Next Article in Journal
Zero Waste as a Determinant of Shaping Green Economy Processes on the Example of Communes of Eastern Poland in 2010–2020
Previous Article in Journal
Comprehensive Study on Reduced DC Source Count: Multilevel Inverters and Its Design Topologies
Previous Article in Special Issue
Phenyl Vinylsulfonate, a Novel Electrolyte Additive to Improve Electrochemical Performance of Lithium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Electrochemical Reduction of Sodium Metaborate

1
Department of Chemistry, Faculty of Science, Universiti Malaya, Kuala Lumpur 50603, Malaysia
2
Nanotechnology and Catalysis Research Center (NanoCat), University Malaya, Kuala Lumpur 50603, Malaysia
3
Department of Chemical Engineering, Faculty of Engineering, University Malaya, Kuala Lumpur 50603, Malaysia
*
Author to whom correspondence should be addressed.
Energies 2023, 16(1), 15; https://doi.org/10.3390/en16010015
Submission received: 22 November 2022 / Revised: 6 December 2022 / Accepted: 14 December 2022 / Published: 20 December 2022
(This article belongs to the Special Issue Energy Electrochemistry)

Abstract

:
The recycling of sodium borohydride poses a huge challenge to the drive towards a hydrogen economy. Currently, mechano-chemical, thermo-chemical and electrochemical are the only reported methods of recycling sodium metaborate into sodium borohydride. Much attention has been devoted to the mechano-chemical and thermo-chemical methods of reduction, but little focus has been devoted to electrochemical methods. This review describes the electrochemical behaviour of borohydride (BH4) and metaborate (BO2) anions in alkaline solutions. The BH4 is stabilized in highly concentrated alkaline solutions, while the electro-oxidation of BH4 is dependent on the type of electrode material. The attempts to electro-reduce the BO2 into BH4 is reviewed and the challenges, suggestions and future outlook of electro-reduction for the recycling of BO2 into BH4 is highlighted.

