Next Article in Journal
Effects of Pressure Rollers with Variable Compliance in the Microfinishing Process Utilizing Abrasive Films
Previous Article in Journal
Data-Driven Design of Nickel-Free Superelastic Titanium Alloys
Previous Article in Special Issue
Rapid Plasma Electrolytic Oxidation Synthesis of Intermetallic PtBi/MgO/Mg Monolithic Catalyst for Efficient Removal of Organic Pollutants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Co and Co3O4 in the Hydrolysis of Boron-Containing Hydrides: H2O Activation on the Metal and Oxide Active Centers

by
Vladislav R. Butenko
1,
Oksana V. Komova
1,*,
Valentina I. Simagina
1,
Inna L. Lipatnikova
1,
Anna M. Ozerova
1,
Natalya A. Danilova
1,2,
Vladimir A. Rogov
1,2,
Galina V. Odegova
1,
Olga A. Bulavchenko
1,
Yuriy A. Chesalov
1 and
Olga V. Netskina
1,2
1
Boreskov Institute of Catalysis SB RAS, 5 Akademika Lavrentieva Ave., Novosibirsk 630090, Russia
2
Department of Natural Sciences, Novosibirsk State University, 1 Pirogova Str., Novosibirsk 630090, Russia
*
Author to whom correspondence should be addressed.
Materials 2024, 17(8), 1794; https://doi.org/10.3390/ma17081794
Submission received: 13 March 2024 / Revised: 31 March 2024 / Accepted: 10 April 2024 / Published: 13 April 2024
(This article belongs to the Special Issue Advances in Multicomponent Catalytic Materials)

Abstract

:
This work focuses on the comparison of H2 evolution in the hydrolysis of boron-containing hydrides (NaBH4, NH3BH3, and (CH2NH2BH3)2) over the Co metal catalyst and the Co3O4-based catalysts. The Co3O4 catalysts were activated in the reaction medium, and a small amount of CuO was added to activate Co3O4 under the action of weaker reducers (NH3BH3, (CH2NH2BH3)2). The high activity of Co3O4 has been previously associated with its reduced states (nanosized CoBn). The performed DFT modeling shows that activating water on the metal-like surface requires overcoming a higher energy barrier compared to hydride activation. The novelty of this study lies in its focus on understanding the impact of the remaining cobalt oxide phase. The XRD, TPR H2, TEM, Raman, and ATR FTIR confirm the formation of oxygen vacancies in the Co3O4 structure in the reaction medium, which increases the amount of adsorbed water. The kinetic isotopic effect measurements in D2O, as well as DFT modeling, reveal differences in water activation between Co and Co3O4-based catalysts. It can be assumed that the oxide phase serves not only as a precursor and support for the reduced nanosized cobalt active component but also as a key catalyst component that improves water activation.

1. Introduction

Hydrogen is a perfect candidate for future fuel use due to its high energy density, oxidation product (H2O), environmental friendliness, and the possibility of H2 recovery from water using renewable energy sources [1,2]. A significant challenge with hydrogen energy lies in its storage and transportation. Compressed gaseous and liquid H2 do not meet the requirements for safety and compactness. This leads to the development of compact hydrogen storage and generation systems based on adsorbed or chemically bonded states [3,4]. For example, solid hydride compounds are actively studied. Among them, sodium borohydride (NaBH4, SBH) [5], ammonia borane (NH3BH3, AB) [6,7], and ethylenediamine bisborane ((CH2NH2BH3)2, EDBB) [8,9] are characterized by high values of hydrogen density (10.8, 19.6, and 16.3 wt%, respectively). Resistance to air moisture, which increases in the row SBH < AB < EDBB, makes them promising for storing and generating hydrogen in fuel cell-based mobile energy devices.
Catalytic hydrolysis of these hydrides is widely studied [10] because it produces H2 at ambient temperatures (1)–(3) [11,12]:
NaBH4 + 4H2O → NaB(OH)4 + 4H2 ΔrH0298 = −217 kJ/mol Wm = 10.2 wt%,
NH3BH3 + 3H2O → NH3 + B(OH)3 + 3H2 ΔrH0298 = −156 kJ/mol Wm = 8.5 wt%,
and (CH2NH2BH3)2 + 6H2O → (CH2NH2)2 + 2B(OH)3 + 6H2 Wm = 7.1 wt%
where Wm is the gravimetric hydrogen capacity of the hydrogen-generating system calculated from the stoichiometry of the reactions (1)–(3).
The catalyst is an important control tool for H2 generation, and the addition of water to the solid-phase composition of the hydride and catalyst can be considered the easiest way to generate H2 on demand [13,14]. The study of the catalytic hydrolysis of AB and EDBB is significant also because it explores the potential for a two-stage hydrothermolysis process. This process operates under a limited water supply, where the heat generated from the exothermic hydride hydrolysis warms the reaction layer. This, in turn, triggers a low-temperature solid-phase hydride dehydrogenation that occurs in the absence of water [15,16].
In most cases, the literature presents data on the high Co3O4 activity in SBH hydrolysis [17,18,19,20,21,22]. In the reaction medium of such a strong reducing agent, in situ activation of Co3O4 occurs through its partial reduction, forming catalytically active centers. Many techniques have been applied to study the in situ-produced catalytically active phase during the reduction in cobalt compounds, including Co3O4 [23]. As a result, amorphous nanoparticles of cobalt borides (CoxB) [17], metal clusters coated with an amorphous layer containing CoxB [24], and metal cobalt trimers stabilized in a borohydride matrix [25] have been proposed as catalytically active phases. Until recently, Co3O4 activity in SBH and AB hydrolysis has been associated only with reduced nanosized cobalt forms (Co0, CoxB then be referred to as CoBn) [23]. Based on traditional knowledge of electroless plating techniques, NaBH4 often exhibits pronounced boron deposition. The application of boranes enables the production of metal depositions devoid of even trace amounts of boron. The impossibility of forming active centers under the action of EDBB seems to explain the lack of reports of Co3O4 activity in the hydrolysis of this hydride.
In the last two to three years, a current approach to designing boron-containing hydride hydrolysis catalysts involves producing bifunctional catalysts [26,27], whose active centers should activate both hydride (4) and water (5).
R-H + 2* = R* + H*, where R = BH3, NH3BH2, BH3NH2CH2CH2NH2BH2
H2O + 2* = OH* + H*
Correlations between the experimentally measured catalyst activity and the DFT-calculated energy parameters of the reagents activation ((4) and (5)) on the catalyst surface suggest that the breakage of the O-H bond of the water molecule (5) on the metal active center can be considered as a rate-determining step (RDS). DFT results confirm that activating water on a metal surface requires overcoming a higher energy barrier compared to the hydride activation stage [28,29,30]. Additionally, the published results on the kinetic isotopic effect measurement (KIE, kH/kD) for the hydrolysis of AB and SBH in the case of replacing H2O with D2O (Table S1 in the Supplementary Materials) also show that the breakage of the O-H bond of the water molecule (5) on the metal active center can be considered as an RDS.
From this point of view, it is interesting to discuss about the catalytic properties of Co3O4 again. The published works make it possible to state that without the formation of a sufficient content of the nanosized metal-like phase (CoBn), hydrolysis of hydride is not carried out, which is due to the low rate of B-H bond activation on the catalyst surface. However, during intense hydrogen evolution, the degree of Co3O4 reduction in the reaction medium is uncertain. The characteristics of the unreduced residual cobalt oxide phase and its influence on the kinetic regularities are also under discussion.
It is well known that Co3O4 treatment with sodium borohydride in aqueous solution is widely used to form anionic vacancies (Vo) in the oxide structure, which has improved the characteristics of catalysts and materials for various applications [31,32,33,34,35]. The results obtained in different research groups are consistent. Though the main crystalline phase remains Co3O4, there is an increase in the Co2+/Co3+ ratio on the surface of the treated sample, with a simultaneous increase in the content of anionic vacancies. In most studies, the relative content of anionic vacancies is judged by the results of O 1s XPS spectra deconvolution, where the ratio of the content of lattice oxygen species and surface adsorbed oxygen is estimated.
The first report on the bifunctional properties of activated Co3O4 was published in 2020 [36]. In this study, the authors, when explaining the high catalytic activity of Co3O4/CN (CN = carbon nitride) in AB hydrolysis, underlined the importance of Co0 in the activation of AB (4) and Co3O4 in the process of water activation (5). They have used DFT modeling to calculate the adsorption energy (Eads) and activation energy (Ea) for the dissociative adsorption of water (5) on CN-supported clusters of Co0, Co3O4, and Co3O4 with introduced anion vacancies in its structure (Co3O4-VO). As expected, water activation on the oxides was shown to be more effective than on the metals. But, the calculations in this paper demonstrate that the formation of anionic vacancies in Co3O4-VO did not decrease the value of Eads and Ea compared to Co3O4. However, the opposite results on the influence of oxygen vacancies in the Co3O4 structure on Eads have been demonstrated in [37,38].
Therefore, this work will focus on the comparison of the kinetics of H2 evolution in the hydrolysis of boron-containing hydrides of different natures and reducing abilities (NaBH4, NH3BH3, and (CH2NH2BH3)2) over the ex situ-formed cobalt metal catalyst and the Co3O4-based catalysts whose activation proceeded in the reaction medium (in situ). In the beginning, DFT calculations for metal clusters of Co, Ni, and Cu will be used to emphasize the importance of the water activation stage in the processes under study and get the answer to why Co0 is considered the most active metal in these processes. The kinetic regularities of the Co0 catalyst in the hydrolysis of the hydrides with KIE measurement at replacing H2O with D2O will then be obtained to confirm the DFT results. Subsequently, the results of the activation of Co3O4 in different hydride environments will be presented, and the effect of the reducing ability of hydrides on the hydrogen generation rate will be discussed. To increase the reduction degree of Co3O4 in the reaction medium of AB and EDBB, the studied cobalt oxide will be modified with CuO. It has been established that the contribution of Cu to the overall process can be ignored. To characterize the oxide phase in activated Co3O4, a Co3O4 sample will be removed from the SBH reaction medium and studied using several methods (TEM, TPR H2, XRD, Raman, and ATR FTIR). Within the one DFT model, the energy parameters (Ea, Eads) of the dissociative adsorption of water on the clusters of Co3O4 and Co3O4 with two types of anionic vacancies will be compared. These results, supported by the KIE measurements in the hydrolysis reaction tests, will be compared with the data for the Co0 catalyst. This will reveal the differences in the water activation between a cobalt metal catalyst and a catalyst containing a cobalt oxide phase.

