In this connection, many works in the last 20 years were devoted to the bulk and supported cobalt-containing systems whose active components were formed in an NaBH4 medium [13,86,113,114,115]. In distinction to platinum group metals, cobalt compounds are reduced with the formation of amorphous particles of the catalyst [116,117,118]. It is noted that the activity of the in situ synthesized catalytic systems exceeds that of the preliminarily (ex situ) synthesized systems [119,120,121,122].
Sodium borohydride is the most optimal reductant for cobalt compounds. This follows from a comparison of data on the reduction of cobalt chloride and oxide in the presence of NaBH4 or another boron-containing hydride—ammonia borane (NH3BH3), which is known to be a weaker reductant. It was shown [120,123] that the use of sodium borohydride in the reduction of cobalt chloride provides a higher reduction rate and results in the formation of a highly dispersed active state of the catalyst (Figure 2). A lower rate of cobalt chloride reduction in NH3BH3 medium is confirmed by the presence of an induction period on the hydrogen generation curve (Figure 2). According to [123], the average size of the particles formed from cobalt chloride under the action of sodium borohydride is ca. 30 nm. When cobalt chloride is reduced in a solution of ammonia borane, this leads to the formation of large particles with the diameter from 150 to 500 nm and the predominant size of ca. 250 nm. On the other hand, particles with the size < 10 nm and a low aggregation degree were obtained upon the reduction of CoCl2 by a NaBH4 solution in the presence of ammonia borane, the hydrolysis of which leads to the formation of NH4+ ions that are likely to have a stabilizing effect [124]. It should be noted that the literature provides a thorough description of the effect exerted by various factors determining the characteristics of amorphous cobalt catalysts, which are formed in a sodium borohydride medium: cobalt salt precursors [91,114,119,122,125], pH of the medium [91,126,127], solvent [91,128], NaBH4/Co ratio in solution [91,125,129], the order and manner of reagents introduction [127,129,130], changes in the composition of NaBH4 + NaOH solution upon storage [131], etc.
In distinction to salts, in the case of introduction of Co3O4 in both the NaBH4 and NH3BH3 media, the induction period with a very low rate of gas generation is observed on the hydrogen generation curve (Figure 2). However, in sodium borohydride solution after a few minutes, H2 generation starts and proceeds quite rapidly, whereas the rate of hydrogen evolution from ammonia borane solution remains insignificant even 6 h after the introduction of cobalt oxide [132]. It was found that within a short induction period, Co3O4 is reduced under the action of NaBH4 to form a new catalytically active ferromagnetic phase containing cobalt and boron [132]. Its accumulation in the course of reaction provided the growth of hydrogen generation rate. Thus, a correlation is observed between the activity and magnetic susceptibility of the cobalt catalyst that is formed in the reaction medium. Characteristics of the cobalt oxide samples (method of synthesis, morphology, defectness, crystallinity, etc.) have a substantial effect on this process [53,114,133]. For example, it was shown in [53] that defectness of the Co3O4 sample, which was synthesized by the calcination of cobalt chloride, provided its rapid reduction in the reaction medium, leading to the formation of the most active state of the catalyst. Significant differences were observed in the temperature dependences of magnetic susceptibility for such catalyst and the catalyst synthesized from the well-crystallized Co3O4. In addition, the presence of alkali in the reaction medium as well as sodium borate, which is accumulated due to the spontaneous hydrolysis of NaBH4 upon storage of the solution, exerts a detrimental effect on the formation of the active phase from Co3O4 [53,134].

5.2. The State of the In Situ Formed Active Component of Co-B Catalysts

The nature of the catalyst active phase formed in the reaction medium from cobalt compounds during the hydrolysis of sodium borohydride is still a debatable question. Investigation of the active phase is complicated by the fact that catalysts are examined after their withdrawal from the reducing reaction medium followed by washing and drying. The number of publications on the subject is ever growing, and review papers summarizing new information are sometimes published [79,86,91,113,115,135].
It was shown that the catalyst particles formed in a sodium borohydride solution from cobalt compounds have a spherical shape [136,137] and are usually in the aggregated state [127,138,139]. The average size of the particles and their size distribution depend on the reduction conditions. For example, in [140], the synthesized samples had the particle diameter from 10 to 100 nm and a narrow particle size distribution with the maximum near 30 nm. Such particle sizes are most typical of the systems under consideration [127,137,138]. It is known that solutions of reverse micelles [136,141] or surfactants [142] used in the synthesis can prevent the aggregation of the particles and reduce their size (4–10 nm) [136,141].
The active phase of the catalysts, which are formed upon the reduction of cobalt compounds in a NaBH4 solution, is X-ray amorphous and ferromagnetic. Its X-ray diffraction (XRD) pattern usually contains a low-informative broad line at 2θ = 45 ± 5 [125,127,129,140,143]. In high-resolution transmission electron microscopy (HR TEM) data, it shows up as the diffuse rings (halo), indicating the amorphous structure of the particles [128,144,145].
It is well known that such catalysts contain Co, B, and O. The Co:B ratio is most often close to 2 [123,129,138,139,140], but it may vary depending on the reduction conditions of samples [113,139,146].
2CoCl2 + 4NaBH4 + 9H2O → Co2B + 4NaCl + 12.5H2 + 3B(OH)3
Different content of oxygen is determined mostly by the degree of inertness of the atmosphere in which drying of the sample was performed [79]. According to HR TEM with energy-dispersive X-ray (EDX) analysis [123,139,147,148,149], nanosized particles of the catalysts are surrounded by a 2–5 nm thick shell containing oxygen compounds of cobalt and boron. This is confirmed by X-ray photoelectron spectroscopy (XPS) and infrared (IR) spectroscopy data indicating that the shell consists mostly of amorphous cobalt borate [139,149,150]. The formation of the shell may be related to the oxidation of an active component in an alkaline solution in the presence of sodium metaborate [131,139].
Indeed, according to XPS data, cobalt is present in the samples in two states, which correspond to Co0 and Co2+ in cobalt oxides, hydroxide and borate [113,149,151]. The Co0/Co2+ ratio considerably increases after the removal of the upper oxidized layer with a thickness no smaller than 3 nm by argon bombardment [147]. In distinction to cobalt, such treatment of the sample has only a slight effect on the ratio of oxidized and non-oxidized boron species when moving from the shell to the core [147,152].
Much attention is still focused on the composition and structure of the X-ray amorphous cobalt-containing core. In 2011, Arzac et al. published a paper [149] where examination by various methods (HR TEM, scanning TEM, electron energy loss spectroscopy, XRD, and XPS) revealed that the core of the 20–40 nm particle covered with the Co(BO2)2 shell contains 1–3 nm microcrystallites of metallic Co with the hexagonal crystal structure that are stabilized in the amorphous single- or multiphase matrix consisting of CoxB, CoxOy, Co(BO2)2, and B2O3. The amorphous matrix prevents the oxidation and growth of cobalt clusters. Such a microstructure was observed in [153] for the nickel–boride system.
It should be noted that important additional information on the state of the core in amorphous catalysts containing Co0 and B0 can be obtained by magnetic methods. The residual magnetization typical of amorphous Co-B samples is lower than that of metallic cobalt and crystalline cobalt borides that are formed in the high-temperature synthesis [117,154]. However, their Curie temperatures (Tc) are very close to each other. For example, for amorphous cobalt borides with the molar ratio Co:B = 2:1 and Co:B = 3:1, Tc is equal to 155 and 452 °C, respectively [155]. These values are close to Tc of crystalline cobalt borides Co2B (156 °C) and Co3B (474 °C) [156]. The magnetic moment per Co atom and the Curie temperature decrease with increasing the boron content in the cobalt environment [155,157,158]. As shown by calculations [158], such weakening of ferromagnetism of the amorphous samples may be caused by changes in the electronic structure owing to hybridization of the electronic states of cobalt (s-d) and boron (s-p). Along with the boron content, magnetic properties of amorphous cobalt borides are affected also by their oxidation state [139], size [136,154], agglomeration of particles [154], and their thermal treatment [117,154]. The informativeness of this method concerning the state of Co-B catalysts synthesized in the medium of boron-containing hydrides was demonstrated in our previous works [53,132,147,148,159].
It was known that hydrogen enters the composition of samples that were reduced by NaBH4 in organic solvent; in this case, the formation of bridge M–H–B structures with terminal B–H bonds was supposed [79]. For example, the reduction of cobalt chloride in ethanol led to the formation of a sample with the stoichiometry Co2B1H0.6, the desorption of hydrogen from which was observed upon heating in vacuum up to 100 °C and higher temperatures [160]. Treatment of the sample with water was also accompanied by a release of gaseous H2 and formation of H3BO3, which decreased the content of hydrogen and boron [160]. On the other hand, according to [116], the presence of hydrogen was detected in the composition of samples obtained in an aqueous solution. In that work, a series of Co-B samples was synthesized with the hydrogen content varying from 0.2 to 0.8 wt.%; therewith, the hydrogen content decreased with increasing the boron content in the samples. Further studies confirmed the presence of hydrogen in the sample obtained upon the reduction of cobalt chloride by sodium borohydride in an aqueous solution [161]. The molar ratio of elements was Co:B:O:H = 3.2:1.5:1.4:1 (or 0.46 wt.% H). Hydrogen release on the differential thermal analysis curve corresponded to the maximum at 230 °C [161], which is the decomposition temperature of higher borohydrides [162].
The magnetic susceptibility technique, XPS, EXAFS, XRD, and thermal analysis [147,161] were used to refine the model of the X-ray amorphous core, the structural elements of which include cobalt tetramers with the shortest Co–Co distance of ca. 2.5 Å and the coordination number equal to 2.7. It was supposed for the first time that such tetramers are stabilized in the hollows of a borohydride matrix. Taking into account X-ray diffraction data on the regular arrangement of cobalt clusters in the amorphous matrix (the distance between adjacent clusters is 5.8 Å), it was hypothesized that cobalt tetramers that formed in the sodium borohydride solution are involved in the build-up of the borohydride structure from BH4 anions. The calculations by Density Functional Theory confirmed that ferromagnetic tetramers can exist in the borohydride matrix isolated from the environment by the amorphous shell, which may be the product of their decomposition in an aqueous medium. We think that further studies are needed to detail the state of hydrogen in such compounds.

5.3. The Role of Boron in the Activity of Co-B Catalysts for the NaBH4 Hydrolysis

In 2016 [163], direct evidence indicating that boron is essential for the activity of cobalt catalysts in the hydrolysis of sodium borohydride was obtained. It was found that the introduction of boron via magnetron sputtering of cobalt on a nickel foil provided a more rapid evolution of hydrogen per cobalt atom as compared to the cobalt catalyst synthesized by a similar method without boron. The greater the boron amount in the sample, the higher the catalyst activity. This effect was attributed to an increase in the dispersion of the metallic phase. The replacement of boron by carbon in the catalyst synthesized by the same method led to a twofold decrease in activity in spite of the similar textural characteristics of cobalt–boride and cobalt–carbon catalysts. Therewith, the activity of cobalt–carbon catalysts increased during five cycles of their testing in NaBH4 hydrolysis [163]. A comparative XPS study of these catalysts before and after the reaction demonstrated the accumulation of boron in the samples. Thus, the formation of the Co-B phase on the surface of the cobalt–carbon catalyst may be considered as the main factor enhancing its activity toward the hydrolysis of sodium borohydride.
A relation between the activity of the catalysts and boron content is noted also in some other studies [114,120,125,164]. This effect may be related to changes in the electronic state of cobalt. Contacting with boron enriches cobalt atoms with electrons, thus protecting cobalt from oxidation and activating BH4 anions on the catalyst surface [120,143]. Therewith, calculations testify to the existence of the optimal stoichiometry between cobalt and boron, which provides high activity of the catalyst [165]. The site comprising two cobalt atoms and one boron atom is the most active one; this may be caused by the electronic structure of such a cluster, which provides the most favorable conditions for hydrogen binding.