1. Introduction

Sodium borohydride (NaBH4) is a hydrogen carrier with a high theoretical hydrogen storage capacity of around 10.8% w/w. In addition, the safe storage and fast kinetics of the release of gaseous hydrogen when in contact with water offers the storage and transport of hydrogen in solid form, which precludes the use of heavy high-pressurized gas storage containers and gas transport pipelines. The release of gaseous hydrogen at room temperature and pressure are described in the following equation:
NaBH4 + 4H2O → NaBO2 · 2H2O + 4H2
The release of hydrogen from concentrated alkaline solutions of NaBH4 necessitates the utilisation of NaBH4 as the active electrochemical reductant in direct borohydride fuel cell (DBFC):
BH4 + 8OH → BO2 + 6H2O + 8e  Eϴ/V = 1.24 vs. SHE
In both cases, the release of molecular hydrogen (Equation (1)) and the utilisation of the BH4 in DBFC (Equation (2)) release the sodium metaborate NaBO2 as the by-product. Therefore, successfully recycling the NaBO2 by-product into NaBH4 is the key for the utilisation of NaBH4 as a solid hydrogen storage material and as the electroactive reductant in DBFC.
The industrial production of NaBH4 developed in the 1950s, known as the Brown—Schlesinger process [1], which is highly inefficient, requires high temperatures, typically around 250 °C, and expensive chemical precursors, as given in the following equation:
B(OCH3)3 + NaH → NaBH4 + NaOCH3
Until now, there has not been much progress in increasing the efficiency of the production of NaBH4.
Another industrial process, known as the Bayer process, utilises the mineral borax (Na2B4O7·10H2O) in powder form but reacted with sodium metal and hydrogen gas with the presence of silica, as given in the equation below:
Na2B4O7 + 16Na + 8H2 + 7SiO2 → 4NaBH4 + 7Na2SiO3
The Bayer process also possesses low conversion efficiency in the production of NaBH4, in addition to being an energy-intensive process with operational temperatures of more than 400 °C [2].
All these factors render NaBH4 an expensive chemical to be utilised even in DBFC and as a hydrogen carrier material. One solution is to recycle the NaBO2 by-product into NaBH4. Among the methods for the recycling of NaBO2 by-product into NaBH4 are thermochemical [3,4,5,6,7,8], which involves the reaction of NaBO2 in the presence of hydrogen gas at high pressures, and metallic Mg as the reducing agent at temperatures more than 500 °C.
Another method for the recycling of the NaBO2 by-product into NaBH4 is by mechano-chemical means, or by ball-milling powder of the NaBO2 in a reactor in the presence of reductants, in the presence of steel balls, to facilitate the mechanical knocking process [9,10,11,12,13,14,15,16,17,18,19].
The electrochemical process, on the other hand, involves the electrochemical reduction of NaBO2 into NaBH4 with the passage of a reduction current. As mentioned in a recent review by Helder X. Nunes et al., 2021 [20], since the year 2000, the number of papers utilising electrochemical reduction for the recycling of NaBO2 into NaBH4 only constitutes around 18% of the total number of papers investigating the regeneration of NaBH4, whereas the bulk are dedicated to mechano-chemical (50%) and thermo-chemical (32%) methods of recycling. This mini-review focuses on the electrochemical behaviour of NaBO2 and NaBH4 in aqueous solution, the electrolytic reduction of NaBO2, and finally, the challenges and future outlook of the process.
Much information regarding the thermodynamics of the competing reactions against the electroreduction of BO2 (Equation (2)) is available in the literature. The electroreduction of BO2 should be performed in concentrated alkaline solutions, as the BH4 is stabilized with the presence of the OH in high concentrations. In concentrated alkaline solutions, there are few reactions which are competing reactions and are more thermodynamically favourable compared to Equation (2), which makes the electroreduction of BO2 difficult in alkaline solutions. Among them are the hydrogen evolution reaction (HER) and reductions involving dissolved oxygen.
2H2O + 2e → 2OH + H2  Eϴ/V = −0.83 vs. SHE
The presence of dissolved oxygen presents an avenue for the reduction of molecular oxygen into various ionic species in alkaline solution.
O2 + 2H2O + 4e → 4OH  Eϴ/V = −0.40 vs. SHE
O2 + H2O + 2e → OH + OH2  Eϴ/V = −0.076 vs. SHE
O2 + 2H2O + 2e → H2O2 + 2OH  Eϴ/V = −0.146 vs. SHE
The reason for the thermodynamic feasibility of these reactions compared to Equation (2), is that these involve uncharged water and molecular oxygen in alkaline solutions, compared to the eight-electron reduction of a negatively charged BO2 anion.

2. Analytical Methods of Detection of NaBH4

Several analytical methods have been used for the quantitative determination of BH4; among them are iodometry, hydrogen evolution, hydride hydrolysis, Fourier transformed infrared spectroscopy (FTIR), and electrochemical methods. The reader is referred to the report by Biljana et al., 2013 [21], which discusses both the disadvantages and advantages of each technique.