2. Materials and Methods

2.1. Catalytic Materials under Study

A commercial sample of Co3O4 (clean for analysis, GOST 4467-79, SoyuzKhimProm, Novosibirsk, Russia) was used. To modify Co3O4 with CuO (10 wt%), 4.0053 g of Co3O4 and 0.6123 g of CuCO3·Cu(OH)2 (clean, GOST 8927-79, Reachem, Moscow, Russia) were ground in a hand mortar and the mixture was calcined at 300 °C for 4 h.
The metal cobalt catalyst was prepared by the galvanic replacement method, as previously reported [39]. Aluminum powder (ASD-0 grade, TU 1791-007-49421776-2011, Sual-PM, Shelekhov, Russia), with an average particle size of 80 ± 18 μm, was used as a template. Initially, 0.5 g of Al powder was degreased in acetone and etched in a 1 M HCl solution (5 mL) to eliminate the surface oxide layer. Next, 15 mL of a 0.23 M solution of cobalt (III) acetylacetonate (pure, TU 6-09-09-520-73, Reakhim, Moscow, Russia) in ethanol was added to the Al suspension. The reaction was carried out in an ultrasonic bath (Sapfir, Moscow, Russia) at 60 °C and 100 W for 2 h. The resulting sample was separated from the reaction medium with a magnet, washed with distilled water, and treated with a 2.5 M NaOH solution for 2 h to remove the residue of Al. Afterward, it was washed with distilled water until a neutral pH was achieved, rinsed with acetone, and then evacuated for 2 h at room temperature. Based on elemental analysis and XRD data, the resulting cobalt catalyst contains 98.1 wt% of Co and 0.06 wt% of Al and consists of one crystalline phase of metallic cobalt (hcp) with a CSR of 12 nm.
The scanning electron microscopy images of Co3O4, 10% CuO-90% Co3O4, and Co0 are presented in Figure S1 in the Supplementary Materials.

2.2. Experimental Methods of Investigations

X-ray diffraction analysis (XRD) was performed on a high-resolution STOE Stadi MP diffractometer in a transmission geometry (STOE & Cie GmbH, Darmstadt, Germany) in the range of angles 2–50° with a step 2θ = 0.015° using a Mythen2 1K (Dectris, Baden-Daettwil, Switzeland). MoKα1 radiation (λ = 0.709 Å) was used. A phase analysis was performed by the Rietveld method. The average coherent scattering regions (CSR) were determined using the Scherrer formula from the following reflections: 111 for CoO and 311 for Co3O4. The phases were identified using the following data: CoO [PDF 42-1300] and Co3O4 [PDF 42-1467].
Temperature-programmed reduction with hydrogen (TPR H2) was carried out in a custom-built unit equipped with a flow quartz reactor and a thermal conductivity detector. Before the experiment, the cobalt oxide sample (7 mg) was mixed with quartz sand and purged with argon for 2 h at room temperature. A gas mixture of 10 vol.% H2 in Ar was supplied at a rate of 40 mL/min. The rate of heating from room temperature to 900 °C was 10 K/min.
The Raman spectrometer T64000 (Horiba Jobin Yvon, Edison, NJ, USA) with a micro-Raman setup was used to measure the Raman spectra. All experimental spectra were collected in the backscattering geometry using the 514.5 nm line of an Ar+ laser. The spectral resolution was not worse than 1.5 cm−1. The detector was a silicon-based CCD matrix cooled with liquid nitrogen. The power of the laser beam reaching the sample was 2 mW. The band at 520.5 cm−1 of Si single crystal was used to calibrate the spectrometer.
High-resolution transmission electron microscopy (HR TEM) studies were carried out using a JEM-2010 instrument (JEOL Ltd., Akishima, Japan) with a lattice resolution of 1.4 Å and an accelerating voltage of 200 kV. Before the experiment, the samples were fixed on “holey” carbon films supported by copper grids.
Attenuated total reflection infrared spectroscopy (ATR FTIR) was performed on an Agilent Cary 630 (Agilent Technologies, Santa Clara, CA, USA) spectrometer in the range of 540–4000 cm−1.
The scanning electron microscopy (SEM) images were obtained with a JEOL JSM-6460 LV (Jeol, Akishima, Japan) instrument.

2.3. Catalytic Hydrolysis of Boron-Containing Hydrides

For catalytic experiments, commercial NaBH4 (98%, CAS 16940-66-2, Chemical Line, Sankt-Petersburg, Russia) was used. Synthesis of NH3BH3 with 97% purity and (CH2NH2BH3)2 with 96% purity was performed according to well-known [8,40] (see Supplementary Materials). The phase composition of the synthesized hydrides was confirmed by XRD and ATR FTIR methods. The purity of hydride samples was estimated in a catalytic hydrolysis experiment in the presence of a preliminary reduced cobalt catalyst, taking into account the stoichiometry of reactions ((2) and (3)).
The SBH, AB, and EDBB hydrolysis was carried out at 40 °C under 750 rpm stirring. First, 10 mL of distilled water was added into a glass reactor (V = 52.5 mL) and preheated to the required temperature; then, the hydride and the catalyst powder were added in succession. The reactor was sealed, and the evolving hydrogen was passed through a condenser to the gas burette (100 mL) equipped with digital pressure Sendo sensor SS312 (Sendo sensor, Shenzhen, China). The tested amount of hydride sample was 35 mg (0.925 mmol) for SBH, 38 mg (1.23 mmol) for AB, and 54 mg (0.615 mmol) for EDBB. In all cases the initial concentration of B-H bond in the reaction medium was 3.7 mmol in 10 mL of solution. In all experiments, the amount of catalyst was 11.7 mg. The obtained values of hydrogen volume were reduced to the N.T.P.

2.4. Density Functional Theory Calculations

As the surfaces under investigation, the 111 facets for metals and CuO and the 311 facets for Co3O4 were selected. The adsorption of H2O on the surfaces of metals and oxides was studied using density functional theory (DFT) calculations, conducted within the standard Kohn–Sham formalism employing a plane-wave basis set with a kinetic energy cutoff of 350 eV. These calculations were carried out using the Vienna ab initio simulation package (VASP). The Perdew–Burke–Ernzerhof (PBE) functional of the generalized gradient approximation (GGA) was chosen for the exchange–correlation potential calculations, disregarding van der Waals interactions. The influence of atomic nuclei on the electron density was accounted for by the projector augmented wave (PAW) method. For integration in reciprocal space, a Monkhorst–Pack grid of (2 × 2 × 1 k-points) was utilized. The conjugate gradient algorithm, with an energy convergence criterion of 1 × 10−5 eV, was employed to ensure the convergence of atomic positions, ensuring that the forces were less than 2 × 10−2 eV/Å.
The adsorption energy (Eads) between the adsorbate molecule and the adsorbent surface was calculated as follows:
Eads = [EAB − (EA + EB)],
where EAB represents the total energy of the surface with the adsorbed molecule A, EA is the total energy of the adsorbate molecule, and EB is the total energy of the adsorbent surface.
The energy (ΔE) of dissociative adsorption was calculated as follows:
ΔE = EFS − EIS,
where EFS and EIS are the final and initial energies, respectively.
The calculation of activation energies was performed using the nudged elastic band method with a climbing image (NEB-CI), where preoptimized structures were employed as initial and final points for the reaction pathway calculations ((4) and (5)). These calculations were continued until the magnitude of the forces orthogonal to the reaction coordinate was reduced to less than 0.05 eV/Å. The activation energy was calculated using the formula:
Ea = ETS − EIS,
where ETS is the energy of the transition state.
Bader analysis was conducted using the code developed by Professor G. Henkelman’s group [41].

3. Results

3.1. The Characteristics of Reactions When Co0 Is Used as a Catalyst

3.1.1. DFT Modeling of H2O and NH3BH3 Activation on Metallic Surfaces of Co, Ni, and Cu

To reveal the RDS of hydrolysis of boron-containing hydrides, the dissociative adsorption processes of AB (4) and water (5) on the 111 surfaces of metal clusters of Co, Ni, and Cu were studied (Figure S2). Ni and Cu are taken as comparison metals. Adsorption has been found to be characterized by insignificant changes in the geometry of both the metal surface and the adsorbate molecules. At the same time, the increase in the values of Eads of H2O is observed in a row: Co < Ni < Cu (Table 1). This indicates a decrease in its bonding strength when moving from Co to Cu. This row is also observed when comparing Eads for AB. But, in the case of AB, the effect of metal nature on Eads values is more pronounced, which is reflected in more noticeable changes in the geometry of the adsorbed hydride molecule (Table 1).
Using preoptimized initial and final states (Figure S2), the reaction profiles for the B-H bond cleavage in the AB molecule (4) and the O-H bond cleavage in the water molecule (5) were constructed through interpolation methods. Local maxima along the reaction coordinate were identified using the NEB-CI method, and the activation energy (Ea) for each metal was determined. The obtained results demonstrate a linear correlation between the reaction energies (ΔE) and Ea (Figure 1a), indicating the well-known Brønsted–Evans–Polanyi (BEP) correlation [42]. The BEP correlation makes it possible to predict the effectiveness of metal catalysts with the same mechanism for activating reagents and, therefore, with the same surface geometry of the resulting transition complex. It is important to note that the dissociation of water is accompanied by a higher activation energy than the hydride dissociation (Figure 1a). The high contribution of water activation to the RDS of AB hydrolysis can be explained by the stronger O-H bond (429.9 kJ/mol) in water than the B-N bond (377.9 kJ/mol) and B-H bond (345.2 kJ/mol) in AB [43]. The published data on DFT results confirm that activating water on a metal surface requires overcoming a higher energy barrier compared to the hydride activation stage [28,29,30].
It was additionally shown that there is a relationship between Eads and Ea for breaking the O-H bond in the water molecule (Figure 1b). This is consistent with the Sabatier principle [44], which states that the adsorption energy of water molecules on the catalyst surface should be strong enough to allow for easier water dissociation. At the same time, the adsorption energies of the intermediates (OH*, H*) should be optimal to ensure their further transformation with the release of active centers for further adsorption of reagents. This relationship is common in DFT calculations for water dissociation. The comparison of the DFT results of this work with literature data for different metal surfaces is presented in Figure S3a.

3.1.2. Activity of Co Metal Catalyst in the Hydrolysis of NaBH4, NH3BH3, and (CH2NH2BH3)2

The DFT results presented above predict higher cobalt activity in chemical processes proceeding with the water dissociation stage (5). This is confirmed by numerous published experimental data on the high catalytic activity of the cobalt-containing catalysts in the hydrolysis reaction of both NaBH4 and NH3BH3 [4,5,23,45]. The activity of cobalt catalysts in hydrolysis (CH2NH2BH3)2 has yet to be well studied, but according to our research, the Co0 catalyst synthesized by galvanic replacement is quite active in this process. The rate of hydrogen release is consistent with the reducing ability of the hydride and decreases in a row: SBH > AB > EDBB (Figure 2a). This appears to be due to the difference in hydride activation on the catalyst surface.
The KIE measurement using isotopically labeled reagents is a widely used experimental technique to reveal the RDS. The results on KIE for catalysts containing metal active components have mostly been investigated in processes (1) and (2). They have shown that for the hydrolysis of both SBH and AB, a primary isotopic effect (kH/kD > 2) is observed when the reaction takes place in deuterated water, and a secondary KIE is noted with isotope-labeled hydrides [43,46,47]. Our tests of the Co0 catalyst in heavy water are shown in Figure 2a. As expected, there is a slowdown in the hydrogen evolution rate. For all studied hydrides, kH/kD > 2 (shown in Figure 2a) is obtained, indicating the direct formation of the deuterium bond in the structure of the activated complex. This observation supports the idea that the dissociation of water (5) can be considered as the RDS when using metallic cobalt as a catalyst. This is consistent with the DFT results and the literature data for other metal catalysts (Table S1).
The results on SBH hydrolysis show that the tested Co0 catalyst is quite stable. In four consecutive reactions, without separating the reaction product from the catalyst, the H2 generation rate remained almost constant despite the accumulation of sodium borate in the reaction medium. Also, changes in kH/kD ratios were no more than 10 rel% (Figure 2b).