5.4. Deactivation of the Co-B Catalysts in the NaBH4 Hydrolysis

Stability of the cobalt-boride catalysts in long-term tests is very important for their practical application. Table 2 lists changes in the activity of bulk and supported Co-B catalysts during their cyclic testing and long-term operation. One can see that the indicated catalytic systems are often deactivated under the conditions of NaBH4 hydrolysis. The activity of these catalysts changes to different extents, which may be related to differences in their physicochemical properties and testing conditions.
Researchers suppose that activity of the cobalt–boride catalysts may decrease due to agglomeration of the particles [159,174,175] caused by their magnetization [140]. In the case of supported catalysts, a decrease in the hydrogen release rate during the long-term testing may occur also due to leaching of the active component under vigorous release of H2 [176,177].
One of the main reasons of deactivation of the Co-B catalysts is the formation on their surface of a layer of strongly adsorbed borate anions [175]. It is difficult to remove this borate layer from the catalyst surface by washing with water [171]. Only the treatment with a diluted solution of sulfuric acid made it possible to destroy the borate layer of deactivated Co/nickel foam catalyst and provide a constant rate of hydrogen evolution in the next five cycles [171]. A similar result was obtained in [178] with the use of diluted hydrochloric acid.
The authors of [168] managed to restore the initial activity of the honeycomb Co-B catalyst after its operation for 30 h by a repeated plasma treatment. This effect was attributed to the reduction of the oxidized boron residing on the surface:
B2O3 + 6e + 3H2O → 2B + 6OH.
It should be noted that a prolonged contact of Co-B catalysts with an aqueous medium generally exerts a detrimental effect on their activity. It is known from the early works [138,179] that in water, cobalt borides are hydrolyzed to form boric acid and may completely transform into metallic cobalt upon boiling:
Co2B + 3H2O → 2Co + H3BO3 + 1.5H2.
The stability of transition metal borides directly depends on the content of boron [131,173]. Leaching of boron from the Co-B catalyst decreases its activity [114,164]. It is known that metallic cobalt nanopowders are low active in the hydrolysis of NaBH4 [119,146]. In addition, a higher magnetization typical of Co0 will lead to a greater aggregation of the catalyst particles in the reaction medium, thus decreasing the total surface area of the catalyst contacting the reaction medium [114,165].
XRD data obtained in our works show that the formation of Co0 and Co(OH)2 occurs when Co-B catalysts are tested in the NaBH4 solution stabilized with an alkali [114,131,180]. The application of IR spectroscopy allowed us to reveal the presence of a certain amount of other oxidized cobalt compounds (Co3O4, CoO, CoOOH) in the catalysts [114]. The surface oxidation of cobalt boride to cobalt hydroxide (7), when the pH of the solution increases due to accumulation of NaBO2, was also reported in [139,175]:
Co2B + OH + 7H2O → 2Co(OH)2 + B(OH)4 + 3.5H2.
A contact with a fresh portion of NaBH4 may lead to reactivation of the catalyst due to reduction of the oxidized cobalt compounds, as was supposed in [139]:
2Co(OH)2 + BH4 → Co2B + 0.5H2 + OH + 3H2O.
However, actually, there are only a few reports on the formation of the catalyst active phase from cobalt hydroxide under the action of NaBH4. Thus, the authors of [127] observed under certain conditions (CoCl2:NaBH4:NaOH = 1:6.5:1.9 in moles) the formation of colloidal Co(OH)2, the further reduction of which resulted in the formation of highly reactive ultradispersed Co-B.
The majority of studies show that cobalt hydroxides, both neat and with intercalated anions of precursor salts, are characterized by a slow reduction in the NaBH4 medium and hence by a low activity [114,119,122,150]. Raising the pH of the NaBH4 solution extends the induction period during which the active Co-B phase is formed from cobalt compounds in the reaction medium [53,131,134].
The oxidation of cobalt and accumulation of borate anions in the reaction medium are accompanied by the formation of cobalt borate, which, according to [129,130], can be summarily expressed as follows:
Co2+ + 2NaBO2 → 2Na+ + Co(BO2)2.
However, in distinction to cobalt hydroxide, cobalt borate is reduced quite readily after the introduction of fresh NaBH4 solution to form the active Co-B phase [150]. The decrease in the activity of Co-B catalysts upon treatment with water or sodium borate solution was not so pronounced as in the case of their treatment with an alkali [114]. This occurs because changes in the phase composition of the catalyst follow different routes. Leaching of boron from the catalyst under the action of water or sodium borate solution is weaker than in the case of alkali solution. In addition, the oxidized cobalt compounds (cobalt borate and oxide) are formed, which transform more readily into the active state upon the addition of a fresh portion of hydride [132,150].

6. Hydrogen Storage Systems Based on Solid-State NaBH4 Composites with Cobalt Catalysts

As it was discussed above in Section 5.4, cobalt-boride catalysts rapidly deactivate at high pH during the hydrolysis of NaBH4 solutions stabilized by alkali. One of the ways to provide sustainable hydrogen generation is to abandon the use of alkaline solutions. Some authors propose solid-phase hydrogen storage systems based on NaBH4 in a composition with cobalt catalyst (or its precursor), which include the dosing of a limited amount of water that initiates hydrogen release [47,181,182]. Another variant implies the dosing of a cobalt chloride solution on solid NaBH4 [43]. The obtained results listed in Table 3 indicate that optimization of the water amount, the content of cobalt catalysts, and the method of its introduction in the reaction medium allow high GHSCs to be achieved, although the 100% conversion of NaBH4 was not reached.
The application of pellets containing NaBH4 and anhydrous CoCl2 for hydrogen generation with the addition of water was first proposed by Schlesinger et al. [61,183], who noted that a decrease in the amount of introduced water leads to undesirable strong foaming and heating of the reaction medium. This study was further developed within the advancement of hydrogen economy. The application of the pellets containing NaBH4 and inexpensive cobalt catalyst has some advantages: a high content of hydrogen in a pellet (up to 10 wt.%), storage under moisture-proof conditions without hydrogen losses, the absence of hydride dust during the operation, and convenient application because gas generation starts immediately after water addition. Water can be taken from any natural source.
It was found that the pressing of a solid mixture of the hydride and catalytic additive as well as the uniform distribution of components over the composition substantially increase the hydrogen evolution rate [47,180,181]. The introduction of a precursor (CoCl2·6H2O) into sodium borohydride is more efficient in comparison with the pre-reduced Co-B catalysts because the formation of the active phase, which starts in the pressing step, provides a more active state of the catalyst [184,185]. It should be noted that the activity and dispersion of the catalyst formed under the action of NaBH4 are strongly affected by the amount of coordinated water in the composition of chloride hydrates (Co, Ni), which are used to prepare the pellets [186]. The rate of hydrogen generation can be controlled by varying the amount of catalytic additive [53] and the dispersion of the Co-B catalyst [180,187,188].
Current studies also reveal a considerable heating of the reaction medium (>100 °C) in the catalytic hydrolysis of NaBH4 performed with a small amount of water. This allows hydrogen to be generated at low ambient temperatures without additional heat input [61,182]. It is noted that the extent of heating of the reaction medium depends on the content of water [61,181] and catalyst [43,182]. As shown by the study reported in [43], a high H2 yield at a low content of water can be reached only upon its heating up to the melting point of NaBO2·2H2O (≈60 °C). On the other hand, heating of the reaction medium to the dehydration temperature of NaBO2·2H2O (>100 °C) increases the amount of water involved in the hydrolysis of NaBH4. This was observed in [182], where a high GHSC equal to 8.7 wt.% was reached. This value exceeds the yield of hydrogen according to the stoichiometry of reaction (1) (see Section 2), which theoretically corresponds to GHSC = 7.3%. The authors attribute this to the formation of NaBO2·nH2O, where n < 2.

7. Hydrogen Generators Based on Hydrolysis of NaBH4 Using Cobalt Catalysts

Fuel cells as electric energy generators are being applied now in unmanned air vehicles (UAV), warehouse equipment, submarines, householding, stationary battery backup, etc. They are simple in design, have zero atmospheric emission, high energy density, and high efficiency because chemical energy is converted directly to electric power. In addition, fuel cells are characterized by weak vibration, low noise, and low maintenance requirements [189]. Prototype power plants using LT PEM FC and hydrogen generator based on the hydrolysis of sodium borohydride in the presence of cobalt catalysts are reported in the literature (Table 4).
To provide the operation of PEM FC, pure hydrogen should be fed at a specified rate. This imposes certain requirements on the design of hydrogen generators and catalysts. As seen in Table 4, the designs ensuring a wide variation of hydrogen feed rate (from several milliliters to more than 100 L H2 per minute) have already been proposed for fuel cell batteries of different power levels intended for remote, portable, and mobile systems—2 W [52], 100 W [192,193,198,201,203,204], 200 W [44,194], 500 W [199], 3 kW [202].
The published studies are devoted mostly to the development of hydrogen generators with a catalytic flow reactor, where an NaBH4 solution stabilized by alkali is fed to the catalytic bed. The hydride solution is stored in a separate reservoir at room temperature and pumped into a reactor at a specified rate. In most cases, temperature is not controlled, and the hydrogen generation rate is determined by the contact time of NaBH4 solution with the catalyst and by its characteristics. The most common scheme of the setup is displayed in Figure 3.
Such arrangement of the process is complicated by safety regulations for fuels of the irritant nature. A more serious problem related to the use of NaBH4 solutions stabilized by alkali is the low GHSC value due to the great weight and dimensions of the fuel tank. The hydrolysis product—sodium metaborate—has a limited solubility in water, which decreases with an increase in pH [205]. To prevent its deposition on the catalyst and equipment, it is necessary to optimize the ratio of the hydride, water, and alkali in the solution [205]. This was made in [190], where a solution with the optimized composition (19% NaBH4 + 4.5% NaOH) was fed to the Co-B/Ni foam catalyst in order to produce hydrogen for the functioning of 60 W-scale LT PEM FC. The GHSC value of 3.5 wt.% was reached (the calculation took into account weights of the hydride solution and catalyst), which is quite high for such systems. However, this value is much lower than the United States Department of Energy targets [62]. The authors of [197] observed also the deposition of sodium borate in the case of 20 wt.% NaBH4 + 1 wt.% NaOH solution. The problem was solved by lowering the NaBH4 concentration to 18 wt.% [198]. Even in the presence of alkali, the hydride slowly interacts with water [206], which is accompanied by a gradual loss in hydrogen capacity. For example, after storage of the 15% NaBH4 + 5% NaOH solution for a year, the content of NaBH4 decreased by a factor of 3, and a partial deposition of the formed NaB(OH)4 (or NaBO2·2H2O) was observed [131].
Another essential aspect is associated with the characteristics of the catalyst. First, when hydrogen generators are functioning in a flow-type ON/OFF mode, it is preferable to use macrostructured granular or monolithic catalysts in order to decrease the gas and hydrodynamic resistance in the system. In addition, such catalysts can easily be separated from the reaction medium and removed from the reactor for replacement. Thus, in [192], a Co-B catalyst supported on ISOLITE ceramic material shaped as the reactor volume was successfully used. In [200], cation-exchange polymeric resin Amberlite IR-120 was used as a support for the Co-B catalyst. The authors of [203,204] used a highly active granular Co/γ-Al2O3 catalyst; however, in the long-term tests, they observed poor durability and adhesion of the supported active component as well as the destruction of the support grains. The most popular support is Ni foam, which not only demonstrates high stability in aggressive reaction medium but also provides efficient heat transfer over the reactor volume [190,193,202]. To increase the adhesion between cobalt-boride coating and Ni foam, the catalyst is heat-treated in an inert medium at moderate temperatures of 200–300 °C [190] and subjected to electroless plating [191,193].
A different approach to mounting a catalyst in the reaction zone was proposed in [50,195]. Magnets were placed in the reactor to hold a powdered ferromagnetic cobalt catalyst and prevent its washout with the liquid flow.
However, along with the washout of the catalyst active component, there is a problem of chemical deactivation of Co-B catalysts in the NaBH4 alkaline solution, as was shown in Section 5.4. To solve the problem, in [190], each cycle of hydrogen release was followed by reactivation of the Co-B/Ni foam catalyst—it was washed with water and immersed in the NaBH4 solution to reduce the oxidized cobalt compounds that formed upon testing. As a result, reusing of the catalyst under these conditions has been demonstrated at least three times without any deterioration of performance. A simple water washing of the catalyst removes the deposited borates from its surface, but it cannot restore the catalytic activity after a long-term cycle [198].
To extend the durability and enhance the activity, catalytic cobalt systems undergo a chemical modification. For example, Korean researchers produce hydrogen generators with the phosphorus-doped catalysts [44,191,197,198]. According to [207], the introduction of phosphorus increases the concentration of Co active sites on the catalyst surface. Another widely employed approach to improving the catalytic properties of cobalt catalysts is based on doping with transition metals. For example, the Co-Fe-Ni catalyst showed a superior stability during 20 hydrogen generation cycles of 2 h each [194].
In most cases, hydrogen generation is carried out without an additional heat supply because the hydrolysis of NaBH4 is an exothermic process. Self-heating is observed in a catalytic reactor; its rate and temperature depend on the catalyst activity [44] and the hydride solution feed rate [197]. Under certain conditions, such self-heating can exceed 100 °C [202] and lead to the evaporation of water. This creates several problems. A deficit of water in the system will result in an incomplete release of hydrogen and a decrease in GHSC. In addition, due to an intense evaporation of water, undissolved sodium borates will actively deposit on the equipment and on the catalyst surface [208]; sodium borates can also be released as impurities into the gas phase together with water vapor [64], thus deteriorating the operation of the Nafion membrane in fuel cells. To lower the temperature, the reactor is equipped with cooling fans [192,193], and the released hydrogen is purified. To this end, hydrogen is passed through reservoirs with water for removing water-soluble impurities [44,190,193,197,198] and then with silica gel for desiccation [44,193,196,197,198,203,204]. In [194], a resin was used for the capture of sodium from H2. Sometimes, H2 cooling devices are also employed—for example, a finned container with water and cooling fan [199].
Note that the optimization of reactor material of the hydrogen generator is one of the key tasks. On the one hand, the reactor material should meet the energy device requirements of light weight, small size, and high power density. On the other hand, to avoid the self-heating of the reaction medium and to control the hydrogen generation rate, materials with high heat transfer properties should be applied. In addition, the reactor material should capable of withstanding pressure and hydrogen embrittlement.
It seems interesting that the contribution of a non-catalytic route of NaBH4 hydrolysis will increase with the growth of reactor temperature. This suggests that the action of the catalyst is most important in the initial step, when the hydrogen generation process is launched. Indeed, the start-up problem is very essential. Thus, the target of the US Department of Energy’s fuel cell start-up time is less than 5 s [62]. This means that the instantaneous hydrogen generation from the sodium borohydride hydrolysis reaction must be achieved. To minimize the time of establishing a stable generation of hydrogen, several approaches were suggested in the literature. The authors of [199] used two catalysts: Co/γ-Al2O3 was placed at the top of the reactor, while the remaining volume was filled with Co-P/Ni foam. Upon feeding, the NaBH4 solution first contacted the highly active Co/γ-Al2O3, which increased the temperature and, accordingly, the rate of hydrogen evolution on the Co-P/Ni foam catalyst. In [197,200], the start-up time was reduced by varying the feed rate of NaBH4 solution. In the case of a highly concentrated 20 wt.% NaBH4 solution, first, a low feed rate of the solution was used, and then, it was increased [197], whereas in the case of a 5 wt.% NaBH4 solution, the process was initiated by the pulsed feeding of solution at a high rate, which was decreased later [200].
A design of the reactor also affects the hydrogen release parameters. The authors of [196] developed a π-shaped catalytic reactor, the shape of which facilitated a rapid removal of hydrogen from the catalyst surface, so as to increase the contact between sodium borohydride and the catalyst. This enhances the efficiency and stability of hydrogen generation in comparison with the reactor having a conventional shape.
Original configurations of hydrogen generators are proposed in order to decrease their weight and dimensions. In [192], the hydrogen generator featured a volume-exchange fuel tank, where the fuel and spent fuel exchange the volume within a single fuel tank. In [193], a novel concept using an all-in-one reactor was proposed in which a catalyst, hydrogen chamber, and by-product separator were combined in a volume. In addition, the reactor as well as a pump, cooling fans, valves, and controller were integrated in a single module. The produced hydrogen was not immediately sent to the fuel cell; it was stored in a hydrogen chamber at a pressure of 4–5 bar. The fuel injection frequency and hydrogen feed rate were controlled automatically in dependence on the electronic load. A sodium metaborate solution was discharged through a blow-down valve at a specified interval. Such a design considerably improved the characteristics of the response “fuel cell power—hydrogen feed rate—injection rate of new portions of the NaBH4 solution”, which provided a stable operation of LT PEM FC and simplified the utilization of sodium borate.
Note that all the flow reactors considered above make it necessary to use dispensers, which allow the control of the gas generation process but increase the weight and size of the hydrogen generators. A simpler instrumentation of the sodium borohydride hydrolysis process is typical of the batch (or autoclave) reactors (Figure 4). Their operation conditions allow catalysts to be used as powders instead of the macrostructured bed.
In [201], hydrogen for the functioning of a fuel cell battery with a power of 100 W was generated in an autoclave reactor, where a cobalt nitrate solution was dosed by a syringe pump on NaBH4 grains. However, this was accompanied by an unstable uncontrolled release of hydrogen with a pronounced hydrogen generation rate maximum, which was associated with a sharp growth of temperature in the reactor.
A constant rate of hydrogen generation without changes in the temperature at a rapid injection of water into spherical pills of a hydrogen-generating material was achieved in [52]. Along with sodium borohydride and a cobalt catalyst (Co2+/IR-120), the composite included a hydrophobic silicone rubber, which facilitated a uniform penetration of water into the pills, thus providing a stable release of hydrogen. The only drawback is that the dilution of sodium borohydride with an inert polymer decreases the GHSC value.
The authors of [202] also used a batch reactor with the CoOx/Ni foam catalyst to which a 15% NaBH4 solution was fed. In addition to parametric studies, the authors developed and successfully implemented a method for controlling time intervals between injections of fresh portions of the NaBH4 solution into reactor in dependence on the PEM FC load. The developed hydrogen generator provided a continuous production of considerable amounts of hydrogen (123 L H2/min) “on demand” to maintain the functioning of PEM FC with a power of 3 kW.
Thus, the main studies of hydrogen generators based on the catalytic hydrolysis of NaBH4 are aimed to develop reactors with the advanced designs depending on the type of cobalt catalyst (supported, macrostructured, bulk, doped, or ferromagnetic catalysts or their precursors with subsequent reduction in the reaction medium). Extended parametric studies are usually performed for revealing the effect of catalyst properties and the method of its loading in the reactor unit, the flow rate, and the concentration of NaBH4 solution as well as the time of its contact with the catalyst. This allows estimating the operational efficiency of a generator from different standpoints (HGR, conversion, start-up time, stability, durability, and integration with fuel cell). The results obtained made it possible to formulate the most acute problems, which require the development of the proposed sci-tech solutions and the creation of the new ones. Particularly, solutions are needed regarding the control of the temperature of reaction medium and its foaming in the reactor, stability of catalysts, purity of hydrogen under variable modes of generator functioning, weight and size characteristics of the generator, time of establishing a stable operation, and obtaining a rapid response to changes in the hydrogen demand of the fuel cell [30].