2.1. Non-Iodometric Detection

Santos et al. [22] developed a hydride hydrolysis method based on the evolution of hydrogen gas upon decomposition of NaBH4, where the volume of H2 is proportional to the amount of NaBH4. It is important that the calibration process and the determination of the NaBH4 in the real sample are performed at same pressure and temperature. According to the authors, this method is simple in fabrication and operation, has high reproducibility but does not possess a low limit of detection (500 ppm) [22]. One suggestion for the improvement of this technique is to couple the setup with a gas chromatography mass spectrometer (GCMS) for the accurate determination of H2 gas, as seen in many photo-catalytic and photo-electrocatalytic water splitting experiments.
Voltammetric methods of detection of BH4 (Section 3), such as cyclic voltammetry (CV), linear scan voltammetry (LSV) and square wave voltammetry (SWV), are based on the electro-oxidation into BO2 and offer high sensitivity with a low limit of detection of 1 ppm [23].
BH4 + xOH → B(OH)4 + (x − 4)H2O + (4 − x/2)H2O + xe
If the n number of transported electrons is eight in highly concentrated alkaline solutions, then it conforms to the complete electro-oxidation into the BO2 without the hydrolysis reaction to release H2 gas [21]. The n number strongly depends on the electrode material and the concentration of the alkaline solution. The competition between hydrolysis (Equation (1)) and electro-oxidation (Equation (2)) of the BH4 determines the success of this method, in which the BH4 is stabilized in highly concentrated alkaline solutions. As will be explained in Section 3, the n value is usually lower than eight, which confirms that the hydrolysis of BH4 to release H2 cannot be completely avoided even in concentrated alkaline solutions.
Amendola et al. [24] reported a relatively simple potentiometric method for the analytical detection of the BH4 by measuring the open circuit potential (OCP) at a metal indicator electrode (IE) versus a reference electrode (RE), preferably mercury/mercury oxide (Hg/HgO) in alkaline solutions. This method is based on the Nernstian relationship, where the OCP is proportional to the log10[BH4] and only requires the potential measurement using a digital multi-meter without the use of a potentiostat/galvanostat, as seen in scanning techniques such as cyclic voltammetry (CV), linear scan voltammetry (LSV) or square wave voltammetry (SWV). By selecting a suitable IE, the concentration of BH4 was varied to obtain the linear regions. The linear regions depend on the type of IE material and the concentration of the alkaline solutions. For example, the linear range for the detection of [BH4] is 10−7–10−3 M for gold IE in 4 M KOH, while Pd IE shows a linear range of 10−5–10−3 M in 4 M NaOH [22].
Spectroscopic detection of BH4 was reported by Elamin et al. [25], which is based on the reaction between the tri-nitrobenzenesulfonate anion with BH4 in alkaline aqueous solutions, which gives a UV absorption maximum at 460 nm [25]. The phosphotungstate anion, [PW12O40]3−, can be reduced into a heteropoly blue species with the characteristic violet-blue colour upon reduction with BH4 in alkaline solutions with a UV absorption maximum at 680 nm and a lowest limit of detection of 10−4 M [26]. However, the reduction of the [PW12O40]3− is reversible; the heteropoly blue species can be oxidized into the colourless [PW12O40]3− in the presence of dissolved oxygen in the alkaline medium.

2.2. Iodometry

In iodometry, the oxidation of the BH4 with the addition of excess of the IO3 highly alkaline solution is shown below:
5BH4 + 8IO3 → 5BO2 + 4I2 + 8OH + 6H2O
The released I2 is titrated with the S2O32− to obtain the amount of I2:
I2 + 2S2O32− → 2I + S4O62−
In highly alkaline solution, the presence of OH promotes the oxidation of the I2 with a different reaction pathway [26]:
4I2 + S2O32− + 10OH → 2SO42− + 8I + 5H2O
The errors associated with the iodometric method of analysis for the quantitative determination of the BH4 are due to two effects:
(a)
the presence of high concentrations of OH enhances the reduction ability of the S2O32−. This effect shows that eight-times less S2O32− is required to reduce the I2 into I compared to Equation (2). This accounts for the erroneously higher values of the BH4 obtained from this method.
(b)
in addition, the BO2 product can also be oxidized into the tetraborate anion B4O72− [27]:
4BO2 + 2OH → B4O72− + H2O + O2 + 4e
and with the presence of excess IO3, as follows:
4IO3 + 8BO2 + 4H2O → 2I2 + 2B4O72− + 8OH + 5O2
This reaction (Equation (9)) also releases I2, which accounts the erroneously higher amount of the BH4 in alkaline solutions.
The iodometric method of analysis has been reported by few researchers in the ball-milling of NaBO2 for the conversion into NaBH4, yielding high conversion efficiencies of nearly 76% [9], 80% [4,15,19] and even 90% [13]. Although the inaccuracies associated with the iodometric method of determination are described above, errors in such conversion efficiencies are easily reported when not confirmed by other methods of analysis. The iodometric method of quantitative detection should be abandoned as it provides an over-exaggeration of the amount of BH4 in the sample. Thus, the electroanalytical behaviour of the BO2 and BH4 should be investigated (Section 3), as it is becoming an important technique for the possibility of sensor fabrication for the quantitative detection of BH4, especially in alkaline solutions.