3.2. Features of Co3O4 Catalyst Use

3.2.1. Catalytic Activity of Co3O4 in the Hydrolysis of NaBH4, NH3BH3, and (CH2NH2BH3)2

It is known that Co3O4 is activated in a hydride reaction medium, and in situ, the formed active phase begins to catalyze the hydrolysis process of hydrides [23]. Our results show that at 40 °C, hydrogen evolution occurs only during SBH hydrolysis after a short induction period (~1 min) (Figure 3). When the reaction temperature rises to 60 °C after a sufficiently long induction period (~10 min), active hydrogen release begins for AB hydrolysis. It was previously shown [17] that the duration of the induction period on the hydrogen generation curve depends on the reduction rate of cobalt oxide.
So, the Co3O4 reduction or activation process is rapid in the reaction medium of a stronger reducing agent such as SBH, providing a short induction period and a high H2 generation rate. Unlike SBH and AB, H2 evolution from EDBB solution was not observed even 60 min later at 60 °C (Figure 3), demonstrating the difficulty in the formation of active phase from Co3O4 under these conditions. On the other hand, the preactivation of Co3O4 under the SBH hydrolysis allowed EDBB to be hydrolyzed at 40 °C during the next addition of EDBB to the reactor. Thus, in the row of SBH < AB < EDBB, the duration of the induction period is increased. This is a result of a decrease in the reductive ability of the studied boron-containing hydrides toward Co3O4.
It is established that modifying cobalt oxide-based catalytic systems with copper enhances their activation rate and activity in the reaction medium of AB hydrolysis, as documented in references [48,49,50,51,52]. When discussing these results, the following explanation is offered (Figure 4): Since metals with high reduction potential are more easily reduced by hydride, Cu2+ (0.337 V) is first reduced to Cu0 in the reaction media, followed by the formation of surface Cu-H centers. The high reducing ability of these centers facilitates the easier reduction in cobalt in oxide compounds. We think that the enhanced durability of the catalysts is also a result of this mechanism. Cu-H species may prevent oxidation of cobalt active component in the reaction medium.
Therefore, to improve the activation and increase the activity of Co3O4 in AB and EDBB hydrolysis at 40 °C, Co3O4 has been modified by a small amount of CuO (10 wt%). To do this, Co3O4 was ground with CuCO3·Cu(OH)2 in a hand mortar and calcined at a low temperature (300 °C) to decompose the basic copper carbonate to CuO and maintain Co3O4 dispersity. The XRD analysis confirmed that the sample was a mixture of two crystalline phases, 9% CuO + 91% Co3O4. The average size of Co3O4 crystallites (CSR) remained unchanged (41 ± 2 nm). As expected, the application of the CuO-Co3O4 catalyst in SBH hydrolysis results in a significant decrease in the induction period as well as an increase in the rate of hydrogen generation, as Figure 5a illustrates. Almost instantaneous hydrogen generation from aqueous solutions of AB and EDBB is also observed. (Figure 5b,c). The activity differences between Co0, Co3O4, and CuO-Co3O4 are thought to be explained by the various types and contents of catalytically active centers involved in the activation of reagents in the reaction medium.
As mentioned earlier, previous studies have already investigated the reduced forms of cobalt (CoBn) that are produced in situ from cobalt precursors and play a crucial role in catalyzing the studied processes [17,23,24,25]. However, the main focus of this study is on the amount and state of the remaining Co3O4 phase and its influence on the results of KIE measurements and DFT modeling of the water activation stage.

3.2.2. Study of Co3O4 Activated in the Reaction Medium of NaBH4

In order to estimate the reduction degree of Co3O4 and study its characteristics, during the stage of vigorous H2 generation, the sample of activated oxide was removed from the reaction medium of NaBH4 hydrolysis at 40 °C after 3 min by a magnet (Figure 3). The sample was then subjected to XRD examination while being analyzed under an alcohol layer to avoid its oxidation by air. The results (Table 2) indicate that the Co3O4 structure is mainly preserved and that the CoO phase content is about 5 wt%.
The study of the activated Co3O4/SBH was continued with TEM (Figure 6). Before the study, the sample was dried in a vacuum and stored under Ar. TEM revealed that this sample largely maintains the original oxide’s morphology (Figure 6a,b,d,e). However, the surface of the oxide particles has been altered by the action of hydride. The formation of translucent films and contrast rounded particles of amorphous CoxB particles is predictably detected (Figure 6d–f). Nanosized CoxB particles are observed in transparent films (Figure 6d,e) and on the surface of Co3O4 (Figure 6f). The formation of the active cobalt boride phase has been discussed previously, for example, in [17]. However, in this study, we were primarily interested in the oxide phase in activated Co3O4. For this purpose, TPR with H2, Raman, and ATR FTIR spectroscopy were further applied.
Note that in the TPR H2 study, the sample briefly came into contact with air at the stage of preparation of the experiment (weighing and loading into the reactor). Figure 7 shows that after the activation of Co3O4 in the sodium borohydride solution, there is a significant change in the TPR H2 spectrum. First of all, there is a shift of the reduction process to high temperatures at 19 °C, indicating a decrease in the reduction potential of cobalt cations in its structure.
It is well known [33] that the traditional Co3O4 reduction process is in two stages: (I) the low-temperature wide peak corresponds to the reduction phase of Co3+ to Co2+, and (II) the high-temperature narrow peak corresponds to the reduction of Co2+ to Co0 (Figure 7b). For the Co3O4 stoichiometry, the ratio of the peak area II to the peak area I must be three, which we have confirmed in our experiment for initial Co3O4. The deconvolution of the TPR H2 spectrum of the activated Co3O4/SBH (Figure 7c) is different. In this case, the higher temperature hydrogen consumption is described more precisely by the three peaks. Two high-temperature peaks (II and III) can be expected to be attributed to the reduction in two Co2+ states. For the initial Co3O4, the amount of hydrogen consumed for the reduction was 1.25 × 10−2 mol/g. This corresponds to the calculated Co3O4 reduction degree of 75%. The loss of oxygen during Co3O4 reductive treatment with SBH solution is low, as the amount of hydrogen consumed during reduction in this sample has decreased only by 9 rel%.
The increase in the Co3O4/SBH reduction temperature and the structural diversity found in its TPR H2 spectrum may be due to defects in the oxide structure. It is known that Raman spectroscopy is often used to detect defects in Co3O4. But, the Raman spectra interpretation for polydisperse Co3O4 is problematic. When analyzing them, the dimensional effects [53], degree of crystallinity [54,55], particle aggregation [56], and defects [57,58] should be taken into account. All these parameters affect the position of the peaks and their half-width. Unfortunately, when analyzing Raman spectra of Co3O4 with structural defects, in most published works, the influence of other factors is not taken into account. This appears to be a reason for numerous contradicting reports found in the literature regarding the relationship between the formation of anionic vacancies in the structure of Co3O4 and the resulting change in its Raman spectrum. There is information that this process may be accompanied by both the red shift of the A1g peak [33,58,59] and its blue shift [60,61]. In this case, the frequently mentioned phonon confinement effect should be applied to nanoscale samples (<10 nm) [62].
The obtained spectrum of initial Co3O4 has five peaks (Figure 8) at 670, 606, 513, 468, and 191 cm−1. These correspond to A1g, F2g, F2g, Eg, and F2g vibrations, respectively [63]. The peak of A1g relates to vibrations of Co3+-O in the octahedron, and the peak of F2g relates to vibrations of Co2+-O in the tetrahedron. Note that our preliminary study of the Co3O4/SBH sample by TPR H2 has shown that the reducing treatment of Co3O4 with SBH only slightly reduced the oxygen content, resulting in structural diversity of the cobalt oxygen environment and a stronger Co-O bond. XRD analysis showed that the Co3O4 crystalline phase characteristics in the Co3O4/SBH did not change, which eliminates the influence of the Co3O4 crystal size on the Raman peak position. However, the appearance of magnetic properties in the Co3O4/SBH sample due to the formation of the ferromagnetic amorphous phase of cobalt boride [17] can increase the degree of aggregation of particles, which, above all, should lead to a shift of A1g to the low-frequency region.
On the other hand, Raman spectra analysis of Co3O4/SBH shows that the A1g peak is shifted into the high-frequency region at 6 cm−1. The shape of the A1g peak and its full width at half maximum (FWHM) also change (Figure 8b). We believe that the observed asymmetry, increased FWHM, and blue shift of A1g are due to the strengthening of Co-O bonds, which is a result of the formation of anionic vacancies in the Co3O4 structure and the redistribution of electron density. Note that, according to DFT calculations, the most likely oxygen removal is from the octahedral oxygen environment.
Similar results have been reported in [60], in which the blue shift of the A1g peak in the Raman spectrum has been observed for the Co3O4 sample, characterized by an increasing hydrogen reduction temperature (TPR H2 data) and a shortening of the Co-O bond (EXAFS data). Changes in the geometry of [CoO6] octahedra must cause geometry changes in adjacent [CoO4] tetrahedra (Figure 8c). In fact, there is a noticeable broadening and shift to the low-frequency region for the F2g peak in the Raman spectrum of Co3O4/SBH, which might correspond to an elongation of the Co-O bonds in tetrahedra.
The comparative study of Co3O4 and Co3O4/SBH was continued with ATR FTIR spectroscopy. Figure 9 shows the spectra of initial Co3O4 and Co3O4/SBH, as well as initial Co3O4, on the surface of which water was preadsorbed. Adsorption was carried out in an equilibrium water-vapor system at room temperature for 20 h. It is evident that the Christiansen effect is present in all spectra. It appears at wavelengths (λChr) when the refractive indices of the matrix (air) and the sample (Co3O4) are equal. It can be observed when the bulk of the sample consists of particles that are evidently larger than the wavelength of the irradiating light (2–25 μm). It agrees with the SEM data of the studied Co3O4 (Figure S1a–c).
Figure 9a shows that the spectra of Co3O4 and Co3O4/SBH samples differ in the intensity of the broad absorption bands (a. b.) with maximums at 3230 and 2720 cm−1 corresponding to the hydroxyl groups’ valence vibration and at 1620 cm−1 corresponding to the deformation vibration of adsorbed water molecules. These a. b. are more intensive in the spectrum of Co3O4/SBH. This indicates it is impossible to thoroughly dehydrate the sample after treating Co3O4 in SBH solution, washing, and vacuum drying for a long period of time at room temperature. In the spectrum of initial Co3O4, these a. b. are barely visible. The long-time saturation of the surface of initial Co3O4 with water vapor increased the intensity of the discussed absorption bands. However, the intensity achieved is significantly lower than the intensity of vibrations of water and hydroxyl groups in the spectrum of Co3O4/SBH. This suggests that the high water content on the surface of the Co3O4/SBH sample is due to its stabilization at the anionic vacancies. In addition, the increased background absorption in the Co3O4/SBH spectrum indicates the presence of free charge carriers (CoBn) and the possible formation of proton conductivity by involving hydroxyl groups in the hydrogen bond.
A more detailed analysis of the a. b. of stretching vibration of Co-O bonds in the octahedral environment with the maximum at 656 cm−1 (Figure 9b) reveals its broadening on the low-frequency slope and a slight low-frequency shift of the maximum at 3 cm−1 (red shift). This suggests that adsorbed water weakens Co-O bonds.
The observation of different types of a. b. shifts corresponding to those for [CoO6] observed by Raman spectroscopy (blue shift) and ATR FTIR spectroscopy (red shift) is related to different conditions for receiving spectra. In the case of Raman spectroscopy, it is known that the surface of the sample is heated by the laser, resulting in adsorbed water being removed from the surface, which leads to a stronger Co-O bond. In the case of ATR FTIR spectroscopy, there is no heating of the sample, allowing us to obtain information about the Co3O4 surface containing adsorbed water on defects.
Thus, the TPR H2 study shows that the reduction degree of Co3O4 in the reaction medium of NaBH4 is not significant. XRD analysis confirms that Co3O4 remains the main crystalline phase. However, the shift of the oxide phase reduction process to higher temperatures indicates an increase in the electronic density of cobalt atoms and the formation of oxygen vacancies. These results are in accordance with the data of [31,32,33,34,35], where the Co3O4 samples treated with NaBH4 solution were examined by a set of methods such as XRD, XPS, EPR, TPR H2, TPD O2, etc. On the other hand, the vibration spectroscopy in this work confirms the differences in the structure of the nearest cobalt environment for Co3O4 and Co3O4/SBH. It is important to note that the formation of anionic vacancies on the Co3O4/SBH surface increases the amount of adsorbed water, which is difficult to remove from the surface under long vacuum drying conditions at room temperature.