8. Conclusions

A significant increase in the number of publications in the last 20 years confirms that there is a great interest in the development of systems for hydrogen storage and generation based on sodium borohydride (NaBH4). This is related to its commercial availability, price, safety, and the possibility to generate H2 via hydrolysis at ambient temperatures. Studies on the development of such systems are carried out in several directions. Their main goal is a search for the methods that could increase the hydrogen capacity and control the gas generation rate. These characteristics, which are essential for the functioning of LT PEM FC, cannot be provided without the use of acids or catalysts.
The acid hydrolysis of NaBH4 is the known process for which special reactors are designed; particular attention is paid to purification of the evolved hydrogen in a highly exothermic process. By now, the highest GHSC values have been achieved with the use of aqueous solutions of strong mineral acids (HCl and H2SO4). Their high aggressiveness imposes considerable limitations on the materials suitable for constructing hydrogen generators, so the possibility to use weaker acids in the process is being explored.
Judging from the number of publications, the catalytic hydrolysis of NaBH4 should be considered as the process attracting the greatest attention. Catalysts based on noble metals (Rh, Ru, Pt) are known to exhibit the highest activity and stability; however, their high price initiates studies on the development of catalysts based on cobalt, which is the most active metal in the iron subgroup. Since NaBH4 is a reductant, a considerable part of works is devoted to the investigation of nanosized amorphous cobalt–boride catalysts that are formed directly in the reaction medium of NaBH4 hydrolysis from precursors (cobalt salts and oxides). Although there is quite a large body of experimental data concerning the composition and structure of the active components of such catalysts, it is still difficult to reliably describe the state of the active site of the Co-B catalyst, and further studies are performed to elucidate this issue.
The published results demonstrate that the application of Co-B catalytic systems in flow reactors for the hydrolysis of NaBH4 alkaline solutions is limited by their strong degradation. So, a different method of hydrogen production in the NaBH4 hydrolysis without an alkali was proposed for the Co-B catalysts. It implies the delivery of water to solid-phase or pelletized compositions of the hydride with a catalytic additive or feeding a solution of the Co-B catalyst precursor to solid NaBH4. This process can be carried out in batch (or autoclave) reactors. Dosing a limited amount of water provides high GHSC values as compared to the hydrolysis of NaBH4 alkaline solutions in a flow reactor. However, in this case, there are problems of self-heating and foaming the reaction medium caused by the exothermicity of the reaction and vigorous gas formation. For estimation of the potential of such a hydrogen generation process, it is necessary to perform further experimental and theoretical studies on the effect of heat and mass transfer on the kinetics of hydrogen generation and its purity when optimizing the design of such generators and varying the conditions of their operating.
Thus, the main lines of investigation include the development and improvement of cobalt-based catalysts and reactors both for the generation of H2 from NaBH4 solution and for the case of solid-phase compositions based on NaBH4. This will make it possible to optimize the weight and size characteristics of the reactor, as well as its heat exchange with the environment, to decrease the start-up time as well as the gas dynamic and hydraulic resistance in the catalyst bed, to solve the problems of catalyst washout and a loss in stability, to enhance the completeness of NaBH4 conversion and increase the purity of evolved hydrogen, and, which is most important, to find flexible solutions for controlling the hydrogen generator capacity in dependence on the power of LT PEM FC and the consumer’s goal. When creating the infrastructure for the application of high-purity hydrogen generation from NaBH4, particular attention should be paid to the convenience and safety of the process (the prevention of flooding of the alkaline reaction medium, decreasing the surface temperature of hydrogen generator, collecting the products and catalysts for their regeneration, etc.) as well as to environmental and economic aspects of the proposed technologies.