3. Electroanalytical Behaviour of NaBH4

Mirkin et al.’s (1991) [28] was perhaps the earliest attempt to understand the electroanalytical behaviour of NaBH4. They reported cyclic voltammetry of a series of NaBH4 in 0.2 M NaOH on gold disc working electrode (WE) with RuO2 counter electrode (CE) and mercury/mercury sulfate (Hg/HgSO4) reference electrode (RE). With five repetitive scans at 0.1 V/s, they reported a single peak at around −0.6 V (vs. Hg/HgSO4) using 3.2 mM NaBH4, which confirms the stable electro-oxidation behaviour of the gold electrode free from any irreversible changes. Linear plots of the current (I/A) vs. the square root of scan rate (ν1/2), and anodic peak current vs. borohydride concentration (mM) were obtained, confirming that the electro-oxidation of BH4 is mass transport controlled. In contrast, several overlapping peaks were found using Pt WE. They concluded that using voltammetric method on gold WE was more advantageous due to the stability and selectivity compared to iodometric method, which often yields an erroneously high concentration of NaBH4. There was no reverse peak obtained for the electro-reduction of BO2 to BH4. The same authors investigated the electro-oxidation of BH4 on gold microelectrode and concluded that the electro-oxidation process involves an 8-electron electrochemical–chemical–electrochemical (ECE) mechanism [29].
They also reported that below pH 12, BH4 undergoes hydrolysis to release hydrogen gas, as follows [29]:
BH4 + H2O → BH3OH + H2
thus, the loss of BH4 should be avoided by storage in highly alkaline solution.
They also reported an n value of 7.9 ± 0.6 electrons using the Levich equation in close agreement with the theoretical value of 8.
Sanli et al. [30] studied the electro-oxidation of NaBH4 in a three-compartment electrochemical cell separated with a Nafion cationic exchange membrane. The WE, CE and RE was Au disc, Pt wire and Ag+/Ag, respectively. They reported a peak for BH4 oxidation at 0.3 V on a clean Ag surface but shifted to 0.1 V with an oxidized Ag electrode. It was obvious that the formation of silver oxides in alkaline solution had a catalytic effect on the electro-oxidation of NaBH4. They also reported linear plots of peak current vs. borohydride concentration (M), confirming that the electro-oxidation is mass transport controlled. Coulometry at 0.45 V in 3M NaOH using 10 mM NaBH4 gave an n number of electrons around 6 for the electro-oxidation of NaBH4. The same authors later investigated the electro-oxidation of BH4 on Ag2O WE and obtained a n number of electrons almost of 6 [31].
Celikkan et al. [23] studied the electrochemical behaviour of NaBH4 in NaOH solution using several WE, such as Au, Pt, Pd and Ni. They reported a peak at −0.4 V (vs. saturated calomel electrode, SCE) at lower scan rate, which shifted to −0.3 V (vs. SCE) at higher scan rates using gold electrode. Linear plots of current (I/A) vs. the square root of scan rate (ν1/2), and anodic peak current vs. borohydride concentration (mM) were also obtained, confirming the electro-oxidation of the BH4 is quasi-reversible with a slight shift in the peak potential of the oxidation currents. No reverse peak was obtained for the electro-reduction of BO2 to BH4. The diffusion coefficient (D) was determined from the Cottrell equation, which gave a value of 4.02 × 10−6 cm2 s−1, while rotating disc electrode (RDE) experiments gave an n number of electrons around 6 and 2, corresponding to the electro-oxidation of the BH4 as:
BH4 + H2O → 3H2 + BH(OH)3  6 electron transfer
BH(OH)3 + H2O → BO2 + H2  2 electron transfer