3.2.3. Modeling of H2O Adsorption on the Surface of Co3O4 and Co3O4 with Oxygen Vacancies

Numerous publications have traditionally reported the formation of anionic vacancies in the structure of transition metal oxides and a decrease in the oxidation state of metals during the hydrolysis tests of SBH and AB [22,31,64,65]. However, the discussion of the role of Co3O4, including defective ones, in the water activation stage (5) is presented in single publications [36,37,38].
In order to study the structural changes in Co3O4 during the formation of anionic vacancies, DFT calculations were performed. When studying the 311 surface of Co3O4 (Fm3m), two positions for oxygen vacancies were considered (Figure 10): Co3+-O-Co2+ (Vo1 –yellow) and Co3+-O-Co3+ (Vo2-green). Our calculations showed that the formation of Vo1 required more energy (2.65 eV) than the formation of Vo2 (2.10 eV). This suggests that anionic vacancies are more likely to be in an octahedral Co3+ environment than in a tetrahedral Co2+ environment.
The optimization of oxide cluster structures has shown that vacancies cause shorter Co-O lengths in the cobalt octahedron environment near the vacancy. When Vo1 and Vo2 appear, the total length of Co-O bonds near the vacancy decreases by 1.8 and 4.7 rel%, respectively, which confirms the blue shift of A1g in the Raman spectrum of Co3O4/SBH (Figure 8b). It is noted that the oxygen atom 2 is moved to the octahedron near the vacancy after the creation of the Vo2 vacancy (Figure 10c). This results in a 3.62 rel% increase in Co-O bond lengths in the nearest tetrahedron and is consistent with the results of Raman spectroscopy on red shift of F2g in the spectrum of Co3O4/SBH (Figure 8c).
As expected, the formation of anionic vacancies in the Co3O4 structure leads to an increase in the electron density on the Co atoms adjacent to the vacancy (Figure 11), which reduces its oxidation state. A similar result was discussed in [66]. As a result, the mobility of lattice oxygen is decreased, which explains the observed increase in the reduction temperature in the TPR H2 spectrum of Co3O4/SBH (Figure 7a), as in the TPR H2 spectra of other similar samples [33,60].
Calculations have confirmed that creating anionic vacancies and increasing the electron density on cobalt atoms enhances the adsorption of water molecules. The adsorption energy decreases from −0.208 eV for Co3O4 to −0.653 eV and −0.432 eV for Co3O4 + Vo1 and Co3O4 + Vo2, respectively. Oxygen vacancies also positively affect the stage of activation of water. The activation energy of O-H breakdown on the Co3O4 surface with vacancies Vo1 (0.248 eV) and Vo2 (0.395 eV) is less than that in the case of initial Co3O4 (0.620 eV) (Figure 12). These results are consistent with the literature [37,38]. It is shown that the OH* group formed during the dissociation of water occupies the position of absent lattice oxygen if this is sterically available. At the same time, there is an increase in the total length of Co-O bonds by 0.99 rel% and 0.11 rel% near Vo1 and Vo2, respectively. This corresponds to the results of ATR FTIR spectroscopy (Figure 9) on the weakening of Co-O bonds in [CoO6] in the case of the Co3O4/SBH containing the adsorbed water.
Similar to metals (Figure 1b), DFT calculations on cobalt oxide show a decrease in activation energy with a reduction in adsorption energy. There is a linear correlation between these values (Figure 13): the lower the Eads of water, the lower the Ea for the O-H bond dissociation. For instance, similar conclusions were obtained in [64] to explain the high activity of Cu0/Cu0.76Co2.24O4−δ in AB hydrolysis. It should be mentioned that the linear Ea-Eads relationships for metals and oxides do not coincide (Figure 13). Metals have higher activation energy values. These results correspond with known statements that the dissociation of water on the polarized oxide surface occurs with lower activation energies than on metal surfaces [36,67]. The comparison of the DFT results of this work with literature data for different metal oxide surfaces is presented in Figure S3b.
Thus, based on the published literature and the results of this study, it can be concluded that the high-activity state of the cobalt oxide catalyst can be considered as CoBn active centers immobilized in the oxide matrix with anionic vacancies (Co3O4−δ). It can be assumed that the oxide phase serves not only as a precursor and support for the metal or metalloid cobalt active component formed during the reduction process but also as a key catalyst component that improves the water activation process. If so, this suggests that when H2O is replaced with D2O in the catalytic hydrolysis of the boron-containing hydrides, a Co3O4-based catalyst is to be characterized with a different kinetic isotope effect compared to a metal catalyst.

3.2.4. Comparison of Kinetic Isotope Effect Results for Co3O4

It should be noted that the kinetic isotope effect (KIE) in the hydrolysis of boron-containing hydrides measured at replacing H2O with D2O for transition metal oxide-based catalysts is almost nonpublished in the literature. The results obtained in our work on the KIE in SBH hydrolysis in the presence of Co3O4 are presented in Figure 14.
Four successive reaction cycles were carried out. First of all, it is evident that the use of D2O almost four times lengthens the induction period. It indicates that D2O is directly involved as a reagent when reducing oxide in a reaction medium takes place. When calculating the KIE, the induction period was not taken into account. Only hydrogen evolution kinetics after the activation stage was analyzed. The kH/kD values for four consecutive tests are shown in Figure 14. It can be seen that the H2 generation rate in the presence of Co3O4 is only slightly reduced when using D2O. The kH/kD value is changed little during cyclic testing, indicating the relative stability of the active component in the activated oxide catalyst. It can be assumed that the reduction degree of Co3O4 changes little after the activation stage.
This result is consistent with the pioneer study published in 2022 [68]. In this work, nickel and cobalt oxides have been shown to have a relatively low reduction rate in the reaction medium since H* formed by the dissociation of B-H, after forming the catalytically active cents on activation reagents in sufficient quantity, is not spent on oxide reduction and reacts mainly with H*, formed during water dissociation, with hydrogen release.
Similar results were obtained in the hydrolysis of AB and EDBB over Co3O4 modified by CuO (Figure S4). Table 3 shows that when Co3O4 is tested in SBH hydrolysis as well as 10% CuO-90% Co3O4 in AB and EDBB hydrolysis, kH/kD < 2 (the secondary kinetic isotope effect) is obtained. Note that in the experiments with the metallic Co0 catalyst, the primary isotope effect has been fixed (kH/kD > 2; see Section 3.1.2).
Thus, the activation of water on the in situ-activated Co3O4 catalyst surface is facilitated. This is confirmed not only by KIE measurements but also by DFT modeling. Based on the discussions mentioned earlier, activated Co3O4 catalysts should be considered as nanosized CoBn immobilized in the anion-deficient Co3O4−δ. If the Co3O4−δ phase was not involved in the water activation process, the kH/kD values would be more than two.