Author Contributions

Conceptualization, V.I.S.; investigation, V.I.S., A.M.O. and O.V.K.; writing—original draft preparation, A.M.O. and O.V.K.; writing—review and editing, V.I.S. and O.V.N.; visualization, O.V.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Higher Education of the Russian Federation within the governmental order for Boreskov Institute of Catalysis (project AAAA-A21-121011390006-0).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Susai, T.; Kawakami, A.; Hamada, A.; Miyake, Y.; Azegami, Y. Development of a 1 kW PEM fuel cell power source. Fuel Cells Bull. 2001, 3, 7–11. [Google Scholar] [CrossRef]
  2. Abdalla, A.M.; Hossain, S.; Nisfindy, O.B.; Azad, A.T.; Dawood, M.; Azad, A.K. Hydrogen production, storage, transportation and key challenges with applications: A review. Energy Convers. Manag. 2018, 165, 602–627. [Google Scholar] [CrossRef]
  3. Züttel, A. Hydrogen storage methods. Naturwissenschaften 2004, 91, 157–172. [Google Scholar] [CrossRef]
  4. Niaz, S.; Manzoor, T.; Pandith, A.H. Hydrogen storage: Materials, methods and perspectives. Renew. Sustain. Energy Rev. 2015, 50, 457–469. [Google Scholar] [CrossRef]
  5. Zacharia, R.; Rather, S.U. Review of solid state hydrogen storage methods adopting different kinds of novel materials. J. Nanomater. 2015, 2015, 1–18. [Google Scholar] [CrossRef] [Green Version]
  6. Durbin, D.; Malardier-Jugroot, C. Review of hydrogen storage techniques for on board vehicle applications. Int. J. Hydrogen Energy 2013, 38, 14595–14617. [Google Scholar] [CrossRef]
  7. Pukazhselvan, D.; Kumar, V.; Singh, S. High capacity hydrogen storage: Basic aspects, new developments and milestones. Nano Energy 2012, 1, 566–589. [Google Scholar] [CrossRef]
  8. Jiang, H.-L.; Singh, S.K.; Yan, J.-M.; Zhang, X.-B.; Xu, Q. Liquid-phase chemical hydrogen storage: Catalytic hydrogen generation under ambient conditions. ChemSusChem 2010, 3, 541–549. [Google Scholar] [CrossRef] [PubMed]
  9. Zheng, J.; Liu, X.; Xu, P.; Liu, P.; Zhao, Y.; Yang, J. Development of high pressure gaseous hydrogen storage technologies. Int. J. Hydrogen Energy 2012, 37, 1048–1057. [Google Scholar] [CrossRef]
  10. Ley, M.B.; Jepsen, L.H.; Lee, Y.-S.; Cho, Y.W.; Von Colbe, J.M.B.; Dornheim, M.; Rokni, M.; Jensen, J.O.; Sloth, M.; Filinchuk, Y.; et al. Complex hydrides for hydrogen storage—New perspectives. Mater. Today 2014, 17, 122–128. [Google Scholar] [CrossRef] [Green Version]
  11. Mohtadi, R.; Remhof, A.; Jena, P. Complex metal borohydrides: Multifunctional materials for energy storage and conversion. J. Phys. Condens. Matter 2016, 28, 353001. [Google Scholar] [CrossRef]
  12. Li, H.-W.; Yan, Y.; Orimo, S.-I.; Zuettel, A.; Jensen, C.M. Recent progress in metal borohydrides for hydrogen storage. Energies 2011, 4, 185–214. [Google Scholar] [CrossRef]
  13. Yadav, M.; Xu, Q. Liquid-phase chemical hydrogen storage materials. Energy Environ. Sci. 2012, 5, 9698–9725. [Google Scholar] [CrossRef]
  14. Kong, V. Development of hydrogen storage for fuel cell generators. i: Hydrogen generation using hydrolysis hydrides. Int. J. Hydrog. Energy 1999, 24, 665–675. [Google Scholar] [CrossRef]
  15. Fakiolu, E. A review of hydrogen storage systems based on boron and its compounds. Int. J. Hydrog. Energy 2004, 29, 1371–1376. [Google Scholar] [CrossRef]
  16. Zhang, J.; Fisher, T.; Gore, J.; Hazra, D.; Ramachandran, P. Heat of reaction measurements of sodium borohydride alcoholysis and hydrolysis. Int. J. Hydrog. Energy 2006, 31, 2292–2298. [Google Scholar] [CrossRef]
  17. Demirci, U.B.; Akdim, O.; Andrieux, J.; Hannauer, J.; Chamoun, R.; Miele, P. Sodium borohydride hydrolysis as hydrogen generator: Issues, state of the art and applicability upstream from a fuel cell. Fuel Cells 2010, 10, 335–350. [Google Scholar] [CrossRef] [Green Version]
  18. Santos, D.; Sequeira, C. Sodium borohydride as a fuel for the future. Renew. Sustain. Energy Rev. 2011, 15, 3980–4001. [Google Scholar] [CrossRef]
  19. Rivarolo, M.; Improta, O.; Magistri, L.; Panizza, M.; Barbucci, A. Thermo-economic analysis of a hydrogen production system by sodium borohydride (NaBH4). Int. J. Hydrog. Energy 2018, 43, 1606–1614. [Google Scholar] [CrossRef]
  20. Hirscher, M.; Yartys, V.A.; Baricco, M.; Von Colbe, J.B.; Blanchard, D.; Bowman, R.C.; Broom, D.P.; Buckley, C.E.; Chang, F.; Chen, P.; et al. Materials for hydrogen-based energy storage—Past, recent progress and future outlook. J. Alloy. Compd. 2020, 827, 153548. [Google Scholar] [CrossRef]
  21. Tedds, S.; Walton, A.; Broom, D.P.; Book, D. Characterisation of porous hydrogen storage materials: Carbons, zeolites, MOFs and PIMs. Faraday Discuss. 2011, 151, 75–94. [Google Scholar] [CrossRef] [PubMed]
  22. Aakko-Saksa, P.T.; Cook, C.; Kiviaho, J.; Repo, T. Liquid organic hydrogen carriers for transportation and storing of renewable energy—Review and discussion. J. Power Sources 2018, 396, 803–823. [Google Scholar] [CrossRef]
  23. Yartys, V.; Lototskyy, M.; Akiba, E.; Albert, R.; Antonov, V.; Ares, J.; Baricco, M.; Bourgeois, N.; Buckley, C.; Von Colbe, J.B.; et al. Magnesium based materials for hydrogen based energy storage: Past, present and future. Int. J. Hydrog. Energy 2019, 44, 7809–7859. [Google Scholar] [CrossRef]
  24. Milanese, C.; Garroni, S.; Gennari, F.; Marini, A.; Klassen, T.; Dornheim, M.; Pistidda, C. Solid state hydrogen storage in alanates and alanate-based compounds: A Review. Metals 2018, 8, 567. [Google Scholar] [CrossRef] [Green Version]
  25. Puszkiel, J.; Garroni, S.; Milanese, C.; Gennari, F.; Klassen, T.; Dornheim, M.; Pistidda, C. Tetrahydroborates: Development and potential as hydrogen storage medium. Inorganics 2017, 5, 74. [Google Scholar] [CrossRef] [Green Version]
  26. Garroni, S.; Santoru, A.; Cao, H.; Dornheim, M.; Klassen, T.; Milanese, C.; Gennari, F.; Pistidda, C. Recent progress and new perspectives on metal amide and imide systems for solid-state hydrogen storage. Energies 2018, 11, 1027. [Google Scholar] [CrossRef] [Green Version]
  27. Paskevicius, M.; Jepsen, L.H.; Schouwink, P.; Černý, R.; Ravnsbæk, D.B.; Filinchuk, Y.; Dornheim, M.; Besenbacher, F.; Jensen, T.R. Metal borohydrides and derivatives—Synthesis, structure and properties. Chem. Soc. Rev. 2017, 46, 1565–1634. [Google Scholar] [CrossRef]
  28. Liu, H.; Zhang, L.; Ma, H.; Lu, C.; Luo, H.; Wang, X.; Huang, X.; Lan, Z.; Guo, J. Aluminum hydride for solid-state hydrogen storage: Structure, synthesis, thermodynamics, kinetics, and regeneration. J. Energy Chem. 2021, 52, 428–440. [Google Scholar] [CrossRef]
  29. Demirci, U.B. Ammonia borane: An extensively studied, though not yet implemented, hydrogen carrier. Energies 2020, 13, 3071. [Google Scholar] [CrossRef]
  30. Demirci, U.B. About the technological readiness of the H2 generation by hydrolysis of B(−N)−H compounds. Energy Technol. 2018, 6, 470–486. [Google Scholar] [CrossRef] [Green Version]
  31. Kumar, D.; Muthukumar, K. An overview on activation of aluminium-water reaction for enhanced hydrogen production. J. Alloy. Compd. 2020, 835, 155189. [Google Scholar] [CrossRef]
  32. Xie, X.; Ni, C.; Wang, B.; Zhang, Y.; Zhao, X.; Liu, L.; Wang, B.; Du, W. Recent advances in hydrogen generation process via hydrolysis of Mg-based materials: A short review. J. Alloy. Compd. 2020, 816, 152634. [Google Scholar] [CrossRef]
  33. Tegel, M.; Schöne, S.; Kieback, B.; Röntzsch, L. An efficient hydrolysis of MgH2-based materials. Int. J. Hydrog. Energy 2017, 42, 2167–2176. [Google Scholar] [CrossRef]
  34. Agrawal, T.; Ajitkumar, R.; Prakash, R.; Nandan, G. Sodium silicide as a hydrogen source for portable energy devices: A review. Mater. Today Proc. 2018, 5, 3563–3570. [Google Scholar] [CrossRef]
  35. Demirci, U.B. The hydrogen cycle with the hydrolysis of sodium borohydride: A statistical approach for highlighting the scientific/technical issues to prioritize in the field. Int. J. Hydrog. Energy 2015, 40, 2673–2691. [Google Scholar] [CrossRef]
  36. Sahin, Ö.; Dolas, H.; Özdemir, M. The effect of various factors on the hydrogen generation by hydrolysis reaction of potassium borohydride. Int. J. Hydrog. Energy 2007, 32, 2330–2336. [Google Scholar] [CrossRef]
  37. Li, C.; Peng, P.; Zhou, D.; Wan, L. Research progress in LiBH4 for hydrogen storage: A review. Int. J. Hydrog. Energy 2011, 36, 14512–14526. [Google Scholar] [CrossRef]
  38. Moury, R.; Demirci, U.B. Hydrazine borane and hydrazinidoboranes as chemical hydrogen storage materials. Energies 2015, 8, 3118–3141. [Google Scholar] [CrossRef]
  39. Amendola, S. A safe, portable, hydrogen gas generator using aqueous borohydride solution and Ru catalyst. Int. J. Hydrog. Energy 2000, 25, 969–975. [Google Scholar] [CrossRef]
  40. Kojima, Y.; Suzuki, K.I.; Fukumoto, K.; Sasaki, M.; Yamamoto, T.; Kawai, Y.; Hayashi, H. Hydrogen generation using sodium borohydride solution and metal catalyst coated on metal oxide. Int. J. Hydrog. Energy 2002, 27, 1029–1034. [Google Scholar] [CrossRef]
  41. Zhang, J.; Zheng, Y.; Gore, J.P.; Mudawar, I.; Fisher, T.S. 1 kWe sodium borohydride hydrogen generation system. Part II: Reactor modeling. J. Power Sources 2007, 170, 150–159. [Google Scholar] [CrossRef]
  42. Richardson, B.S.; Birdwell, J.F.; Pin, F.G.; Jansen, J.F.; Lind, R.F. Sodium borohydride based hybrid power system. J. Power Sources 2005, 145, 21–29. [Google Scholar] [CrossRef]
  43. Liu, B.H.; Li, Z.P.; Suda, S. Solid sodium borohydride as a hydrogen source for fuel cells. J. Alloys Compd. 2009, 468, 493–498. [Google Scholar] [CrossRef]
  44. Oh, T.H.; Gang, B.G.; Kim, H.; Kwon, S. Sodium borohydride hydrogen generator using Co-P/Ni foam catalysts for 200 W proton exchange membrane fuel cell system. Energy 2015, 90, 1163–1170. [Google Scholar] [CrossRef]
  45. Gang, B.G.; Jung, W.; Kwon, S. Transient behavior of proton exchange membrane fuel cells over a cobalt-phosphorous/nickel foam catalyst with sodium borohydride. Int. J. Hydrog. Energy 2016, 41, 524–533. [Google Scholar] [CrossRef]
  46. Kim, T. Fully-integrated micro PEM fuel cell system with NaBH4 hydrogen generator. Int. J. Hydrog. Energy 2012, 37, 2440–2446. [Google Scholar] [CrossRef]
  47. Akdim, O.; Demirci, U.B.; Miele, P. More reactive cobalt chloride in the hydrolysis of sodium borohydride. Int. J. Hydrog. Energy 2009, 34, 9444–9449. [Google Scholar] [CrossRef]
  48. Nunes, H.X.; Ferreira, M.J.F.; Rangel, C.M.; Pinto, A.M.F.R. Hydrogen generation and storage by aqueous sodium borohydride (NaBH4) hydrolysis for small portable fuel cells (H2—PEMFC). Int. J. Hydrog. Energy 2016, 41, 15426–15432. [Google Scholar] [CrossRef]
  49. Sousa, T.; Fernandes, V.R.; Pinto, P.J.R.; Slavkov, Y.; Bosukov, L.; Rangel, C.M. A sodium borohydride hydrogen generation reactor for stationary applications: Experimental and reactor simulation studies. Chem. Eng. Sci. 2012, 84, 70–79. [Google Scholar] [CrossRef] [Green Version]
  50. Galli, S.; De Francesco, M.; Monteleone, G.; Oronzio, R.; Pozio, A. Development of a compact hydrogen generator from sodium borohydride. Int. J. Hydrog. Energy 2010, 35, 7344–7349. [Google Scholar] [CrossRef]
  51. Gislon, P.; Prosini, P.P. Devices for producing hydrogen via NaBH4 and LiH hydrolysis. Int. J. Hydrog. Energy 2011, 36, 240–246. [Google Scholar] [CrossRef]
  52. Hsueh, C.L.; Liu, C.H.; Chen, B.H.; Lee, M.S.; Chen, C.Y.; Lu, Y.W.; Tsau, F.; Ku, J.R. A novel design of solid-state NaBH4 composite as a hydrogen source for 2 W PEMFC applications. J. Power Sources 2011, 196, 3530–3538. [Google Scholar] [CrossRef]
  53. Simagina, V.I.; Ozerova, A.M.; Komova, O.V.; Odegova, G.V.; Kellerman, D.G.; Fursenko, R.V.; Odintsov, E.S.; Netskina, O.V. Cobalt boride catalysts for small-scale energy application. Catal. Today 2015, 242, 221–229. [Google Scholar] [CrossRef]
  54. Netskina, O.V.; Fursenko, R.V.; Komova, O.V.; Odintsov, E.S.; Simagina, V.I. NaBH4 generator integrated with energy conversion device based on hydrogen combustion. J. Power Sources 2015, 273, 278–281. [Google Scholar] [CrossRef]
  55. Komova, O.V.; Netskina, O.V.; Ozerova, A.M.; Odegova, G.V.; Arzumanov, S.S.; Simagina, V.I. The effect of storage-induced changes in ammonia borane on hydrogen release during its low-temperature thermolysis. Inorganics 2019, 7, 96. [Google Scholar] [CrossRef] [Green Version]
  56. Damjanović, L.; Bennici, S.; Auroux, A. A direct measurement of the heat evolved during the sodium and potassium borohydride catalytic hydrolysis. J. Power Sources 2010, 195, 3284–3292. [Google Scholar] [CrossRef]