Santos et al. [32] investigated the electro-oxidation of BH4 on gold electrode in 2 M NaOH, using gold CE and SCE RE. Figure 1 shows the cyclic voltammetry of NaBH4 in 2 M NaOH solution. They reported the oxidation peak at −0.2 V (vs. SCE) while a gold oxidation peak at higher potential of 0.3 V (vs. SCE). They obtained n values between 5 and 8; n values of less than 8 could be due the formation of boron oxides film on the gold electrode [33]. The a1 peak in Figure 1 is attributed to the 8-electron electro-oxidation of BH4, while peaks a2 and c2 belong to the adsorbed BH3OH, which is an intermediate in the electro-oxidation of BH4 on gold electrode [34].
The competition between the electro-oxidation of BH4 and hydrolysis is present especially at pH lower than 12 as described in Equation (2). Thus a lower number of electrons (6) released for the electro-oxidation of BH4 compared to Equation (2) is due to the presence of the BH3OH:
BH3OH + 6OH → B(OH)4 + 3H2O + 6e
The peak a1 in Figure 1 could be due to the presence of BH3OH [34]:
BH3OH + 3OH → B(OH)4 + 3/2H2 + 3e
The detail oxidation behaviour of the BH4 on platinum anode is to date still largely unknown due to the platinum activity towards hydrogen adsorption and hydrogen evolution reaction (HER), which are the competing reactions towards the electro-oxidation of the BH4 [34,35]. Thus, the electro-oxidation of the BH4 on platinum does not involve eight electrons and the Faradaic efficiency is lower compared to gold electrode [34,36,37].
Thus, gold anode is generally the preferred electrode for the investigation of the electro-oxidation of BH4, as it does not oxidize at positive potentials in alkaline solutions, and is inactive towards the heterogenous hydrolysis of the BH4 and the HER [29,38,39]. As previously confirmed, the electro-oxidation of the BH4 on gold electrode involves eight electrons (Equation (2)).
E. Gyenge [34] studied the voltammetry of BH4 and reported that the concentration of the hydroxide solution affects the oxidation potential of the BH4. The standard potential for the reduction of the BH4 where the oxidation potentials of the BH4 on gold electrode has been reported to occur at positive potentials (vs. SHE) has been compared to the standard potential of the BH4 (Equation (2)). Thus, the kinetics of the oxidation of the BH4 are controlled by the concentration of the hydroxide solution, and the pH of 12 or less promoted the homogenous hydrolysis of the BH4; however, it is stable in concentrations of 2 M OH and above.
S. Colominas et al. [40] investigated the electrochemical behaviour of NaBH4 on gold electrode in 2 M NaOH solution. Figure 2 shows an oxidation peak around −0.4 V (vs. mercury/mercury oxide, Hg/HgO), which was obtained as reported by other researchers [34] with no reduction peak at the reverse scan.
Chatenet et al. [41] also reported n = 7 as the number of electrons involved in the electro-oxidation of the BH4 on gold electrode using the Levich equation in 1 M NaOH. As confirmed from previous studies [42], the n number of electrons in the electro-oxidation of BH4 on gold electrode is controlled by the ratio of OH to BH4 ([OH]/[BH4]) concentration. Smaller ratios inhibit the oxidation process and give smaller n values than eight, but higher ratios give n values slightly higher than eight. Different ratios of the [OH]/[BH4] also gave different onset oxidation potentials at the voltammograms. For example, a unit ratio of the [OH]/[BH4] gave an onset oxidation potential of around −1.2 V (vs. NHE) compared to −0.5 V (vs. NHE) at higher ratios. Thus, this ratio controls the overall oxidation kinetics of the BH4 at gold electrode. They suggest a large excess of OH, typically a ratio of [OH]/[BH4] ≥ 100, for the sufficient oxidation of the BH4, for faster kinetics of oxidation of n values higher than 7, and for total suppression of heterogenous hydrolysis of the BH4 anion.