4. Conclusions

In the hydrolysis processes of boron-containing hydrides like SBH and AB, the roles of Co3O4 and Co0 have previously been examined, with these findings discussed in the introduction. The new study of them is explained by the need to clarify their catalytic properties from the modern position of a bifunctional catalyst with centers activating hydride and water. It was also interesting to consider the results for SBH and AB hydrolysis together with the results for EDBB hydrolysis, where the use of such catalysts had not yet been described.
DFT calculations for the dissociative adsorption of water and AB on metal clusters Co, Ni, and Cu showed an implementation of the Brønsted–Evans–Polanyi principle and confirmed that the water activation stage is characterized by a higher activation energy. For this stage, a linear correlation between Eads and Ea (the Sabatier principle) was found among the metals studied. Cobalt has the lowest Eads and Ea, which is supported by numerous literature data on the high activity of cobalt-containing active components. The study of the kinetic isotope effect by replacing H2O with D2O confirmed that in the case of Co0, the primary isotopic effect is found, i.e., the breaking of the OH-bond in the water molecule determines the rate of hydrogen generation and can be considered as a rate-determining stage.
It is known that Co3O4 is activated in a reaction medium of hydrides, and the in situ formed active phase begins to catalyze the hydrolysis process of boron-containing hydrides. It was shown that in the row of SBH > AB > EDBB, Co3O4 activity decreases. This is a result of a decrease in the reductive ability of the studied boron-containing hydrides toward Co3O4. In the case of AB and EDBB hydrolysis, this problem can be solved by adding copper compounds to the cobalt oxide catalysts. According to our study (TEM, TPR H2, Raman, and ATR FTIR spectroscopy), the reduction degree of Co3O4 in the reaction medium of SBH with a high reducing ability is low, and the active state of the oxide catalyst providing a high H2 generation rate should be considered as nanodispersed CoBn immobilized in the anion-deficient Co3O4−δ. According to ATR FTIR spectroscopy, the presence of defects on the surface of Co3O4 enhances water adsorption.
DFT modeling shows that the formation of an anion-deficient structure of Co3O4 leads to an increase in the electron density on Co atoms near the vacancies. The presence of anionic vacancies on the oxide surface facilitates the adsorption and activation of water. The localization of oxygen defects in the oxide structure determines the values of Eads and Ea. According to the Sabatier principle, these parameters are also related to a linear dependence. As expected, the Ea value for the dissociative adsorption of water on the oxide surface is significantly lower than on the metal surface. KIE measurements for Co3O4 in SBH hydrolysis and 10% CuO-90% Co3O4 in SBH, AB, and EDBB hydrolysis reveal a secondary KIE, indicating that the activation of water is not part of RDS. The data obtained support the statement that the creation of anionic vacancies in the structure of catalysts is a standard technique to improve the transformation of reagents containing an electron-donor atom.
Although the activity of the Co0 catalyst depends on water activation, the hydrogen generation rate is high. We believe this may be attributed to the relatively easier process of hydride activation. To enhance the water activation processes, we propose the addition of activating water oxide to the catalytic composition containing Co0. Oxides can serve not only as supports but also as water activation components. When developing a catalyst, finding the optimum proportion of metal/metalloid to oxide is crucial. This allows one to not only stabilize and modify the active metal-like component but also introduce the requisite number of water activation centers.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17081794/s1, Description of ammonia borane and ethylenediamine bisborane synthesis; Table S1: Kinetic isotope effect (KIE) in AB hydrolysis when replacing H2O by D2O; Figure S1: SEM images of studied catalytic materials: (a–c) Co3O4, (d–f) 10%CuO-90%Co3O4, and (g–i) Co0; Figure S2: Example of optimized structures of initial and final states for dissociative adsorption of (a,b) NH3BH3 (4) and (c,d) H2O (5) on 111 Co0. The atoms of Co are pink, B are peach, N are blue, H are grey, and O are red; Figure S3: For metals (a) and oxides (b), comparison of the energetic parameters (Ea, Eads) calculated by DFT in this work with the literature data; Figure S4: Experimental curves of H2 evolution for hydrolysis of NaBH4, NH3BH3, and (CH2NH2BH3)2 measured at replacement of H2O by D2O over 10% CuO-90% Co3O4 catalyst at 40 °C [26,36,40,43,46,47,64,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89].