  57. Marrero-Alfonso, E.Y.; Beaird, A.M.; Davis, T.A.; Matthews, M.A. Hydrogen generation from chemical hydrides. Ind. Eng. Chem. Res. 2009, 48, 3703–3712. [Google Scholar] [CrossRef]
  58. Marrero-Alfonso, E.Y.; Gray, J.R.; Davis, T.A.; Matthewa, M.A. Minimizing water utilization in hydrolysis of sodium borohydride: The role of sodium metaborate hydrates. Int. J. Hydrog. Energy 2007, 32, 4723–4730. [Google Scholar] [CrossRef]
  59. Minkina, V.G.; Shabunya, S.I.; Kalinin, V.I.; Martynenko, V.V.; Smirnova, A.L. Long-term stability of sodium borohydrides for hydrogen generation. Int. J. Hydrog. Energy 2008, 33, 5629–5635. [Google Scholar] [CrossRef]
  60. Minkina, V.G.; Shabunya, S.I.; Kalinin, V.I.; Martynenko, V.V.; Smirnova, A.L. Stability of alkaline aqueous solutions of sodium borohydride. Int. J. Hydrog. Energy 2012, 37, 3313–3318. [Google Scholar] [CrossRef]
  61. Schlesinger, H.I.; Brown, H.C.; Finholt, A.E.; Gilbreath, J.R.; Hoekstra, H.R.; Hyde, E.K. Sodium borohydride, its hydrolysis and its use as a reducing agent and in the generation of hydrogen. J. Am. Chem. Soc. 1953, 75, 215–219. [Google Scholar] [CrossRef]
  62. DOE Technical Targets for Onboard Hydrogen Storage for Light-Duty Vehicles | Department of Energy. Available online: https://www.energy.gov/eere/fuelcells/doe-technical-targets-onboard-hydrogen-storage-light-duty-vehicles (accessed on 6 June 2020).
  63. Akdim, O.; Demirci, U.; Miele, P. Acetic acid, a relatively green single-use catalyst for hydrogen generation from sodium borohydride. Int. J. Hydrog. Energy 2009, 34, 7231–7238. [Google Scholar] [CrossRef]
  64. Kim, H.J.; Shin, K.-J.; Kim, H.-J.; Han, M.; Kim, H.; Shul, Y.-G.; Jung, K.T. Hydrogen generation from aqueous acid-catalyzed hydrolysis of sodium borohydride. Int. J. Hydrog. Energy 2010, 35, 12239–12245. [Google Scholar] [CrossRef]
  65. Murugesan, S.; Subramanian, V.R. Effects of acid accelerators on hydrogen generation from solid sodium borohydride using small scale devices. J. Power Sources 2009, 187, 216–223. [Google Scholar] [CrossRef]
  66. Kwon, S.-M.; Kim, M.J.; Kang, S.; Kim, T. Development of a high-storage-density hydrogen generator using solid-state NaBH4 as a hydrogen source for unmanned aerial vehicles. Appl. Energy 2019, 251, 113331. [Google Scholar] [CrossRef]
  67. Lee, C.J.; Kim, T. Hydrogen supply system employing direct decomposition of solid-state NaBH4. Int. J. Hydrog. Energy 2015, 40, 2274–2282. [Google Scholar] [CrossRef]
  68. Xia, Z.; Shen, Y.; Wang, Y.; Poh, C.K.; Lin, J. Development of a portable hydrogen generator with differential pressure-driven control. Int. J. Hydrog. Energy 2014, 39, 16187–16194. [Google Scholar] [CrossRef]
  69. Shavi, R.; Jadhav, A.H.; Lee, K.; Gil Seo, J. Sulfonated nanolayers of H+-montmorillonite as an efficient acidic catalyst for hydrogen generation from hydrolysis of sodium borohydride. J. Nanosci. Nanotechnol. 2016, 16, 10980–10985. [Google Scholar] [CrossRef]
  70. Manna, J.; Roy, B.; Sharma, P. Efficient hydrogen generation from sodium borohydride hydrolysis using silica sulfuric acid catalyst. J. Power Sources 2015, 275, 727–733. [Google Scholar] [CrossRef]
  71. Chen, J.; Wang, C.; Wu, Y.; Fu, H.; Zheng, J.; Li, X. Hydrogen generation from reactions of hydrides with hydrated solids in the solid state. RSC Adv. 2016, 6, 36863–36869. [Google Scholar] [CrossRef]
  72. Wang, M.; Ouyang, L.; Liu, J.; Wang, H.; Zhu, M. Hydrogen generation from sodium borohydride hydrolysis accelerated by zinc chloride without catalyst: A kinetic study. J. Alloy. Compd. 2017, 717, 48–54. [Google Scholar] [CrossRef]
  73. Balbay, A.; Saka, C. Effect of phosphoric acid addition on the hydrogen production from hydrolysis of NaBH4 with Cu based catalyst. Energy Sources Part A Recover. Util. Environ. Eff. 2018, 40, 794–804. [Google Scholar] [CrossRef]
  74. Balbay, A.; Saka, C. Effect of phosphoric acid and acetic acid addition on the hydrogen evolution using Ni based catalyst prepared in ethanol, methanol, and water solvents. Energy Sources Part A Recover. Util. Environ. Eff. 2018, 40, 2442–2450. [Google Scholar] [CrossRef]
  75. Saka, C.; Balbay, A. Efficient hydrogen production with controlled hydrochloric acid addition using Ni-based catalyst synthesized in ethanol solvent. Energy Sources Part A Recover. Util. Environ. Eff. 2020, 42, 1193–1205. [Google Scholar] [CrossRef]
  76. Abdul-Majeed, W.S.; Arslan, M.T.; Zimmerman, W.B. Application of acidic accelerator for production of pure hydrogen from NaBH4. Int. J. Ind. Chem. 2014, 5, 1–9. [Google Scholar] [CrossRef] [Green Version]
  77. Duke, B.J.; Gilbert, J.R.; Read, I.A. Preparation and purification of diborane. J. Chem. Soc. 1964, 540–541. [Google Scholar]
  78. Netskina, O.V.; Filippov, T.N.; Komova, O.V.; Simagina, V.I. Hydrogen generation by both acidic and catalytic hydrolysis of sodium borohydride. Catal. Sustain. Energy 2018, 5, 41–48. [Google Scholar] [CrossRef]
  79. Carenco, S.; Portehault, D.; Boissière, C.; Mézailles, N.; Sanchez, C. Nanoscaled metal borides and phosphides: Recent developments and perspectives. Chem. Rev. 2013, 113, 7981–8065. [Google Scholar] [CrossRef] [PubMed]
  80. Liu, B.; Li, Z. A review: Hydrogen generation from borohydride hydrolysis reaction. J. Power Sources 2009, 187, 527–534. [Google Scholar] [CrossRef]
  81. Muir, S.S.; Yao, X. Progress in sodium borohydride as a hydrogen storage material: Development of hydrolysis catalysts and reaction systems. Int. J. Hydrog. Energy 2011, 36, 5983–5997. [Google Scholar] [CrossRef]
  82. Brack, P.; Dann, S.E.; Wijayantha, K.G.U. Heterogeneous and homogenous catalysts for hydrogen generation by hydrolysis of aqueous sodium borohydride (NaBH4) solutions. Energy Sci. Eng. 2015, 3, 174–188. [Google Scholar] [CrossRef] [Green Version]
  83. Liu, B.H.; Li, Z.P.; Suda, S. Thermal properties of alkaline sodium borohydride solutions. Thermochim. Acta 2008, 471, 103–105. [Google Scholar] [CrossRef]
  84. Demirci, U.B.; Miele, P. Reaction mechanisms of the hydrolysis of sodium borohydride: A discussion focusing on cobalt-based catalysts. Comptes Rendus Chim. 2014, 17, 707–716. [Google Scholar] [CrossRef]
  85. Retnamma, R.; Novais, A.Q.; Rangel, C. Kinetics of hydrolysis of sodium borohydride for hydrogen production in fuel cell applications: A review. Int. J. Hydrog. Energy 2011, 36, 9772–9790. [Google Scholar] [CrossRef]
  86. Patel, N.; Miotello, A. Progress in Co–B related catalyst for hydrogen production by hydrolysis of boron-hydrides: A review and the perspectives to substitute noble metals. Int. J. Hydrog. Energy 2015, 40, 1429–1464. [Google Scholar] [CrossRef]
  87. Holbrook, K.A.; Twist, P.J. Hydrolysis of the borohydride ion catalyzed by metal–boron alloys. J. Chem. Soc. A 1971, 890–894. [Google Scholar] [CrossRef]
  88. Guella, G.; Zanchetta, C.; Patton, B.; Miotello, A. New insights on the mechanism of palladium-catalyzed hydrolysis of sodium borohydride from 11B NMR measurements. J. Phys. Chem. B 2006, 110, 17024–17033. [Google Scholar] [CrossRef] [PubMed]
  89. Walter, J.C.; Zurawski, A.; Montgomery, D.; Thornburg, M.; Revankar, S. Sodium borohydride hydrolysis kinetics comparison for nickel, cobalt, and ruthenium boride catalysts. J. Power Sources 2008, 179, 335–339. [Google Scholar] [CrossRef]
  90. Penaalonso, R.; Sicurelli, A.; Callone, E.; Carturan, G.; Raj, R. A picoscale catalyst for hydrogen generation from NaBH4 for fuel cells. J. Power Sources 2007, 165, 315–323. [Google Scholar] [CrossRef]
  91. Sun, H.; Meng, J.; Jiao, L.; Cheng, F.; Chen, J. A review of transition-metal boride/phosphide-based materials for catalytic hydrogen generation from hydrolysis of boron-hydrides. Inorg. Chem. Front. 2018, 5, 760–772. [Google Scholar] [CrossRef]
  92. Andrieux, J.; Demirci, U.; Miele, P. Langmuir–Hinshelwood kinetic model to capture the cobalt nanoparticles-catalyzed hydrolysis of sodium borohydride over a wide temperature range. Catal. Today 2011, 170, 13–19. [Google Scholar] [CrossRef]
  93. Netskina, O.V.; Kochubey, D.I.; Prosvirin, I.P.; Kellerman, D.G.; Simagina, V.I.; Komova, O.V. Role of the electronic state of rhodium in sodium borohydride hydrolysis. J. Mol. Catal. A Chem. 2014, 390, 125–132. [Google Scholar] [CrossRef]
  94. Liu, Z.; Guo, B.; Chan, S.H.; Tang, E.H.; Hong, L. Pt and Ru dispersed on LiCoO2 for hydrogen generation from sodium borohydride solutions. J. Power Sources 2008, 176, 306–311. [Google Scholar] [CrossRef]
  95. Özkar, S.; Zahmakıran, M. Hydrogen generation from hydrolysis of sodium borohydride using Ru(0) nanoclusters as catalyst. J. Alloys Compd. 2005, 404–406, 728–731. [Google Scholar] [CrossRef]
  96. Simagina, V.I.; Storozhenko, P.A.; Netskina, O.V.; Komova, O.V.; Odegova, G.V.; Larichev, Y.V.; Ishchenko, A.V.; Ozerova, A.M. Development of catalysts for hydrogen generation from hydride compounds. Catal. Today 2008, 138, 253–259. [Google Scholar] [CrossRef]
  97. Simagina, V.I.; Storozhenko, P.A.; Netskina, O.V.; Komova, O.V.; Odegova, G.V.; Samoilenko, T.Y.; Gentsler, A.G. Effect of the nature of the active component and support on the activity of catalysts for the hydrolysis of sodium borohydride. Kinet. Catal. 2007, 48, 168–175. [Google Scholar] [CrossRef]
  98. Rodrigues, T.S.; Zhao, M.; Yang, T.; Gilroy, K.D.; Da Silva, A.G.M.; Camargo, P.H.C.; Xia, Y. Synthesis of colloidal metal nanocrystals: A comprehensive review on the reductants. Chem. A Eur. J. 2018, 24, 16944–16963. [Google Scholar] [CrossRef]
  99. Ewers, T.D.; Sra, A.K.; Norris, B.C.; Cable, R.E.; Cheng, C.-H.; Shantz, D.F.; Schaak, R.E. Spontaneous hierarchical assembly of rhodium nanoparticles into spherical aggregates and superlattices. Chem. Mater. 2005, 17, 514–520. [Google Scholar] [CrossRef]
  100. Bozkurt, G.; Özer, A.; Yurtcan, A.B. Development of effective catalysts for hydrogen generation from sodium borohydride: Ru, Pt, Pd nanoparticles supported on Co3O4. Energy 2019, 180, 702–713. [Google Scholar] [CrossRef]
  101. Lee, D.-S.; Liu, T.-K. Characteristics of nanosized ruthenium metal by chemical reduction method. J. Non-Crystalline Solids 2002, 311, 323–327. [Google Scholar] [CrossRef]
  102. Tan, Y.; Yu, X. Chemical regeneration of hydrogen storage materials. RSC Adv. 2013, 3, 23879–23894. [Google Scholar] [CrossRef]
  103. Ouyang, L.; Zhong, H.; Li, H.-W.; Zhu, M. A Recycling hydrogen supply system of NaBH4 based on a facile regeneration process: A review. Inorganics 2018, 6, 10. [Google Scholar] [CrossRef] [Green Version]
  104. Ou, T.; Giuliano, A.; Panizza, M.; Barbucci, A.; Cerisola, G. Thermochemical recycling of hydrolyzed NaBH4. Part I: In-situ and ex-situ evaluations. Int. J. Hydrog. Energy 2013, 38, 15269–15274. [Google Scholar] [CrossRef]
  105. Ou, T.; Panizza, M.; Barbucci, A. Thermochemical recycling of hydrolyzed NaBH4. Part II: Systematical study of parameters dependencies. Int. J. Hydrog. Energy 2013, 38, 15940–15945. [Google Scholar] [CrossRef]
  106. Liu, B.; Li, Z.; Morigasaki, N.; Suda, S. Kinetic characteristics of sodium borohydride formation when sodium meta-borate reacts with magnesium and hydrogen. Int. J. Hydrog. Energy 2008, 33, 1323–1328. [Google Scholar] [CrossRef]
  107. Kojima, Y.; Haga, T. Recycling process of sodium metaborate to sodium borohydride. Int. J. Hydrog. Energy 2003, 28, 989–993. [Google Scholar] [CrossRef]
  108. Liu, B.H.; Li, Z.P.; Suda, S. Improving MgH2 formation kinetics and its effect on NaBH4 synthesis. J. Alloy. Compd. 2009, 474, 321–325. [Google Scholar] [CrossRef]
  109. Eom, K.; Cho, E.; Kim, M.; Oh, S.; Nam, S.-W.; Kwon, H. Thermochemical production of sodium borohydride from sodium metaborate in a scaled-up reactor. Int. J. Hydrog. Energy 2013, 38, 2804–2809. [Google Scholar] [CrossRef]
  110. Hsueh, C.-L.; Liu, C.-H.; Chen, B.-H.; Chen, C.-Y.; Kuo, Y.-C.; Hwang, K.-J.; Ku, J.-R. Regeneration of spent-NaBH4 back to NaBH4 by using high-energy ball milling. Int. J. Hydrog. Energy 2009, 34, 1717–1725. [Google Scholar] [CrossRef]
  111. Çakanyıldırım, Ç.; Gürü, M. Processing of NaBH4 from NaBO2 with MgH2 by ball milling and usage as hydrogen carrier. Renew. Energy 2010, 35, 1895–1899. [Google Scholar] [CrossRef]
  112. Lang, C.; Jia, Y.; Liu, J.; Wang, H.; Ouyang, L.; Zhu, M.; Yao, X. NaBH4 regeneration from NaBO2 by high-energy ball milling and its plausible mechanism. Int. J. Hydrog. Energy 2017, 42, 13127–13135. [Google Scholar] [CrossRef]