On the other hand, we have the diffusion coefficient (D) of the BH4 oxidation current magnitude on gold electrodes. Other researchers [43] reported that high concentrations of alkaline solution decrease the diffusion coefficient of the BH4, as also predicted by the Stokes–Einstein equation.
M. Chatenet et al. [44] reported that the kinematic viscosity of the NaBH4 in NaOH solution increases with higher concentrations of NaBH4 and NaOH, which influences the diffusion coefficient value of the BH4 from RDE experiments in different concentrations of NaBH4 in alkaline solutions (Figure 3). The diffusion coefficient obtained using the rotating-ring disc electrode (RRDE) technique was nearly twice larger in a solution of 0.01 M NaBH4 in 1 M NaOH at 25 °C, compared to the diffusion coefficient in 0.02 M NaBH4 and 1 M NaOH at 30 °C obtained using chronoamperometry technique on gold microsphere electrode. They suggest the formation of boron oxide films on the gold surface and heterogenous hydrolysis releasing hydrogen gas in an insufficiently low concentration of NaOH solution could affect the value of the diffusion coefficients [44]. The presence of boron oxides and hydrogen gas bubbles yield diffusion coefficients smaller than expected compared to those obtained in their absence, as both phenomena hinder the mass transport of the BH4 to the gold electrode [44].
They reported that the D values were dependent on the concentration of both NaOH and BH4, which also influenced the n values of the number of electrons transferred [44]. At the same concentration of the BH4 (0.001 M), higher concentration of NaOH (1 M) gave higher values of n (5.8–6.6), but lower values of D (2.26–3.71 × 10−5 cm2 s−1), compared to lower concentration of NaOH (0.1 M) with n (5.1–4.8) and D (2.78–4.49 × 10−5 cm2 s−1), at all temperatures (10, 25 and 40 °C), though 1 M NaOH possesses slightly higher viscosity than 0.1 M NaOH [44]. With the same concentration of BH4 but an increase in NaOH concentration to 4 M, the D values decreased further (1.41–2.31 × 10−5 cm2 s−1) and a slight decrease in the n values (5.3-6.0) was observed at all temperatures. The decrease in the D values in 4 M NaOH was due to the increase in the kinematic viscosity. With an increase in the BH4 concentration (0.01 M) in 4 M NaOH, the D values were the smallest at all temperatures (0.74–2.21 × 10−5 cm2 s−1) due to the increase in the viscosity of the solution, while the n values did not show an obvious trend.
Previous measurements using polarography on dropping mercury electrode (DME) gave higher D values of 2.1 × 10−5 cm2 s−1 in 0.11 M NaOH [45] and 10−4 cm2 s−1 in 0.1 M NaOH [46]. The higher D values could be due to the absence of boron oxide film formation and the formation of hydrogen bubbles on the mercury electrode surface.
An n value greater than 8 was reported by H. Cheng et al. [47] and Atwan et al. [48], 9.7 and 12, respectively, using the RDE method in 2 M NaOH at 25 °C. Therefore, these n values greater than 8 are not represented from a thermodynamic viewpoint, but the increase in the solution viscosity must be considered for more accurate calculation of the n values [44].
Thus, the D and n values are dependent on many factors, among which are the type of experimental technique involved and concentration of BH4 and NaOH, together with the temperature. The importance of understanding the electrochemical behaviour of the BH4 with regard to the D and n values is an important step in the quantitative electrochemical detection of the BH4, as seen in the Section 4.