Author Contributions

Writing—review and editing, O.V.K. and O.V.N.; conceptualization, V.I.S. and O.V.K.; investigation, V.R.B., I.L.L., V.A.R., G.V.O., O.A.B. and Y.A.C., validation, A.M.O. and N.A.D.; software, V.R.B.; visualization, I.L.L.; formal analysis, A.M.O. and N.A.D.; writing—original draft preparation, V.A.R., G.V.O., O.A.B. and Y.A.C.; data curation, V.I.S.; project administration, 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 FWUR-2024-0034).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors are very grateful to A.V. Ishchenko for the study of the samples by transmission electron microscopy (TEM). Raman spectra were measured using the equipment of the Center for Collective Use “VTAN” of the Analytical and Technological Research Center of the Department of Physics of NSU.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Baum, Z.J.; Diaz, L.L.; Konovalova, T.; Zhou, Q.A. Materials Research Directions Toward a Green Hydrogen Economy: A Review. ACS Omega 2022, 7, 32908–32935. [Google Scholar] [CrossRef] [PubMed]
  2. Vivanco-Martín, B.; Iranzo, A. Analysis of the European Strategy for Hydrogen: A Comprehensive Review. Energies 2023, 16, 3866. [Google Scholar] [CrossRef]
  3. Salman, M.S.; Pratthana, C.; Lai, Q.; Wang, T.; Rambhujun, N.; Srivastava, K.; Aguey-Zinsou, K.F. Catalysis in Solid Hydrogen Storage: Recent Advances, Challenges, and Perspectives. Energy Technol. 2022, 10, 2200433. [Google Scholar] [CrossRef]
  4. Luconi, L.; Tuci, G.; Giambastiani, G.; Rossin, A.; Peruzzini, M. H2 Production from Lightweight Inorganic Hydrides Catalyzed by 3d Transition Metals. Int. J. Hydrogen Energy 2019, 44, 25746–25776. [Google Scholar] [CrossRef]
  5. Abdelhamid, H.N. A Review on Hydrogen Generation from the Hydrolysis of Sodium Borohydride. Int. J. Hydrogen Energy 2021, 46, 726–765. [Google Scholar] [CrossRef]
  6. Guan, S.; Liu, Y.; Zhang, H.; Shen, R.; Wen, H.; Kang, N.; Zhou, J.; Liu, B.; Fan, Y.; Jiang, J.; et al. Recent Advances and Perspectives on Supported Catalysts for Heterogeneous Hydrogen Production from Ammonia Borane. Adv. Sci. 2023, 10, 2300726. [Google Scholar] [CrossRef]
  7. Wang, M.; Wang, J.; Zhao, X.; Qin, G.; Zhang, X.; Lu, Z.; Yu, X.; Li, L.; Yang, X. Hydrogen Production from Ammonia Borane Hydrolysis Catalyzed by Metal Oxide-Based Materials. Russ. J. Phys. Chem. A 2023, 97, 854–877. [Google Scholar] [CrossRef]
  8. Engin, M.; Ozay, O. The First Catalytic Hydrolysis of Ethylenediamine Bisborane with Hydrogel-Supported Metallic Nanoparticles. Int. J. Hydrogen Energy 2018, 43, 15083–15094. [Google Scholar] [CrossRef]
  9. Coşkuner, Ö.; Kantürk Figen, A. Catalytic Semi-Continuous Operation Modes for Hydrogen Generation from Carbon Derivatives of Ammonia Boranes. Int. J. Hydrogen Energy 2022, 47, 40304–40316. [Google Scholar] [CrossRef]
  10. Yao, J.; Wu, Z.; Wang, H.; Yang, F.; Ren, J.; Zhang, Z. Application-Oriented Hydrolysis Reaction System of Solid-State Hydrogen Storage Materials for High Energy Density Target: A Review. J. Energy Chem. 2022, 74, 218–238. [Google Scholar] [CrossRef]
  11. Chen, W.; Ouyang, L.Z.; Liu, J.W.; Yao, X.D.; Wang, H.; Liu, Z.W.; Zhu, M. Hydrolysis and Regeneration of Sodium Borohydride (NaBH4)—A Combination of Hydrogen Production and Storage. J. Power Sources 2017, 359, 400–407. [Google Scholar] [CrossRef]
  12. Xu, Q.; Chandra, M. A Portable Hydrogen Generation System: Catalytic Hydrolysis of Ammonia-Borane. J. Alloys Compd. 2007, 446–447, 729–732. [Google Scholar] [CrossRef]
  13. 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]
  14. Keskin, E.; Coşkuner Filiz, B.; Kılıç Depren, S.; Kantürk Figen, A. Recommendations for Ammonia Borane Composite Pellets as a Hydrogen Storage Medium. Int. J. Hydrogen Energy 2018, 43, 20354–20371. [Google Scholar] [CrossRef]
  15. Komova, O.V.; Simagina, V.I.; Pochtar, A.A.; Bulavchenko, O.A.; Ishchenko, A.V.; Odegova, G.V.; Gorlova, A.M.; Ozerova, A.M.; Lipatnikova, I.L.; Tayban, E.S.; et al. Catalytic Behavior of Iron-Containing Cubic Spinel in the Hydrolysis and Hydrothermolysis of Ammonia Borane. Materials 2021, 14, 5422. [Google Scholar] [CrossRef]
  16. Coşkuner, Ö.; Kantürk Figen, A. Hydro-Catalytic Treatment of Organoamine Boranes for Efficient Thermal Dehydrogenation for Hydrogen Production. Int. J. Hydrogen Energy 2021, 46, 35641–35652. [Google Scholar] [CrossRef]
  17. 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]
  18. 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. Hydrogen Energy 2008, 33, 7095–7102. [Google Scholar] [CrossRef]
  19. Tomboc, G.R.M.; Tamboli, A.H.; Kim, H. Synthesis of Co3O4 Macrocubes Catalyst Using Novel Chitosan/Urea Template for Hydrogen Generation from Sodium Borohydride. Energy 2017, 121, 238–245. [Google Scholar] [CrossRef]
  20. Pfeil, T.L.; Pourpoint, T.L.; Groven, L.J. Effects of Crystallinity and Morphology of Solution Combustion Synthesized Co3O4 as a Catalyst Precursor in Hydrolysis of Sodium Borohydride. Int. J. Hydrogen Energy 2014, 39, 2149–2159. [Google Scholar] [CrossRef]
  21. Zhang, X.; Zhang, Q.; Xu, B.; Liu, X.; Zhang, K.; Fan, G.; Jiang, W. Efficient Hydrogen Generation from the NaBH4 Hydrolysis by Cobalt-Based Catalysts: Positive Roles of Sulfur-Containing Salts. ACS Appl. Mater. Interfaces 2020, 12, 9376–9386. [Google Scholar] [CrossRef] [PubMed]
  22. Ugale, A.D.; Ghodke, N.P.; Kang, G.S.; Nam, K.B.; Bhoraskar, S.V.; Mathe, V.L.; Yoo, J.B. Cost-Effective Synthesis of Carbon Loaded Co3O4 for Controlled Hydrogen Generation via NaBH4 Hydrolysis. Int. J. Hydrogen Energy 2022, 47, 16–29. [Google Scholar] [CrossRef]
  23. 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. [Google Scholar] [CrossRef]
  24. 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]
  25. 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]
  26. Chen, W.; Fu, W.; Qian, G.; Zhang, B.; Chen, D.; Duan, X.; Zhou, X. Synergistic Pt-WO3 Dual Active Sites to Boost Hydrogen Production from Ammonia Borane. iScience 2020, 23, 100922. [Google Scholar] [CrossRef]
  27. Xu, W.; Zhang, S.; Shen, R.; Peng, Z.; Liu, B.; Li, J.; Zhang, Z.; Li, B. A Catalytic Copper/Cobalt Oxide Interface for Efficient Hydrogen Generation. Energy Environ. Mater. 2023, 6, e12279. [Google Scholar] [CrossRef]
  28. Mo, B.; Li, S.; Wen, H.; Zhang, H.; Zhang, H.; Wu, J.; Li, B.; Hou, H. Functional Group Regulated Ni/Ti3C2Tx (Tx = F, -OH) Holding Bimolecular Activation Tunnel for Enhanced Ammonia Borane Hydrolysis. ACS Appl. Mater. Interfaces 2022, 14, 16320–16329. [Google Scholar] [CrossRef]
  29. Yao, F.; Guan, S.; Bian, L.; Fan, Y.; Liu, X.; Zhang, H.; Li, B.; Liu, B. Ensemble-Exciting Effect in Pd/Alk-Ti3C2 on the Activity for Efficient Hydrogen Production. ACS Sustain. Chem. Eng. 2021, 9, 12332–12340. [Google Scholar] [CrossRef]
  30. Karakaya Akbaş, N.; Kutlu, B. Effect of Hydroxyl (OH) Radicals on the Progression of NaBH4 Hydrolysis Reaction on Fcc-Co Surfaces: A DFT Study. Phys. B Condens. Matter. 2022, 647, 414385. [Google Scholar] [CrossRef]
  31. Ji, J.; Deng, K.; Li, J.; Zhang, Z.; Duan, X.; Huang, H. In Situ Transformation of 3D Co3O4 Nanoparticles to 2D Nanosheets with Rich Surface Oxygen Vacancies to Boost Hydrogen Generation from NaBH4. Chem. Eng. J. 2021, 424, 130350. [Google Scholar] [CrossRef]
  32. Yuan, H.; Wang, S.; Ma, Z.; Kundu, M.; Tang, B.; Li, J.; Wang, X. Oxygen Vacancies Engineered Self-Supported B Doped Co3O4 Nanowires as an Efficient Multifunctional Catalyst for Electrochemical Water Splitting and Hydrolysis of Sodium Borohydride. Chem. Eng. J. 2021, 404, 126474. [Google Scholar] [CrossRef]
  33. Wang, X.; Li, X.; Mu, J.; Fan, S.; Chen, X.; Wang, L.; Yin, Z.; Tadé, M.; Liu, S. Oxygen Vacancy-Rich Porous Co3O4 Nanosheets toward Boosted NO Reduction by CO and CO Oxidation: Insights into the Structure-Activity Relationship and Performance Enhancement Mechanism. ACS Appl. Mater. Interfaces 2019, 11, 41988–41999. [Google Scholar] [CrossRef] [PubMed]
  34. Xiang, K.; Xu, Z.; Qu, T.; Tian, Z.; Zhang, Y.; Wang, Y.; Xie, M.; Guo, X.; Ding, W.; Guo, X. Two Dimensional Oxygen-Vacancy-Rich Co3O4 Nanosheets with Excellent Supercapacitor Performances. Chem. Commun. 2017, 53, 12410–12413. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, Y.; Zhou, T.; Jiang, K.; Da, P.; Peng, Z.; Tang, J.; Kong, B.; Cai, W.B.; Yang, Z.; Zheng, G. Reduced Mesoporous Co3O4 Nanowires as Efficient Water Oxidation Electrocatalysts and Supercapacitor Electrodes. Adv. Energy Mater. 2014, 4, 1400696. [Google Scholar] [CrossRef]
  36. Guan, S.; An, L.; Ashraf, S.; Zhang, L.; Liu, B.; Fan, Y.; Li, B. Oxygen Vacancy Excites Co3O4 Nanocrystals Embedded into Carbon Nitride for Accelerated Hydrogen Generation. Appl. Catal. B Environ. 2020, 269, 118775. [Google Scholar] [CrossRef]
  37. Dong, G.; Hu, H.; Huang, X.; Zhang, Y.; Bi, Y. Rapid Activation of Co3O4 Cocatalysts with Oxygen Vacancies on TiO2 Photoanodes for Efficient Water Splitting. J. Mater. Chem. A 2018, 6, 21003–21009. [Google Scholar] [CrossRef]
  38. Lu, Y.; Li, C.; Zhang, Y.; Cao, X.; Xie, G.; Wang, M.; Peng, D.; Huang, K.; Zhang, B.; Wang, T.; et al. Engineering of Cation and Anion Vacancies in Co3O4 Thin Nanosheets by Laser Irradiation for More Advancement of Oxygen Evolution Reaction. Nano Energy 2021, 83, 105800. [Google Scholar] [CrossRef]
  39. Ozerova, A.M.; Skobelkina, A.A.; Simagina, V.I.; Komova, O.V.; Prosvirin, I.P.; Bulavchenko, O.A.; Lipatnikova, I.L.; Netskina, O.V. Magnetically Recovered Co and Co@Pt Catalysts Prepared by Galvanic Replacement on Aluminum Powder for Hydrolysis of Sodium Borohydride. Materials 2022, 15, 3010. [Google Scholar] [CrossRef]
  40. Komova, O.V.; Simagina, V.I.; Butenko, V.R.; Odegova, G.V.; Bulavchenko, O.A.; Nikolaeva, O.A.; Ozerova, A.M.; Lipatnikova, I.L.; Tayban, E.S.; Mukha, S.A.; et al. Dehydrogenation of Ammonia Borane Recrystallized by Different Techniques. Renew. Energy 2022, 184, 460–472. [Google Scholar] [CrossRef]
  41. Tang, W.; Sanville, E.; Henkelman, G. A Grid-Based Bader Analysis Algorithm without Lattice Bias. J. Phys. Condens. Matter 2009, 21, 084204. [Google Scholar] [CrossRef] [PubMed]