  113. Demirci, U.B.; Miele, P. Cobalt in NaBH4 hydrolysis. Phys. Chem. Chem. Phys. 2010, 12, 14651–14665. [Google Scholar] [CrossRef]
  114. Simagina, V.I.; Komova, O.V.; Netskina, O.V. Nanosized Cobalt Catalysts for Hydrogen Storage Systems Based on Ammonia Borane and Sodium Borohydride. In Metal Nanopowders; Gromov, A., Teipel, U., Eds.; Wiley: Hoboken, NJ, USA, 2014; pp. 199–226. [Google Scholar]
  115. Demirci, U.B.; Akdim, O.; Hannauer, J.; Chamoun, R.; Miele, P. Cobalt, a reactive metal in releasing hydrogen from sodium borohydride by hydrolysis: A short review and a research perspective. Sci. China Ser. B Chem. 2010, 53, 1870–1879. [Google Scholar] [CrossRef]
  116. Dragieva, I.; Mehandjiev, D.; Lefterova, E.; Stoycheva, M.; Stoynov, Z. Mechanism of borohydride reduction process and properties of obtained nanosized ferromagnetic particles. J. Magn. Magn. Mater. 1995, 140, 455–456. [Google Scholar] [CrossRef]
  117. Lefterova, E.; Dragieva, I.; Tchanev, V.; Mehandjiev, D.; Mikhov, M. Crystallographic phases in nanosized ferromagnetic particles obtained by two different methods. J. Magn. Magn. Mater. 1995, 140, 457–458. [Google Scholar] [CrossRef]
  118. Chen, Y. Chemical preparation and characterization of metal–metalloid ultrafine amorphous alloy particles. Catal. Today 1998, 44, 3–16. [Google Scholar] [CrossRef]
  119. Akdim, O.; Demirci, U.; Muller, D.; Miele, P. Cobalt (II) salts, performing materials for generating hydrogen from sodium borohydride. Int. J. Hydrog. Energy 2009, 34, 2631–2637. [Google Scholar] [CrossRef]
  120. Cavaliere, S.; Hannauer, J.; Demirci, U.B.; Akdim, O.; Miele, P. Ex situ characterization of N2H4-, NaBH4- and NH3BH3-reduced cobalt catalysts used in NaBH4 hydrolysis. Catal. Today 2011, 170, 3–12. [Google Scholar] [CrossRef]
  121. Liu, B.H.; Li, Z.P.; Suda, S. Nickel- and cobalt-based catalysts for hydrogen generation by hydrolysis of borohydride. J. Alloy. Compd. 2006, 415, 288–293. [Google Scholar] [CrossRef]
  122. Wu, C.; Bai, Y.; Wu, F.; Yi, B.-L.; Zhang, H.-M. Highly active cobalt-based catalysts in situ prepared from CoX2 (X=Cl, NO3) and used for promoting hydrogen generation from NaBH4 solution. Int. J. Hydrog. Energy 2010, 35, 2675–2679. [Google Scholar] [CrossRef]
  123. Ozerova, A.M.; Bulavchenko, O.A.; Komova, O.V.; Netskina, O.V.; Zaikovskii, V.I.; Odegova, G.V.; Simagina, V.I. Cobalt boride catalysts for hydrogen storage systems based on NH3BH3 and NaBH4. Kinet. Catal. 2012, 53, 511–520. [Google Scholar] [CrossRef]
  124. Yan, J.-M.; Zhang, X.-B.; Shioyama, H.; Xu, Q. Room temperature hydrolytic dehydrogenation of ammonia borane catalyzed by Co nanoparticles. J. Power Sources 2010, 195, 1091–1094. [Google Scholar] [CrossRef]
  125. Jeong, S.; Cho, E.; Nam, S.; Oh, I.; Jung, U.; Kim, S. Effect of preparation method on Co–B catalytic activity for hydrogen generation from alkali NaBH4 solution. Int. J. Hydrog. Energy 2007, 32, 1749–1754. [Google Scholar] [CrossRef]
  126. Ingersoll, J.; Mani, N.; Thenmozhiyal, J.; Muthaiah, A. Catalytic hydrolysis of sodium borohydride by a novel nickel–cobalt–boride catalyst. J. Power Sources 2007, 173, 450–457. [Google Scholar] [CrossRef]
  127. Liu, B.H.; Li, Q. A highly active Co-B catalyst for hydrogen generation from sodium borohydride hydrolysis. Int. J. Hydrog. Energy 2008, 33, 7385–7391. [Google Scholar] [CrossRef]
  128. Shen, X.; Dai, M.; Gao, M.; Zhao, B.; Ding, W. Solvent effects in the synthesis of CoB catalysts on hydrogen generation from hydrolysis of sodium borohydride. Chin. J. Catal. 2013, 34, 979–985. [Google Scholar] [CrossRef]
  129. Glavee, G.N.; Klabunde, K.J.; Sorensen, C.M.; Hadjipanayis, G.C. Borohydride reduction of cobalt ions in water. Chemistry leading to nanoscale metal, boride, or borate particles. Langmuir 1993, 9, 162–169. [Google Scholar] [CrossRef]
  130. Glavee, G.N.; Klabunde, K.J.; Sorensen, C.M.; Hadjapanayis, G.C. Borohydride reductions of metal ions. A new understanding of the chemistry leading to nanoscale particles of metals, borides, and metal borates. Langmuir 1992, 8, 771–773. [Google Scholar] [CrossRef]
  131. Netskina, O.V.; Komova, O.V.; Simagina, V.I.; Odegova, G.V.; Prosvirin, I.P.; Bulavchenko, O.A. Aqueous-alkaline NaBH4 solution: The influence of storage duration of solutions on reduction and activity of cobalt catalysts. Renew. Energy 2016, 99, 1073–1081. [Google Scholar] [CrossRef]
  132. Simagina, V.I.; Komova, O.V.; Ozerova, A.M.; Netskina, O.V.; Odegova, G.V.; Kellerman, D.G.; Bulavchenko, O.A.; Ishchenko, A.V. Cobalt oxide catalyst for hydrolysis of sodium borohydride and ammonia borane. Appl. Catal. A Gen. 2011, 394, 86–92. [Google Scholar] [CrossRef]
  133. Krishnan, P.; Advani, S.G.; Prasad, A.K. Cobalt oxides as Co2B catalyst precursors for the hydrolysis of sodium borohydride solutions to generate hydrogen for PEM fuel cells. Int. J. Hydrog. Energy 2008, 33, 7095–7102. [Google Scholar] [CrossRef]
  134. Netskina, O.V.; Komova, O.V.; Mukha, S.A.; Simagina, V.I. Aqueous-alkaline NaBH4 solutions: The influence of hydride decomposition on catalytic properties of Co3O4. Catal. Commun. 2016, 85, 9–12. [Google Scholar] [CrossRef]
  135. Demirci, U.B.; Miele, P. Cobalt-based catalysts for the hydrolysis of NaBH4 and NH3BH3. Phys. Chem. Chem. Phys. 2014, 16, 6872–6885. [Google Scholar] [CrossRef] [PubMed]
  136. Petit, C.; Pileni, M. Nanosize cobalt boride particles: Control of the size and properties. J. Magn. Magn. Mater. 1997, 166, 82–90. [Google Scholar] [CrossRef]
  137. Saida, J.; Inoue, A.; Masumoto, T. The effect of reaction condition on composition and properties of ultrafine amorphous powders in (Fe, Co, Ni)-B systems prepared by chemical reduction. Met. Mater. Trans. A 1991, 22, 2125–2132. [Google Scholar] [CrossRef]
  138. Lu, J.; Dreisinger, D.; Cooper, W. Cobalt precipitation by reduction with sodium borohydride. Hydrometall. 1997, 45, 305–322. [Google Scholar] [CrossRef] [Green Version]
  139. Garron, A.; Swierczynski, D.; Bennici, S.; Auroux, A. New insights into the mechanism of H2 generation through NaBH4 hydrolysis on Co-based nanocatalysts studied by differential reaction calorimetry. Int. J. Hydrog. Energy 2009, 34, 1185–1199. [Google Scholar] [CrossRef]
  140. Corrias, A.; Ennas, G.; Musinu, A.; Marongiu, G.; Paschina, G. Amorphous transition metal-boron ultrafine particles prepared by chemical methods. Chem. Mater. 1993, 5, 1722–1726. [Google Scholar] [CrossRef]
  141. Petit, C.; Pileni, M.P.A. Cobalt nanosized particles organized in a 2D superlattice: Synthesis, characterization, and magnetic properties. J. Phys. Chem. B 1999, 103, 1805–1810. [Google Scholar] [CrossRef]
  142. Metin, O.; Özkar, S. Hydrogen generation from the hydrolysis of ammonia-borane and sodium borohydride using water-soluble polymer-stabilized cobalt(0) nanoclusters catalyst. Energy Fuels 2009, 23, 3517–3526. [Google Scholar] [CrossRef]
  143. Patel, N.; Guella, G.; Kale, A.; Miotello, A.; Patton, B.; Zanchetta, C.; Mirenghi, L.; Rotolo, P. Thin films of Co–B prepared by pulsed laser deposition as efficient catalysts in hydrogen producing reactions. Appl. Catal. A Gen. 2007, 323, 18–24. [Google Scholar] [CrossRef]
  144. Dávila-Ibáñez, A.B.; Legido-Soto, J.L.; Rivas, J.; Salgueirino, V. Amorphous tunable-size Co–B magnetic nanoparticles from the cobalt-catalyzed NaBH4 hydrolysis. Phys. Chem. Chem. Phys. 2011, 13, 20146–20154. [Google Scholar] [CrossRef] [PubMed]
  145. Wu, Z.; Ge, S. Facile synthesis of a Co–B nanoparticle catalyst for efficient hydrogen generation via borohydride hydrolysis. Catal. Commun. 2011, 13, 40–43. [Google Scholar] [CrossRef]
  146. Wu, C.; Wu, F.; Bai, Y.; Yi, B.; Zhang, H. Cobalt boride catalysts for hydrogen generation from alkaline NaBH4 solution. Mater. Lett. 2005, 59, 1748–1751. [Google Scholar] [CrossRef]
  147. Netskina, O.V.; Kochubey, D.I.; Prosvirin, I.P.; Malykhin, S.E.; Komova, O.V.; Kanazhevskiy, V.V.; Chukalkin, Y.G.; Bobrovskii, V.I.; Kellerman, D.G.; Ishchenko, A.V.; et al. Cobalt-boron catalyst for NaBH4 hydrolysis: The state of the active component forming from cobalt chloride in a reaction medium. Mol. Catal. 2017, 441, 100–108. [Google Scholar] [CrossRef]
  148. Netskina, O.V.; Kellerman, D.G.; Ishchenko, A.V.; Komova, O.V.; Simagina, V.I. Amorphous ferromagnetic cobalt-boron composition reduced by sodium borohydride: Phase transformation at heat-treatment and its influence on the catalytic properties. Colloids Surf A Physicochem. Eng. Asp. 2018, 537, 485–494. [Google Scholar] [CrossRef]
  149. Arzac, G.M.; Rojas, T.C.; Fernández, A. Boron compounds as stabilizers of a complex microstructure in a Co-B-based catalyst for NaBH4 hydrolysis. ChemCatChem 2011, 3, 1305–1313. [Google Scholar] [CrossRef] [Green Version]
  150. Ozerova, A.M.; Simagina, V.I.; Komova, O.V.; Netskina, O.V.; Odegova, G.V.; Bulavchenko, O.A.; Rudina, N.A. Cobalt borate catalysts for hydrogen production via hydrolysis of sodium borohydride. J. Alloy. Compd. 2012, 513, 266–272. [Google Scholar] [CrossRef]
  151. Xue, T.; Lee, J.-M. Capacitive behavior of mesoporous Co(OH)2 nanowires. J. Power Sources 2014, 245, 194–202. [Google Scholar] [CrossRef]
  152. Mavel, G.; Escard, J.; Costa, P.; Castaing, J. ESCA surface study of metal borides. Surf. Sci. 1973, 35, 109–116. [Google Scholar] [CrossRef]
  153. Geng, J.; Jefferson, D.A.; Johnson, B.F.G. Exploring the structural complexities of metal-metalloid nanoparticles: The case of NiB as catalyst. Chem. A Eur. J. 2008, 15, 1134–1143. [Google Scholar] [CrossRef] [PubMed]
  154. Yiping, L.; Hadjipanayis, G.; Sorensen, C.; Klabunde, K. Magnetic and structural properties of ultrafine Co-B particles. J. Magn. Magn. Mater. 1989, 79, 321–326. [Google Scholar] [CrossRef]
  155. Hasegawa, R.; Ray, R. Magnetization of glassy Co-B alloys. J. Appl. Phys. 1979, 50, 1586–1588. [Google Scholar] [CrossRef]
  156. Zhurakovskii, E.A.; Shashkina, T.B.; Kotlyar, V.I. X-ray absorption spectra of cobalt in ferromagnetic borides. Russ. Phys. J. 1970, 13, 14–17. [Google Scholar] [CrossRef]
  157. Mizutani, U.; Hasegawa, M.; Fukamichi, K.; Hattori, Y.; Yamada, Y.; Tanaka, H.; Takayama, S. Magnetic, electronic, and electron-transport properties of amorphous (Co0.85B0.15)100−xXx(X = B, Al, Si, and V) alloys. Phys. Rev. B 1993, 47, 2678–2688. [Google Scholar] [CrossRef] [PubMed]
  158. Tanaka, H.; Takayama, S.; Hasegawa, M.; Fukunaga, T.; Mizutani, U.; Fujita, A.; Fukamichi, K. Electronic structure and magnetism of amorphousCo1−xBx alloys. Phys. Rev. B 1993, 47, 2671–2677. [Google Scholar] [CrossRef] [PubMed]
  159. Komova, O.V.; Simagina, V.I.; Netskina, O.V.; Kellerman, D.G.; Ishchenko, A.V.; Rudina, N.A. LiCoO2-based catalysts for generation of hydrogen gas from sodium borohydride solutions. Catal. Today 2008, 138, 260–265. [Google Scholar] [CrossRef]
  160. Maybury, P.C.; Mitchell, R.W.; Hawthorne, M.F. Hydrogen adducts of cobalt and nickel boride. J. Chem. Soc. Chem. Commun. 1974, 534–535. [Google Scholar] [CrossRef]
  161. Netskina, O.V.; Ozerova, A.M.; Komova, O.V.; Kochubey, D.I.; Kanazhevskiy, V.V.; Ishchenko, A.V.; Simagina, V.I. The Effect of heat-treatment temperature of cobalt–boron catalysts on their activity in sodium borohydride hydrolysis. Top. Catal. 2016, 59, 1431–1437. [Google Scholar] [CrossRef]
  162. Beachell, H.C.; Haugh, J.F. The pyrolysis of decaborane. J. Am. Chem. Soc. 1958, 80, 2939–2942. [Google Scholar] [CrossRef]
  163. Paladini, M.; Godinho, V.; Arzac, G.M.; Haro, M.D.C.J.D.; Beltrán, A.M.; Fernández-Camacho, A. Tailor-made preparation of Co–C, Co–B, and Co catalytic thin films using magnetron sputtering: Insights into structure–composition and activation effects for catalyzed NaBH4 hydrolysis. RSC Adv. 2016, 6, 108611–108620. [Google Scholar] [CrossRef] [Green Version]
  164. Wang, Y.; Li, T.; Bai, S.; Qi, K.; Cao, Z.; Zhang, K.; Wu, S.; Wang, D. Catalytic hydrolysis of sodium borohydride via nanostructured cobalt–boron catalysts. Int. J. Hydrog. Energy 2016, 41, 276–284. [Google Scholar] [CrossRef]
  165. Kadrekar, R.; Patel, N.; Arya, A. Understanding the role of boron and stoichiometric ratio in the catalytic performance of amorphous Co-B catalyst. Appl. Surf. Sci. 2020, 518, 146199. [Google Scholar] [CrossRef]
  166. Garron, A.; Bennici, S.; Auroux, A. In situ generated catalysts for NaBH4 hydrolysis studied by liquid-phase calorimetry: Influence of the nature of the metal. Appl. Catal. A Gen. 2010, 378, 90–95. [Google Scholar] [CrossRef]