4. Electrolysis of BO2 into BH4

A few patents describe the electro-reduction of BO2 into BH4 in an aqueous medium [49,50,51], but later reports [26] did not verify any viable conversion of BO2 into BH4 in an aqueous medium, although the current efficiencies of the conversion, 20–25%, were reported in those patents [49,50,51]. This could be due to the iodometric titration, which yields higher NaBH4 formation, which was adopted by the patents [26].
McLafferty et al. [52] also performed experiments to verify several patents [52,53,54] that claim the successful electrochemical reduction of BO2 into BH4. They performed electrolysis of boric acid on mercury pool cathode in a divided electrolysis cell with cationic exchange membrane with sulfuric acid as the anolyte and BO2 in 2 M NaOH as the catholyte. Since mercury has high overpotential for the hydrogen evolution reaction (HER), it was selected as the cathode in this study, and the potential of the mercury cathode was maintained between −2.16 and −2.60 V (vs. Hg/HgO). The authors found no presence of the BH4 using cyclic voltammetry of alkaline solution even after two days of electrolysis. The same results were obtained with the hydrolysis of BO2 in the catholyte at gold and Cu electrodes, to reproduce the work of Jiang et al. [55], but no traces of the BH4 were detected using CV of the catholyte after electrolysis. Gold cathode was used again in pulsed potential electrolysis and there was no presence of any BH4 detected using CV. Titanium, Ag and Pb, which possess high overpotentials of HER, also did not show any conversion into BH4 in the alkaline reduction of the BO2 in the divided electrolysis cell. The authors used tetraalkylammonium hydroxide (TAAH) to increase the overpotential of the HER. All experiments concluded that the TAA+ cation was ineffective in absorbing at the cathode surface due to the replacement of water molecules at higher negative potentials, which promoted the HER. The authors also tried to modify the cathode surface using a chelating compound or a cationic polyelectrolyte to coat the gold electrode surface, forming a self-assembled monolayer (SAM), which promotes the attraction of the BO2 at small distances from the gold electrode surface. The chelating agent utilised for this purpose was 1-thioglycerol, which consists of a thiol group, which strongly bounds to the gold surface, and a diol group, which could bind with the BO2, while a quaternized poly(4-vinylpyridine) was selected as the cationic polyelectrolyte. The electrolysis was repeated as with the high overpotential metal cathode, but no presence of the BH4 ion was detected in the CV.
Sanli et al. [56] carried out electrolyses in 1 M NaOH + 0.1 M NaBO2 at 0.5 V for 24 h and 48 h with Ag gauze cathode at room temperature and atmospheric pressure in an undivided electrochemical cell. The authors performed iodometric titration for the determination of the conversion efficiency of sodium metaborate into sodium borohydride. It was observed that there was 10% conversion after 24 h of electrolysis, reaching 17% (by iodometric titration) after 48 h. It was obvious that the low conversion efficiency was due to the HER at the cathode, which is a strong competing reaction with the electro-reduction of metaborate into borohydride.
The hydrogen is produced in situ at the cathode:
2H2O + 2e ⇌ 2OH + H2  Eϴ/V = −0.83 vs. SHE
which was used to reduce the borate into borohydride:
BO2 + 4H2 ⇌ BH4 + 2H2O
The authors performed cyclic voltammetry of NaBO2 in 1 M NaOH and found that the electro-reduction of BO2 occurred at 0.5 V (vs. SCE) and the corresponding electro-oxidation of BH4 into the BO2 at 0.1 V (vs. SCE). They also reported that gold electrode is the most effective for the reduction of BO2 compared to Pt, Pd and Ni electrodes.