  42. Bligaard, T.; Nørskov, J.K.; Dahl, S.; Matthiesen, J.; Christensen, C.H.; Sehested, J. The Brønsted-Evans-Polanyi Relation and the Volcano Curve in Heterogeneous Catalysis. J. Catal. 2004, 224, 206–217. [Google Scholar] [CrossRef]
  43. Wang, C.; Tuninetti, J.; Wang, Z.; Zhang, C.; Ciganda, R.; Salmon, L.; Moya, S.; Ruiz, J.; Astruc, D. Hydrolysis of Ammonia-Borane over Ni/ZIF-8 Nanocatalyst: High Efficiency, Mechanism, and Controlled Hydrogen Release. J. Am. Chem. Soc. 2017, 139, 11610–11615. [Google Scholar] [CrossRef] [PubMed]
  44. Medford, A.J.; Vojvodic, A.; Hummelshøj, J.S.; Voss, J.; Abild-Pedersen, F.; Studt, F.; Bligaard, T.; Nilsson, A.; Nørskov, J.K. From the Sabatier Principle to a Predictive Theory of Transition-Metal Heterogeneous Catalysis. J. Catal. 2015, 328, 36–42. [Google Scholar] [CrossRef]
  45. Demirci, U.B.; Miele, P. Cobalt in NaBH4 Hydrolysis. Phys. Chem. Chem. Phys. 2010, 12, 14651–14665. [Google Scholar] [CrossRef]
  46. Li, Z.; He, T.; Liu, L.; Chen, W.; Zhang, M.; Wu, G.; Chen, P. Covalent Triazine Framework Supported Non-Noble Metal Nanoparticles with Superior Activity for Catalytic Hydrolysis of Ammonia Borane: From Mechanistic Study to Catalyst Design. Chem. Sci. 2016, 8, 781–788. [Google Scholar] [CrossRef] [PubMed]
  47. Paterson, R.; Alharbi, A.A.; Wills, C.; Dixon, C.; Šiller, L.; Chamberlain, T.W.; Griffiths, A.; Collins, S.M.; Wu, K.; Simmons, M.D.; et al. Heteroatom Modified Polymer Immobilized Ionic Liquid Stabilized Ruthenium Nanoparticles: Efficient Catalysts for the Hydrolytic Evolution of Hydrogen from Sodium Borohydride. Mol. Catal. 2022, 528, 112476. [Google Scholar] [CrossRef]
  48. Feng, Y.; Liao, J.; Chen, X.; Wang, H.; Guo, B.; Li, H.; Zhou, L.; Huang, J.; Li, H. Synthesis of Rattle-Structured CuCo2O4 Nanospheres with Tunable Sizes Based on Heterogeneous Contraction and Their Ultrahigh Performance toward Ammonia Borane Hydrolysis. J. Alloys Compd. 2021, 863, 158089. [Google Scholar] [CrossRef]
  49. Liu, Q.; Zhang, S.; Liao, J.; Feng, K.; Zheng, Y.; Pollet, B.G.; Li, H. CuCo2O4 Nanoplate Film as a Low-Cost, Highly Active and Durable Catalyst towards the Hydrolytic Dehydrogenation of Ammonia Borane for Hydrogen Production. J. Power Sources 2017, 355, 191–198. [Google Scholar] [CrossRef]
  50. Lu, D.; Liao, J.; Zhong, S.; Leng, Y.; Ji, S.; Wang, H.; Wang, R.; Li, H. Cu0.6Ni0.4Co2O4 Nanowires, a Novel Noble-Metal-Free Catalyst with Ultrahigh Catalytic Activity towards the Hydrolysis of Ammonia Borane for Hydrogen Production. Int. J. Hydrogen Energy 2018, 43, 5541–5550. [Google Scholar] [CrossRef]
  51. Lu, D.; Li, J.; Lin, C.; Liao, J.; Feng, Y.; Ding, Z.; Li, Z.; Liu, Q.; Li, H. A Simple and Scalable Route to Synthesize CoxCu1−xCo2O4@CoyCu1−yCo2O4 Yolk–Shell Microspheres, A High-Performance Catalyst to Hydrolyze Ammonia Borane for Hydrogen Production. Small 2019, 15, 1805460. [Google Scholar] [CrossRef]
  52. Lu, D.; Liao, J.; Leng, Y.; Zhong, S.; He, J.; Wang, H.; Wang, R.; Li, H. Mo-Doped Cu0.5Ni0.5Co2O4 Nanowires, a Strong Substitute for Noble-Metal-Based Catalysts towards the Hydrolysis of Ammonia Borane for Hydrogen Production. Catal. Commun. 2018, 114, 89–92. [Google Scholar] [CrossRef]
  53. Gupta, S.K.; Desai, R.; Jha, P.K.; Sahoo, S.; Kirin, D. Titanium Dioxide Synthesized Using Titanium Chloride: Size Effect Study Using Raman Spectroscopy and Photoluminescence. J. Raman Spectrosc. 2010, 41, 350–355. [Google Scholar] [CrossRef]
  54. Gawali, S.R.; Gandhi, A.C.; Gaikwad, S.S.; Pant, J.; Chan, T.S.; Cheng, C.L.; Ma, Y.R.; Wu, S.Y. Role of Cobalt Cations in Short Range Antiferromagnetic Co3O4 Nanoparticles: A Thermal Treatment Approach to Affecting Phonon and Magnetic Properties. Sci. Rep. 2018, 8, 249. [Google Scholar] [CrossRef]
  55. Zhang, R.; Zhang, Y.C.; Pan, L.; Shen, G.Q.; Mahmood, N.; Ma, Y.H.; Shi, Y.; Jia, W.; Wang, L.; Zhang, X.; et al. Engineering Cobalt Defects in Cobalt Oxide for Highly Efficient Electrocatalytic Oxygen Evolution. ACS Catal. 2018, 8, 3803–3811. [Google Scholar] [CrossRef]
  56. Lorite, I.; Del Campo, A.; Romero, J.J.; Fernández, J.F. Isolated Nanoparticle Raman Spectroscopy. J. Raman Spectrosc. 2012, 43, 889–894. [Google Scholar] [CrossRef]
  57. Pasquini, C.; D’Amario, L.; Zaharieva, I.; Dau, H. Operando Raman Spectroscopy Tracks Oxidation-State Changes in an Amorphous Co Oxide Material for Electrocatalysis of the Oxygen Evolution Reaction. J. Chem. Phys. 2020, 152, 194202. [Google Scholar] [CrossRef]
  58. Zhang, R.; Pan, L.; Guo, B.; Huang, Z.F.; Chen, Z.; Wang, L.; Zhang, X.; Guo, Z.; Xu, W.; Loh, K.P.; et al. Tracking the Role of Defect Types in Co3O4 Structural Evolution and Active Motifs during Oxygen Evolution Reaction. J. Am. Chem. Soc. 2023, 145, 2271–2281. [Google Scholar] [CrossRef]
  59. Zhao, L.; Zhang, J.; Zhang, Z.; Wei, T.; Wang, J.; Ma, J.; Ren, Y.; Zhang, H. Co3O4 Crystal Plane Regulation to Efficiently Activate Peroxymonosulfate in Water: The Role of Oxygen Vacancies. J. Colloid Interface Sci. 2022, 623, 520–531. [Google Scholar] [CrossRef]
  60. Mo, S.; Zhang, Q.; Li, S.; Ren, Q.; Zhang, M.; Xue, Y.; Peng, R.; Xiao, H.; Chen, Y.; Ye, D. Integrated Cobalt Oxide Based Nanoarray Catalysts with Hierarchical Architectures: In Situ Raman Spectroscopy Investigation on the Carbon Monoxide Reaction Mechanism. ChemCatChem 2018, 10, 3012–3026. [Google Scholar] [CrossRef]
  61. Wang, Z.; Wang, W.; Zhang, L.; Jiang, D. Surface Oxygen Vacancies on Co3O4 Mediated Catalytic Formaldehyde Oxidation at Room Temperature. Catal. Sci. Technol. 2016, 6, 3845–3853. [Google Scholar] [CrossRef]
  62. Arora, A.K.; Rajalakshmi, M.; Ravindran, T.R.; Sivasubramanian, V. Raman spectroscopy of optical phonon confinement in nanostructured materials. J. Raman Spectrosc. 2007, 38, 604–617. [Google Scholar] [CrossRef]
  63. Hadjiev, V.G.; Iliev, M.N.; Vergilov, I.V. The Raman Spectra of Co3O4. J. Phys. C Solid State Phys. 1988, 21, L199–L201. [Google Scholar] [CrossRef]
  64. Wang, C.; Ren, Y.; Zhao, J.; Sun, S.; Du, X.; Wang, M.; Ma, G.; Yu, H.; Li, L.; Yu, X.; et al. Oxygen Vacancy-Attired Dual-Active-Sites Cu/Cu0.76Co2.24O4 Drives Electron Transfer for Efficient Ammonia Borane Dehydrogenation. Appl. Catal. B Environ. 2022, 314, 121494. [Google Scholar] [CrossRef]
  65. Patil, K.N.; Prasad, D.; Bhanushali, J.T.; Kim, H.; Atar, A.B.; Nagaraja, B.M.; Jadhav, A.H. Sustainable Hydrogen Generation by Catalytic Hydrolysis of NaBH4 Using Tailored Nanostructured Urchin-like CuCo2O4 Spinel Catalyst. Catal. Lett. 2020, 150, 586–604. [Google Scholar] [CrossRef]
  66. Yang, S.; Liu, Y.; Hao, Y.; Yang, X.; Goddard, W.A.; Zhang, X.L.; Cao, B. Oxygen-Vacancy Abundant Ultrafine Co3O4/Graphene Composites for High-Rate Supercapacitor Electrodes. Adv. Sci. 2018, 5, 1700659. [Google Scholar] [CrossRef]
  67. Zhang, H.; Zhang, K.; Ashraf, S.; Fan, Y.; Guan, S.; Wu, X.; Liu, Y.; Liu, B.; Li, B. Polar O–Co–P Surface for Bimolecular Activation in Catalytic Hydrogen Generation. Energy Environ. Mater. 2023, 6, e12273. [Google Scholar] [CrossRef]
  68. Gao, Z.; Wang, G.; Lei, T.; Lv, Z.; Xiong, M.; Wang, L.; Xing, S.; Ma, J.; Jiang, Z.; Qin, Y. Enhanced Hydrogen Generation by Reverse Spillover Effects over Bicomponent Catalysts. Nat. Commun. 2022, 13, 118. [Google Scholar] [CrossRef]
  69. Gorlova, A.M.; Kayl, N.L.; Komova, O.V.; Netskina, O.V.; Ozerova, A.M.; Odegova, G.V.; Bulavchenko, O.A.; Ishchenko, A.V.; Simagina, V.I. Fast Hydrogen Generation from Solid NH3BH3 under Moderate Heating and Supplying a Limited Quantity of CoCl2 or NiCl2 Solution. Renew. Energy 2018, 121, 722–729. [Google Scholar] [CrossRef]
  70. Rueda, M.; Sanz-Moral, L.M.; Segovia, J.J.; Martín, Á. Enhancement of Hydrogen Release Kinetics from Ethane 1,2 Diamineborane (EDAB) by Micronization Using Supercritical Antisolvent (SAS) Precipitation. Chem. Eng. J. 2016, 306, 164–173. [Google Scholar] [CrossRef]
  71. Leardini, F.; Valero-Pedraza, M.J.; Perez-Mayoral, E.; Cantelli, R.; Bañares, M.A. Thermolytic Decomposition of Ethane 1,2-Diamineborane Investigated by Thermoanalytical Methods and in Situ Vibrational Spectroscopy. J. Phys. Chem. C 2014, 118, 17221–17230. [Google Scholar] [CrossRef]
  72. Li, H.; He, W.; Xu, L.; Pan, Y.; Xu, R.; Sun, Z.; Wei, S. Synergistic Interface between Metal Cu Nanoparticles and CoO for Highly Efficient Hydrogen Production from Ammonia Borane. RSC Adv. 2023, 13, 11569–11576. [Google Scholar] [CrossRef]
  73. Fu, F.; Wang, C.; Wang, Q.; Martinez-Villacorta, A.M.; Escobar, A.; Chong, H.; Wang, X.; Moya, S.; Salmon, L.; Fouquet, E.; et al. Highly Selective and Sharp Volcano-Type Synergistic Ni2Pt@ZIF-8-Catalyzed Hydrogen Evolution from Ammonia Borane Hydrolysis. J. Am. Chem. Soc. 2018, 140, 10034–10042. [Google Scholar] [CrossRef]
  74. Wang, Q.; Fu, F.; Escobar, A.; Moya, S.; Ruiz, J.; Astruc, D. “Click” Dendrimer-Stabilized Nanocatalysts for Efficient Hydrogen Release upon Ammonia-Borane Hydrolysis. ChemCatChem 2018, 10, 2673–2680. [Google Scholar] [CrossRef]
  75. Zhao, Q.; Espuche, B.; Kang, N.; Moya, S.; Astruc, D. Cobalt Sandwich-Stabilized Rhodium Nanocatalysts for Ammonia Borane and Tetrahydroxydiboron Hydrolysis. Inorg. Chem. Front. 2022, 9, 4651–4660. [Google Scholar] [CrossRef]
  76. Kang, N.; Wang, Q.; Djeda, R.; Wang, W.; Fu, F.; Moro, M.M.; Ramirez, M.D.L.A.; Moya, S.; Coy, E.; Salmon, L.; et al. Visible-Light Acceleration of H2 Evolution from Aqueous Solutions of Inorganic Hydrides Catalyzed by Gold-Transition-Metal Nanoalloys. ACS Appl. Mater. Interfaces 2020, 12, 53816–53826. [Google Scholar] [CrossRef]
  77. Meng, Y.; Sun, Q.; Zhang, T.; Zhang, J.; Dong, Z.; Ma, Y.; Wu, Z.; Wang, H.; Bao, X.; Sun, Q.; et al. Cobalt-Promoted Noble-Metal Catalysts for Efficient Hydrogen Generation from Ammonia Borane Hydrolysis. J. Am. Chem. Soc. 2023, 145, 5486–5495. [Google Scholar] [CrossRef]
  78. Song, J.; Gu, X.; Zhang, H. Electrons and Hydroxyl Radicals Synergistically Boost the Catalytic Hydrogen Evolution from Ammonia Borane over Single Nickel Phosphides under Visible Light Irradiation. ChemistryOpen 2020, 9, 366–373. [Google Scholar] [CrossRef]
  79. Yang, L.; Liu, Z.; Qu, B.; Tao, Y.; Liu, Y. Highly Efficient Dehydrogenation of Ammonia Borane over Reduced Graphene Oxide-Supported Pd@NiP Nanoparticles at Room Temperature. Int. J. Energy Res. 2023, 2023, 9889312. [Google Scholar] [CrossRef]