  167. Vernekar, A.A.; Bugde, S.T.; Tilve, S. Sustainable hydrogen production by catalytic hydrolysis of alkaline sodium borohydride solution using recyclable Co–Co2B and Ni–Ni3B nanocomposites. Int. J. Hydrog. Energy 2012, 37, 327–334. [Google Scholar] [CrossRef]
  168. Tong, D.G.; Chu, W.; Wu, P.; Zhang, L. Honeycomb-like Co–B amorphous alloy catalysts assembled by a solution plasma process show enhanced catalytic hydrolysis activity for hydrogen generation. RSC Adv. 2012, 2, 2369–2376. [Google Scholar] [CrossRef]
  169. Sahiner, N.; Ozay, O.; Inger, E.; Aktas, N. Superabsorbent hydrogels for cobalt nanoparticle synthesis and hydrogen production from hydrolysis of sodium boron hydride. Appl. Catal. B Environ. 2011, 102, 201–206. [Google Scholar] [CrossRef]
  170. Tian, H.; Guo, Q.; Xu, D. Hydrogen generation from catalytic hydrolysis of alkaline sodium borohydride solution using attapulgite clay-supported Co-B catalyst. J. Power Sources 2010, 195, 2136–2142. [Google Scholar] [CrossRef]
  171. Akdim, O.; Demirci, U.B.; Miele, P. Deactivation and reactivation of cobalt in hydrolysis of sodium borohydride. Int. J. Hydrog. Energy 2011, 36, 13669–13675. [Google Scholar] [CrossRef]
  172. Liang, J.; Li, Y.; Huang, Y.; Yang, J.; Tang, H.; Wei, Z.; Shen, P.K. Sodium borohydride hydrolysis on highly efficient Co–B/Pd catalysts. Int. J. Hydrog. Energy 2008, 33, 4048–4054. [Google Scholar] [CrossRef]
  173. Krishnan, P.; Advani, S.G.; Prasad, A.K. Thin-film CoB catalyst templates for the hydrolysis of NaBH4 solution for hydrogen generation. Appl. Catal. B Environ. 2009, 86, 137–144. [Google Scholar] [CrossRef]
  174. Kim, J.-H.; Kim, K.-T.; Kang, Y.-M.; Kim, H.-S.; Song, M.-S.; Lee, Y.-J.; Lee, P.S.; Lee, J.-Y. Study on degradation of filamentary Ni catalyst on hydrolysis of sodium borohydride. J. Alloys Compd. 2004, 379, 222–227. [Google Scholar] [CrossRef]
  175. Arzac, G.M.; Hufschmidt, D.; Jiménez De Haro, M.C.; Fernández, A.; Sarmiento, B.; Jiménez, M.A.; Jiménez, M.M. Deactivation, reactivation and memory effect on Co–B catalyst for sodium borohydride hydrolysis operating in high conversion conditions. Int. J. Hydrog. Energy 2012, 37, 14373–14381. [Google Scholar] [CrossRef] [Green Version]
  176. Dai, H.-B.; Liang, Y.; Wang, P.; Cheng, H.M. Amorphous cobalt–boron/nickel foam as an effective catalyst for hydrogen generation from alkaline sodium borohydride solution. J. Power Sources 2008, 177, 17–23. [Google Scholar] [CrossRef]
  177. Lee, J.; Kong, K.Y.; Jung, C.R.; Cho, E.; Yoon, S.P.; Han, J.; Lee, T.G.; Nam, S.W. A structured Co-B catalyst for hydrogen extraction from NaBH4 solution. Catal. Today 2007, 120, 305–310. [Google Scholar] [CrossRef]
  178. Paladini, M.; Arzac, G.M.; Godinho, V.; De Haro, M.C.J.; Fernández, A. Supported Co catalysts prepared as thin films by magnetron sputtering for sodium borohydride and ammonia borane hydrolysis. Appl. Catal. B Environ. 2014, 158–159, 400–409. [Google Scholar] [CrossRef] [Green Version]
  179. Glavee, G.N.; Klabunde, K.J.; Sorensen, C.M.; Hadjipanayis, G.C. Sodium borohydride reduction of cobalt ions in nonaqueous media. Formation of ultrafine particles (nanoscale) of cobalt metal. Inorg. Chem. 1993, 32, 474–477. [Google Scholar] [CrossRef]
  180. Netskina, O.V.; Ozerova, A.M.; Komova, O.V.; Odegova, G.V.; Simagina, V.I. Hydrogen storage systems based on solid-state NaBH4/CoxB composite: Influence of catalyst properties on hydrogen generation rate. Catal. Today 2015, 245, 86–92. [Google Scholar] [CrossRef]
  181. Liu, C.-H.; Kuo, Y.-C.; Chen, B.-H.; Hsueh, C.-L.; Hwang, K.-J.; Ku, J.-R.; Tsau, F.; Jeng, M.-S. Synthesis of solid-state NaBH4/Co-based catalyst composite for hydrogen storage through a high-energy ball-milling process. Int. J. Hydrog. Energy 2010, 35, 4027–4040. [Google Scholar] [CrossRef]
  182. Delmas, J.; Laversenne, L.; Rougeaux, I.; Capron, P.; Garron, A.; Bennici, S.; Swierczynski, D.; Auroux, A. Improved hydrogen storage capacity through hydrolysis of solid NaBH4 catalyzed with cobalt boride. Int. J. Hydrog. Energy 2011, 36, 2145–2153. [Google Scholar] [CrossRef]
  183. Demirci, U.B. Impact of H.I. Schlesinger’s discoveries upon the course of modern chemistry on B−(N−)H hydrogen carriers. Int. J. Hydrog. Energy 2017, 42, 21048–21062. [Google Scholar] [CrossRef]
  184. Netskina, O.V.; Tayban, E.S.; Prosvirin, I.P.; Komova, O.V.; Simagina, V.I. Hydrogen storage systems based on solid-state NaBH4/Co composite: Effect of catalyst precursor on hydrogen generation rate. Renew. Energy 2020, 151, 278–285. [Google Scholar] [CrossRef]
  185. Netskina, O.V.; Tayban, E.S.; Ozerova, A.M.; Komova, O.V.; Simagina, V.I. Solid-state NaBH4/Co composite as hydrogen storage material: Effect of the pressing pressure on hydrogen generation rate. Energies 2019, 12, 1184. [Google Scholar] [CrossRef] [Green Version]
  186. Netskina, O.V.; Pochtar, A.A.; Komova, O.V.; Simagina, V.I. Solid-state NaBH4 composites as hydrogen generation material: Effect of thermal treatment of a catalyst precursor on the hydrogen generation rate. Catal. 2020, 10, 201. [Google Scholar] [CrossRef] [Green Version]
  187. Netskina, O.V.; Komova, O.V.; Prosvirin, I.P.; Pochtar, A.A.; Ozerova, A.M.; Simagina, V.I. Solid-state hydrogen-generating composites based on sodium borohydride: Effect of the heat treatment of boron–cobalt catalysts on the hydrogen generation rate. Russ. J. Appl. Chem. 2017, 90, 1666–1673. [Google Scholar] [CrossRef]
  188. Netskina, O.V.; Komova, O.V.; Simagina, V.I. Developing effective cobalt catalysts for hydrogen-generating solid-state NaBH4 composite. Catal. Ind. 2018, 10, 166–172. [Google Scholar] [CrossRef]
  189. Wang, Y.; Diaz, D.F.R.; Chen, K.S.; Wang, Z.; Adroher, X.C. Materials, technological status, and fundamentals of PEM fuel cells—A review. Mater. Today 2020, 32, 178–203. [Google Scholar] [CrossRef]
  190. Arzac, G.; Fernandez, A.; Justo, A.; Sarmiento, B.; Jimenez, M.M. Optimized hydrogen generation in a semicontinuous sodium borohydride hydrolysis reactor for a 60W-scale fuel cell stack. J. Power Sources 2011, 196, 4388–4395. [Google Scholar] [CrossRef]
  191. Lee, J.; Kim, T. Micro PEM fuel cell system with NaBH4 hydrogen generator. Sens. Actuators A Phys. 2012, 177, 54–59. [Google Scholar] [CrossRef]
  192. Kim, T. NaBH4 (sodium borohydride) hydrogen generator with a volume-exchange fuel tank for small unmanned aerial vehicles powered by a PEM (proton exchange membrane) fuel cell. Energy 2014, 69, 721–727. [Google Scholar] [CrossRef]
  193. Kim, J.; Kim, T. Compact PEM fuel cell system combined with all-in-one hydrogen generator using chemical hydride as a hydrogen source. Appl. Energy 2015, 160, 945–953. [Google Scholar] [CrossRef]
  194. Okumus, E.; San, F.G.B.; Okur, O.; Turk, B.E.; Cengelci, E.; Kilic, M.; Karadag, C.; Cavdar, M.; Turkmen, A.; Yazici, M.S. Development of boron-based hydrogen and fuel cell system for small unmanned aerial vehicle. Int. J. Hydrog. Energy 2017, 42, 2691–2697. [Google Scholar] [CrossRef]
  195. Oronzio, R.; Monteleone, G.; Pozio, A.; De Francesco, M.; Galli, S. New reactor design for catalytic sodium borohydride hydrolysis. Int. J. Hydrog. Energy 2009, 34, 4555–4560. [Google Scholar] [CrossRef]
  196. Li, H.-Y.; Chen, Y.-T.; Lu, M.-T.; Lai, Y.-H.; Yang, J.-T. Design and testing of a novel catalytic reactor to generate hydrogen. Int. J. Hydrog. Energy 2014, 39, 11945–11954. [Google Scholar] [CrossRef]
  197. Kim, H.; Oh, T.H.; Kwon, S. Simple catalyst bed sizing of a NaBH4 hydrogen generator with fast startup for small unmanned aerial vehicles. Int. J. Hydrog. Energy 2016, 41, 1018–1026. [Google Scholar] [CrossRef]
  198. Jung, E.S.; Kim, H.; Kwon, S.; Oh, T.H. Fuel cell system with sodium borohydride hydrogen generator for small unmanned aerial vehicles. Int. J. Green Energy 2018, 15, 385–392. [Google Scholar] [CrossRef]
  199. Gang, B.G.; Kwon, S. All-in-one portable electric power plant using proton exchange membrane fuel cells for mobile applications. Int. J. Hydrog. Energy 2018, 43, 6331–6339. [Google Scholar] [CrossRef]
  200. Chen, Y.-H.; Lin, J.-C. Reactant feeding strategy analysis of sodium borohydride hydrolysis reaction systems for instantaneous hydrogen generation. Energies 2020, 13, 4674. [Google Scholar] [CrossRef]
  201. Kim, T. Hydrogen production from solid sodium borohydride with hydrogen peroxide decomposition reaction. Int. J. Hydrog. Energy 2010, 35, 12870–12877. [Google Scholar] [CrossRef]
  202. Li, S.-C.; Wang, F.-C. The development of a sodium borohydride hydrogen generation system for proton exchange membrane fuel cell. Int. J. Hydrog. Energy 2016, 41, 3038–3051. [Google Scholar] [CrossRef]
  203. Kim, K.; Kim, T.; Lee, K.; Kwon, S. Fuel cell system with sodium borohydride as hydrogen source for unmanned aerial vehicles. J. Power Sources 2011, 196, 9069–9075. [Google Scholar] [CrossRef]
  204. Kim, T.; Kwon, S. Design and development of a fuel cell-powered small unmanned aircraft. Int. J. Hydrog. Energy 2012, 37, 615–622. [Google Scholar] [CrossRef]
  205. Shang, Y.; Chen, R. Hydrogen storage via the hydrolysis of NaBH4 basic solution: Optimization of NaBH4 concentration. Energy Fuels 2006, 20, 2142–2148. [Google Scholar] [CrossRef]
  206. Shabunya, S.I.; Nesteruk, A.A. Hydrolysis of aqueous-alkaline solutions of sodium borohydride. Asymptotics of low concentrations of the alkali. J. Eng. Phys. Thermophys. 2012, 85, 73–78. [Google Scholar] [CrossRef]
  207. Patel, N.; Fernandes, R.; Miotello, A. Hydrogen generation by hydrolysis of NaBH4 with efficient Co–P–B catalyst: A kinetic study. J. Power Sources 2009, 188, 411–420. [Google Scholar] [CrossRef]
  208. Petit, E.; Miele, P.; Demirci, U.B. By-product carrying humidified hydrogen: An underestimated issue in the hydrolysis of sodium borohydride. ChemSusChem 2016, 9, 1777–1780. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Number of results for a bibliometric resource search made on the Web of Science website (December 2020). The search combination was “compound” AND “hydrogen storage material” AND “hydrolysis”. An exception was made for sodium silicide. In this case, the search combination was “compound” AND “hydrogen storage”.
Figure 1. Number of results for a bibliometric resource search made on the Web of Science website (December 2020). The search combination was “compound” AND “hydrogen storage material” AND “hydrolysis”. An exception was made for sodium silicide. In this case, the search combination was “compound” AND “hydrogen storage”.
Catalysts 11 00268 g001
Figure 2. The influence of the nature of the hydride (NaBH4, NH3BH3) on the formation rate and dispersion of Co-B catalysts forming in situ. The Co-B catalyst instantaneously formed in the NaBH4 hydrolysis medium has a smaller particle size and shows a higher hydrogen generation rate. Conditions—1.2 mmol NaBH4, 10 mL H2O, 0.0117 g CoCl2·6H2O or Co3O4, 40 °C. Adapted from Refs. [123,132].
Figure 2. The influence of the nature of the hydride (NaBH4, NH3BH3) on the formation rate and dispersion of Co-B catalysts forming in situ. The Co-B catalyst instantaneously formed in the NaBH4 hydrolysis medium has a smaller particle size and shows a higher hydrogen generation rate. Conditions—1.2 mmol NaBH4, 10 mL H2O, 0.0117 g CoCl2·6H2O or Co3O4, 40 °C. Adapted from Refs. [123,132].
Catalysts 11 00268 g002
Figure 3. A schematic diagram of the hydrogen generation system with a flow reactor.
Figure 3. A schematic diagram of the hydrogen generation system with a flow reactor.
Catalysts 11 00268 g003
Figure 4. A schematic diagram of the hydrogen generation system with a batch reactor. Two approaches are possible: (a) injection of water into a reactor with solid-state mixture of NaBH4 with a catalyst in the form of powder, pellets, or tablets; (b) injection of solution with catalysts or catalyst precursor into a reactor with solid NaBH4; * indicates the composition of liquid phase; ** indicates the composition of solid phase.
Figure 4. A schematic diagram of the hydrogen generation system with a batch reactor. Two approaches are possible: (a) injection of water into a reactor with solid-state mixture of NaBH4 with a catalyst in the form of powder, pellets, or tablets; (b) injection of solution with catalysts or catalyst precursor into a reactor with solid NaBH4; * indicates the composition of liquid phase; ** indicates the composition of solid phase.