  80. Zhang, L.; Ye, J.; Tu, Y.; Wang, Q.; Pan, H.; Wu, L.; Zheng, X.; Zhu, J. Oxygen Modified CoP2 Supported Palladium Nanoparticles as Highly Efficient Catalyst for Hydrolysis of Ammonia Borane. Nano Res. 2022, 15, 3034–3041. [Google Scholar] [CrossRef]
  81. Chen, W.; Li, D.; Wang, Z.; Qian, G.; Sui, Z.; Duan, X.; Zhou, X.; Yeboah, I.; Chen, D. Reaction Mechanism and Kinetics for Hydrolytic Dehydrogenation of Ammonia Borane on a Pt/CNT Catalyst. AIChE J. 2017, 63, 60–65. [Google Scholar] [CrossRef]
  82. Wang, Q.; Fu, F.; Yang, S.; Martinez Moro, M.; Ramirez, M.D.L.A.; Moya, S.; Salmon, L.; Ruiz, J.; Astruc, D. Dramatic Synergy in CoPt Nanocatalysts Stabilized by “Click” Dendrimers for Evolution of Hydrogen from Hydrolysis of Ammonia Borane. ACS Catal. 2019, 9, 1110–1119. [Google Scholar] [CrossRef]
  83. Fu, W.; Wang, Q.; Chen, W.; Qian, G.; Zhang, J.; Chen, D.; Yuan, W.-K.; Zhou, X.; Duan, X.; Yuan, W. Fabrication and Engineering of Ru Local Structures toward Enhanced Kinetics of Hydrogen Generation. Authorea Prepr. 2020, 2. [Google Scholar] [CrossRef]
  84. Mohsenzadeh, A.; Bolton, K.; Richards, T. DFT Study of the Adsorption and Dissociation of Water on Ni(111), Ni(110) and Ni(100) Surfaces. Surf. Sci. 2014, 627, 1–10. [Google Scholar] [CrossRef]
  85. Cui, C.; Liu, Y.; Mehdi, S.; Wen, H.; Zhou, B.; Li, J.; Li, B. Enhancing Effect of Fe-Doping on the Activity of Nano Ni Catalyst towards Hydrogen Evolution from NH3BH3. Appl. Catal. B Environ. 2020, 265, 118612. [Google Scholar] [CrossRef]
  86. Liu, R. Adsorption and Dissociation of H2O on Au(1 1 1) Surface: A DFT Study. Comput. Theor. Chem. 2013, 1019, 141–145. [Google Scholar] [CrossRef]
  87. Petersen, T.; Klüner, T. Water Adsorption on Ideal Anatase-TiO2(101)—An Embedded Cluster Model for Accurate Adsorption Energetics and Excited State Properties. Z. Phys. Chem. 2020, 234, 813–834. [Google Scholar] [CrossRef]
  88. Zhao, W.; Bajdich, M.; Carey, S.; Vojvodic, A.; Nørskov, J.K.; Campbell, C.T. Water Dissociative Adsorption on NiO(111): Energetics and Structure of the Hydroxylated Surface. ACS Catal. 2016, 6, 7377–7384. [Google Scholar] [CrossRef]
  89. Chen, W.; Zheng, W.; Cao, J.; Fu, W.; Qian, G.; Chen, D.; Zhou, X.; Duan, X. Atomic Insights into Robust Pt-PdO Interfacial Site-Boosted Hydrogen Generation. ACS Catal. 2020, 10, 11417–11429. [Google Scholar] [CrossRef]
Figure 1. (a) The correlation between ΔE and Ea (the Brønsted–Evans–Polanyi principle) for dissociative adsorption of NH3BH3 (4) and H2O (5), and (b) the correlation between Eads and Ea (the Sabatier principle) for dissociative adsorption of H2O (5) on different metal surfaces.
Figure 1. (a) The correlation between ΔE and Ea (the Brønsted–Evans–Polanyi principle) for dissociative adsorption of NH3BH3 (4) and H2O (5), and (b) the correlation between Eads and Ea (the Sabatier principle) for dissociative adsorption of H2O (5) on different metal surfaces.
Materials 17 01794 g001
Figure 2. (a) The effect of hydride nature and replacement of H2O with D2O on the rate of H2 evolution in the presence of the Co0 catalyst. (b) The calculated kH/kD ratios at cycling durability tests with the Co0 catalyst in NaBH4 hydrolysis.
Figure 2. (a) The effect of hydride nature and replacement of H2O with D2O on the rate of H2 evolution in the presence of the Co0 catalyst. (b) The calculated kH/kD ratios at cycling durability tests with the Co0 catalyst in NaBH4 hydrolysis.
Materials 17 01794 g002
Figure 3. The activity of Co3O4 in the hydrolysis of studied boron-containing hydrides at 40 and 60 °C. *—sample extraction from the reaction medium for study.
Figure 3. The activity of Co3O4 in the hydrolysis of studied boron-containing hydrides at 40 and 60 °C. *—sample extraction from the reaction medium for study.
Materials 17 01794 g003
Figure 4. The proposed scheme of Co3O4 activation by copper compounds in the reaction medium of hydrides.
Figure 4. The proposed scheme of Co3O4 activation by copper compounds in the reaction medium of hydrides.
Materials 17 01794 g004
Figure 5. The comparison of activity of Co0, Co3O4, and CuO (10 wt%)-Co3O4 in the hydrolysis of (a) NaBH4, (b) NH3BH3, and (c) (CH2NH2BH3)2 at 40 °C.
Figure 5. The comparison of activity of Co0, Co3O4, and CuO (10 wt%)-Co3O4 in the hydrolysis of (a) NaBH4, (b) NH3BH3, and (c) (CH2NH2BH3)2 at 40 °C.
Materials 17 01794 g005
Figure 6. The TEM images of different magnitudes for initial Co3O4 (ac) and Co3O4 extracted from the reaction medium of NaBH4 (df).
Figure 6. The TEM images of different magnitudes for initial Co3O4 (ac) and Co3O4 extracted from the reaction medium of NaBH4 (df).
Materials 17 01794 g006
Figure 7. (a) TPR H2 spectra for initial Co3O4 and after its treatment in NaBH4 solution (Co3O4/SBH), deconvolution of spectra of initial Co3O4, (b) and Co3O4/SBH (c).
Figure 7. (a) TPR H2 spectra for initial Co3O4 and after its treatment in NaBH4 solution (Co3O4/SBH), deconvolution of spectra of initial Co3O4, (b) and Co3O4/SBH (c).
Materials 17 01794 g007
Figure 8. Raman spectra for Co3O4 and Co3O4/SBH (a), A1g peak region (b), and F2g peak region (c).
Figure 8. Raman spectra for Co3O4 and Co3O4/SBH (a), A1g peak region (b), and F2g peak region (c).
Materials 17 01794 g008
Figure 9. (a) ATR FTIR spectra of initial Co3O4 and Co3O4/SBH, as well as initial Co3O4 with preadsorbed water on its surface, and (b) Co-O vibration region.
Figure 9. (a) ATR FTIR spectra of initial Co3O4 and Co3O4/SBH, as well as initial Co3O4 with preadsorbed water on its surface, and (b) Co-O vibration region.
Materials 17 01794 g009
Figure 10. (a) The optimized structure of initial 311 Co3O4 clusters and Co3O4 clusters with oxygen vacancies (Vo1 (b) and Vo2 (c)). The atoms of Co are pink, and O are red.
Figure 10. (a) The optimized structure of initial 311 Co3O4 clusters and Co3O4 clusters with oxygen vacancies (Vo1 (b) and Vo2 (c)). The atoms of Co are pink, and O are red.
Materials 17 01794 g010
Figure 11. Changes in the electron density of cobalt atoms relative to Co3O4. The numbers of the Co atoms correspond to the atoms in Figure 10.
Figure 11. Changes in the electron density of cobalt atoms relative to Co3O4. The numbers of the Co atoms correspond to the atoms in Figure 10.
Materials 17 01794 g011
Figure 12. Energy profile of dissociative adsorption of water on the surfaces of Co3O4 and Co3O4 with oxygen vacancies with structures of initial (IS), transition (TS), and final (FS) states. The atoms of Co are pink, O are red, and H are grey.
Figure 12. Energy profile of dissociative adsorption of water on the surfaces of Co3O4 and Co3O4 with oxygen vacancies with structures of initial (IS), transition (TS), and final (FS) states. The atoms of Co are pink, O are red, and H are grey.
Materials 17 01794 g012
Figure 13. Comparison of linear correlations between Eads and Ea (the Sabatier principle) for the dissociative adsorption of H2O on modeled surfaces of cobalt oxide and metals.
Figure 13. Comparison of linear correlations between Eads and Ea (the Sabatier principle) for the dissociative adsorption of H2O on modeled surfaces of cobalt oxide and metals.
Materials 17 01794 g013
Figure 14. Change in kinetics of H2 generation over Co3O4 in NaBH4 hydrolysis at replacement H2O by D2O (40 °C).
Figure 14. Change in kinetics of H2 generation over Co3O4 in NaBH4 hydrolysis at replacement H2O by D2O (40 °C).
Materials 17 01794 g014
Table 1. Adsorption energy values (Eads) for H2O and NH3BH3 and corresponding changes in the geometry of adsorbed molecules calculated in this work.
Table 1. Adsorption energy values (Eads) for H2O and NH3BH3 and corresponding changes in the geometry of adsorbed molecules calculated in this work.
MetalEads H2O, eVM-O, Å∠H-O-H, °Eads AB, eVM-H, Å∠H-B-H, °
Co−0.3642.200105.7−2.4511.162142.7
Ni−0.3172.167105.6−1.8641.594137.2
Cu−0.1842.349105.0−1.2331.814115.6
H2O/NH3BH3 104.4 113.4
Table 2. XRD data for initial sample Co3O4 and Co3O4 sample activated in the reaction medium of NaBH4 hydrolysis.
Table 2. XRD data for initial sample Co3O4 and Co3O4 sample activated in the reaction medium of NaBH4 hydrolysis.
SamplePhase Composition, wt%SCR 1, nma(Co3O4) 2, Å
Co3O4100% Co3O4418.084 (±0.001)
Co3O4/SBH95% Co3O4408.085 (±0.001)
5% CoO 317
1 The average size of crystallites (CSR) was calculated using the Sherrer formula for 311 reflections for Co3O4 and 111 for CoO. 2 a(Co3O4)—lattice parameter of Co3O4 (sp. gr. Fd3m). 3 The oxidation of CoO takes place when the protective alcohol layer evaporates and the sample comes into contact with air.
Table 3. Kinetic isotope effect (kH/kD) for the cobalt-based catalysts in the catalytic hydrolysis of boron-containing hydrides at 40 °C (first test).
Table 3. Kinetic isotope effect (kH/kD) for the cobalt-based catalysts in the catalytic hydrolysis of boron-containing hydrides at 40 °C (first test).
SampleNaBH4NH3BH3(CH2NH2BH3)2
Co02.42.52.3
Co3O4 11.2--
10% CuO-90% Co3O41.71.61.7
1 Kinetic data after the activation period were analyzed.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Butenko, V.R.; Komova, O.V.; Simagina, V.I.; Lipatnikova, I.L.; Ozerova, A.M.; Danilova, N.A.; Rogov, V.A.; Odegova, G.V.; Bulavchenko, O.A.; Chesalov, Y.A.; et al. Co and Co3O4 in the Hydrolysis of Boron-Containing Hydrides: H2O Activation on the Metal and Oxide Active Centers. Materials 2024, 17, 1794. https://doi.org/10.3390/ma17081794

AMA Style

Butenko VR, Komova OV, Simagina VI, Lipatnikova IL, Ozerova AM, Danilova NA, Rogov VA, Odegova GV, Bulavchenko OA, Chesalov YA, et al. Co and Co3O4 in the Hydrolysis of Boron-Containing Hydrides: H2O Activation on the Metal and Oxide Active Centers. Materials. 2024; 17(8):1794. https://doi.org/10.3390/ma17081794

Chicago/Turabian Style

Butenko, Vladislav R., Oksana V. Komova, Valentina I. Simagina, Inna L. Lipatnikova, Anna M. Ozerova, Natalya A. Danilova, Vladimir A. Rogov, Galina V. Odegova, Olga A. Bulavchenko, Yuriy A. Chesalov, and et al. 2024. "Co and Co3O4 in the Hydrolysis of Boron-Containing Hydrides: H2O Activation on the Metal and Oxide Active Centers" Materials 17, no. 8: 1794. https://doi.org/10.3390/ma17081794

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