Catalysts 11 00268 g004
Table 1. Water-reactive materials for hydrogen storage.
Table 1. Water-reactive materials for hydrogen storage.
MaterialReaction∆H, kJ/mol H2Calc. GHSC, wt.%AdvantagesDisadvantages
Al
[31]
2Al + 6H2O → 2Al(OH)3 + 3H2−280 (at ~50–100 °C)3.7+ low cost, abundance on the earth
+ light weight
+ environmental-friendly products, may be raw materials of ceramics
+ improved and tunable properties at additions of dopants (salts, metals, oxides, hydrides, carbon materials, etc.), alloying with metals, reducing the particle size (BM), etc.
− surface passivation of native oxide layer (more relevant to Al)
− the hydrolysis products deposition on metal particles surface (more relevant to Mg)
− heat and mass transfer problem
− unpromising regeneration process of solid hydrolysis products
Mg
[32]
Mg + 2H2O → Mg(OH)2 + H2−3543.3
MgH2
[33]
MgH2 + 2H2O → Mg(OH)2 + 2H2−1606.4+ low cost
+ can be industrially produced with a high energy efficiency
+ improved and tunable properties at addition of Brønsted acids, salts, metal catalysts, and alloying with rare earth elements, BM, etc.
− unstable in the presence of moisture
− reacts with water very slowly and incompletely
− formation of dense passivation layers of Mg(OH)2
− complicated regeneration process Mg(OH)2 to Mg (by reactive hydrogen plasma process)
NaSi
[34]
2NaSi + 5H2O → Na2Si2O5 + 5H2−355.2+ commercial availability
+ no activation procedure is required
+ by-product is of market value
− unstable in the presence of moisture
− use of liquid hydrocarbons for stabilization
− excess water is required
− problem with design of the reactor to control the rate of gas evolution
NaBH4 [17,18,35]NaBH4 + 2H2O → NaBO2 + 4H2−6010.8+ commercial availability
+ high GHSC
+ safe and controllable on-board H2 generation, commercialized generators
+ producing pure H2 at ambient temperatures without energy supply
+ inflammable
+ high solubility (14.5 mol·L−1): 30 wt.% NaBH4 solution contains 6.7 wt.% H2
+ improved and tunable properties at addition of catalysts or acids
+ concentrated water solutions or solid pellets with catalysts may be used as H2 sources
− instability of aqueous solutions, stabilization with NaOH is required
− deactivation of catalysts
− hydrolysis products deposition on the catalyst surface and reactor walls in limited water conditions
− off-board and high-cost regeneration of solid hydrolysis products
− heat and mass transport problem at design of reactors for NaBH4 interaction with stoichiometric amount of water
KBH4
[36]
KBH4 + 2H2O → KBO2 + 4H2−558.9+ improved properties at addition of catalysts− instability of aqueous solutions
− low solubility (3.5 mol·L−1): low GHSC
LiBH4
[37]
LiBH4 + 2H2O → LiBO2 + 4H2−9013.8+ high GHSC
+ improved properties at addition of catalysts, dopants (MWCNTs 1, NH3BH3), use of double-solvents, etc.
− preparation from NaBH4
− flammable and unstable in the presence of moisture, instability of water solutions
− high exothermic hydrolysis process
− incomplete hydrolysis
− hydrolysis products deposition on the catalyst surface and reactor walls in limited water conditions
NH3BH3 [29]NH3BH3 + 2H2O → NH4BO2 + 3H2<−509+ high stability of aqueous solution
+ high solubility (11.4 mol·L−1)
+ improved and tunable properties at addition of catalysts
− high cost
− the presence of NH3 impurity in the H2 stream
− deactivation of catalysts
− hydrolysis products deposition on the catalyst surface and reactor walls in limited water conditions
− off-board and high-cost regeneration of solid hydrolysis products
N2H4BH3 [38]N2H4BH3 + 3H2O → N2H4 + B(OH)3 + 3H2
N2H4BH3 + 3H2O → N2 + B(OH)3 + 5H2
81.5
39
6
10
+ possibility of catalytic dehydrogenation of the N2H4 moiety
+ high theoretical GHSC
+ stability of water solutions at pH ≥ 7/5
+ improved properties at addition of catalysts
− high cost
− search for effective, low-cost, and stable catalysts for the hydrolytic dehydrogenation of both the BH3 moiety and the N2H4 moiety
− low solubility (1.3 mol·L−1)
− lack of knowledge of the characteristics of the process in limited water conditions, providing GHSC
Table 2. Changes in the catalytic activity of bulk and supported Co-B catalysts during their long-term testing in the hydrolysis of NaBH4.
Table 2. Changes in the catalytic activity of bulk and supported Co-B catalysts during their long-term testing in the hydrolysis of NaBH4.
CatalystTest ConditionsThe Number of Cycles or Test Duration% of Initial ActivityRef.
Co-B obtained in situ from CoCl22% NaBH4, 4% NaOH, 30 °C389[166]
19% NaBH4, 4% NaOH, 30 °C389
Co-B obtained in situ from LiCoO20.5% NaBH4, 40 °C6100[159]
1440
Co-Co2B obtained by reduction of CoCl2 in a methanol solution of NH3BH30.76% NaBH4, 0,1% NaOH, 30 °C12100[167]
Honeycomb Co-B obtained by plasma treatment of
Co(NH3)62+ + KBH4 + triethanolamine solution
2% NaBH4, 7% NaOH, 25 °C30 h100[168]
thendeactivation
Co-B obtained in situ from Co3O410% NaBH4, 5% NaOH, 25 °C3 h≈70[133]
20 h≈70
Co-B in hydrogel obtained by impregnation with CoCl2 solution and reduced by NaBH40.2% NaBH4, 5% NaOH, 30 °C593[169]
C-B/attapulgite clay obtained by impregnation with Co(NO3)2 and reduced by NaBH410% NaBH4, 5% NaOH, 25 °C931[170]
Co/nickel foam obtained by electrodeposition from CoCl210% NaBH4, 2% NaOH, 80 °C253[171]
335
Co-B/Pd-nickel foam obtained by dip-coating from CoSO4 and reduced by NaBH420% NaBH4, 0.4% NaOH, 30 °C70≈100[172]
Co-B/nickel foam obtained by electrodeposition from CoSO45% NaBH4, 5% NaOH, 25 °C4≈88[173]
10% NaBH4, 5% NaOH, 25 °C5 h≈60–70
60 h≈60–70
Table 3. Characteristics of H2 evolution from the solid-phase system based on NaBH4.
Table 3. Characteristics of H2 evolution from the solid-phase system based on NaBH4.
Solid Phase CompositionH2O:NaBH4
in Moles
Conversion
(%)
GHSC
(wt.%)
Reference
80 wt.% NaBH4,
20 wt.% CoCl2
9:1783.4[47]
85 wt.% NaBH4
16 wt.% CoCl2·6H2O 1
4:181.26.7[43]
87.5 wt.% NaBH4,
12.5 wt.% Co2+/IR-120
4:190.956.7[181]
90 wt.% NaBH4,
10 wt.% Co2B
3:1-8.7[182]
1 a CoCl2 solution was dosed.
Table 4. The reported hydrogen generators with Co-containing catalysts.
Table 4. The reported hydrogen generators with Co-containing catalysts.
Hydrogen Generator
Design
CatalystConditionsOperating TimeCommentsOutput
Characteristics
Ref.
Semicontinuous reactor, H2 washing flask (water)Co-B/Ni foam19% NaBH4, 4.5% NaOH, 2.5 mL/min3 cycles of 60 minStable H2 generation with cat. re-activation between cycles1160 mL H2/min, conversion 90%, start-up time 1 < 5 min[190]
Flow catalytic reactor, gas–liquid separator, mesh filter, H2 dehumidifier (silica)Co/Al2O3 pellets15% NaBH4, 5% NaOH,
3 mL/min
2.5 h UAV
flying test
Crushing Al2O3946 mL H2/min, integration with 100 W PEMFC stack, specific energy density 165 Wh/kg[185,186]
Flow catalytic microreactor, hydrogen separator with gas–liquid separation membraneCo-P-B/Ni foam10% NaBH4, 5% NaOH, 0.06 mL/min10 cycles of 30 minDecrease in max power output by 8% in 6 cycles16.1 mL H2/min, conversion 98.8%, integration with micro PEMFC, max power output 174.6 mW[191]
Flow catalytic reactor with cooling fans, volume-exchange fuel tank (fuel and spent fuel exchange the volume within a single fuel tank)Co-B/
ISOLITE monolith
Alkaline 15% NaBH4, 2.5 mL/min2 cycles of 30 min with 5 min rest in between, 1 h UAV flying testRestart of H2 generation with quick start-up time and decrease in rate by 9%1331 mL H2/min, integration with 100-W PEMFC stack[192]
All-in-one reactor (cat., chamber of pressurized H2 (4–5 bar) and by-product separator are combined in single space) with cooling devices, H2 wash tank; periodic pressure regulation be NaBH4 supply; autonomous operationCo-B/Ni foam Alkaline 25% NaBH42 cycles of 5 minIncrease in start-up time, stable pressure in H2 chamber with cycles; cat. washing between cycles; stable HGR 2 and temperature during 17 min operation≈1200 mL H2/min, integration with 100 W PEMFC stack, total energy density 360 Wh/kg[193]
Flow catalytic reactor with cooling jacket, by-product level sensor, H2 cooling device, resin for capture of sodium from H2; removed water fed back to the reactorCo-Fe-Ni powder20% NaBH4,
3% NaOH
20 cycles of 2 h, 3 h FC stack operatingDecrease in HGR by 1% in 10 cycles, constant rate over next 10 cycles5400 mL H2/min, integration with 200 W PEMFC stack, total energy density 325 Wh/kg[194]
Reactor with inner magnet attracting the magnetic cat. powder, H2 filter and drying device; recycling spent fuel back to fuel tankModified comm. Co catalyst20% NaBH4, 10% NaOH,
10 mL/min, preheated to 35 °C
20 hGood cat. stability25 mL H2/min, conversion 90%[195]
π-shaped catalytic reactor, gas–liquid separator, H2 dehumidifier (silica)CoOx + Ni + PVDF 3/Ni foamAlkaline 15% NaBH4, 0.5–2 mL/min Improved characteristics compared to conventional reactor via rapid H2 discharge from cat. surfaceconversion 90.2%[196]
Flow catalytic reactor, by-product separator, H2 wash (water) and dehumidifier (silica) tanksCo-P/Ni foam20% NaBH4,
5% NaOH, 5.22 mL/min, preheated to 30 °C
3 hCat. high durability2670 mL H2/min, conversion 90%, start-up time 71 s, integration with 200 W PEMFC stack, energy density 252.1 Wh/kg[44]
20% NaBH4,
1% NaOH,
2.5 mL/min for start-up, 12.9 mL/min next, thermostated to 20 °C
2 cycles of 2 h with 2 h rest in between Restart of H2 generation with the same start-up time (2.5 min) and decrease in rate by 12.4%; NaBO2 deposition on cat.≈6500 mlH2/min, conversion 94%, start-up time 2.5 min[197]
15–18% NaBH4, 5% NaOH, 3.07–3.5 mL/min, preheated to 30 °C2 cycles of 1 h, 3 h FC stack operatingCat. washing and drying in air between cycles; increase in start-up time, decrease in conversion by 42.5% in 2nd cycle; stable H2 generation during cycle1570 mL H2/min, 97.8% conversion, integration with 100 W PEMFC stack, specific energy density 185.2 Wh/kg, max power 95.96 W[198]
Flow catalytic reactor, liquid–gas separator with purge pump, H2 cooling device (finned container with water + cooling fan), H2 dehumidifier (silica)Co/Al2O3 on the top of Co-P/Ni foam20% NaBH4,
12 mL/min
2 h2 catalysts provide fast start-up5900 mL H2/min, integration with 500 W PEMFC stack, specific energy density 211 Wh/kg[199]
Packed-bed catalytic reactor with two exit channels (top for H2 and side for byproduct) placed in water bath (70 °C),
2 water traps, dessicator
Co-B/
IR-120
5% NaBH4, 2% NaOH, 2.4 mL/min for start-up, 1.02 mL/min next3 cycles for ≈30 minModeling and validating a strategy of combining reactant feed rates for instant and stable H2 generation130 mL H2/min, conversion 99.3%, start-up time 70 s[200]
Batch reactor with NaBH4 pellets, injection of Co(NO3)2 solution, H2 separatorCo-B reduced in situ1.6 g of NaBH4 pellets, 2 mL of 0.1 M Co(NO3)2
solution
30 minInstable HGRMax 913.5 mL H2/min, 100% conversion, integration with 100 W PEMFC stack[201]
Batch reactor with pills of solid-state NaBH4 composites, rapid injection of waterCo-B/IR-120 reduced in situNaBH4 (3 g), cat. (0.6 g) and silicon rubber (2.5 g) mixed in 20 solid-state pills composites; 20 g H2O2 hRegulation HGR depending on the composition, shape and number of solid-state composites25 mL H2/min, integration with 2 W PEMFC stack to power cellular phone[52]
Batch reactor with cat., heat exchanger, H2 buffer tank, batch control moduleCoOx/Ni foam15% NaBH46 cycles for 6 min; 1 h FC stack operatingEfficiency depends on NaBH4 concentration, volume of bath reactor and algorithm of next batch additionMax 128 L H2/min, conversion > 90%, integration with 3 kW PEMFC stack[202]
1 Start-up time—the time for stabilization constant hydrogen flow. 2 HGR—hydrogen generation rate. 3 PVDF—polyvinylidene fluorid.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Simagina, V.I.; Ozerova, A.M.; Komova, O.V.; Netskina, O.V. Recent Advances in Applications of Co-B Catalysts in NaBH4-Based Portable Hydrogen Generators. Catalysts 2021, 11, 268. https://doi.org/10.3390/catal11020268

AMA Style

Simagina VI, Ozerova AM, Komova OV, Netskina OV. Recent Advances in Applications of Co-B Catalysts in NaBH4-Based Portable Hydrogen Generators. Catalysts. 2021; 11(2):268. https://doi.org/10.3390/catal11020268

Chicago/Turabian Style

Simagina, Valentina I., Anna M. Ozerova, Oksana V. Komova, and Olga V. Netskina. 2021. "Recent Advances in Applications of Co-B Catalysts in NaBH4-Based Portable Hydrogen Generators" Catalysts 11, no. 2: 268. https://doi.org/10.3390/catal11020268

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