Santos et al. [22] performed controlled potential (1.8 V) bulk electrolysis for the electroreduction of 2 M NaBO2 at cadmium cathode in 4 M NaOH in a divided electrochemical cell with a cationic exchange membrane (CEM), as shown in Figure 4. The cathodic compartment consists of an indicator electrode connected to a high impedance voltmeter for the potentiometric monitoring of the concentration of the sodium borohydride. However, the authors did not report any conversion efficiency for this electrochemical method. Upon completion of electro-reduction, the NaBH4 can be isolated from aqueous solution by freeze-drying to eliminate water, leaving the NaBO2, NaOH and NaBH4. Then, the solid free water is dissolved in tetraglyme. The NaBH4 is only solubilised in tetraglyme, leaving the NaBO2 and NaOH undissolved. Tetraglyme is removed by thermal evaporation, leaving the NaBH4, tetraglyme (C10H22O5) is also known as tetraethylene glycol dimethyl ether, butyl diglyme (C12H26O3) or diethylene glycol dibutyl ether.
Shen Y. et al. [57] carried the desulfurization of dibenzothiophene with a potential difference 3.5 V and an electrolytic time 4 h in 0.025 M NaOH using Pb cathode in an undivided cell for the regeneration of the NaBH4. Pb cathode was selected due to the large overpotential for the HER, which could not have interfered with the electroreduction of BO2 to BH4. The authors reported that the desulfurization efficiency was increased with the amount of added NaBO2, due to the increased regeneration of the BH4 with the presence of BO2; however, no efficiency conversion of the BO2 into BH4 was reported in this study.
Zhu et al. [58] performed cyclic voltammetry at Cu cathode and Ni counter electrode (CE) and reported that the cathodic peak at 1.2 V (vs. SCE) was due to the formation of the BH4. In addition, they reported a conversion efficiency of 15% from electrolysis using an undivided electrolysis cell, with a potential difference of 1.7 V, in 1 M NaOH as the catholyte, at Cu cathode.
Chenhua et al. [59] used NaBH4 for the de-sulfurization of fuel and regenerated the NaBH4 through electroreduction of the NaBO2. The NaBO2 was converted into sodium borohydride (NaBH4) through electroreduction on boron doped diamond (BDD) cathode. The authors used a divided cell, using a cationic exchange membrane (CEM) to separate the CE from the WE and RE and used a pulsed current for the electro-reduction of the BO2 into BH4. The negatively charged BO2 experiences repulsion at the cathode. Thus, a pulse cycle of positive followed by negative potential facilitated the electro-reduction of the BO2. When the WE was charged positive, the BO2 was attracted to the WE surface, and reduced into the BH4, as the potential was pulsed into the negative potential. The BDD showed the onset potential for the HER is at −1.8 V (vs. SCE), while the onset potential for the oxygen evolution reaction (OER) is around +0.6 V (vs. SCE), as shown in Figure 5. The negative pulse is more positive than −1.8 V to avoid the HER but more negative than 0.6 V to avoid the OER. The electro-reduction of BO2 into BH4 showed a peak at −1.4 V (vs. SCE). The authors did not report any conversion efficiency for the electro-reduction of the BO2 into borohydride BH4. The BH4 was used as an intermediate for the desulfurization by supplying hydrogen for the reductive de-sulfurization of the gasoline samples. Upon reductive desulfurization, the BH4 is oxidized into BO2.
The same authors also used the same method of pulsed potential to re-generate BH4 from BO2 but applied a larger negative pulse potential (−2–2 V to −3.0 V) compared to the positive pulse potential (0.1 V to 0.9 V) [60]. From the CV in ionic liquids (IL), the reduction of the BO2 occurred at −2.5 V vs. Ag/Ag+ using glassy carbon electrode (GCE) as the WE, where the HER was ineffective in providing competition against the electroreduction of the BO2 as it occurred beyond −3 V vs. Ag/Ag+, is shown in Figure 6. The whole regeneration of BH4 from BO2 and desulfurization of a model fuel oil was carried out in an IL, 1-ethyl-3-methylimidazolium bis (trifluoromethanesulfonyl) imide (EMI[TFSI]) in an undivided electrochemical cell. A higher reverse pulse towards the positive potential was avoided due to the effect of oxygen evolution reaction (OER) and the oxidation into borax (sodium tetraborate). The time duration for the negative and positive pulse increased (0.1 s to 1.0 s) the desulfurization efficiency of the model oil; however, the conversion efficiency of the BO2 into BH4 was not calculated.
Hydrogen carrier electrode materials are also hydrogen release materials, as seen in rechargeable nickel metal hydride battery electrodes, such as NiMH and Ni(OH)2 in alkaline solutions.
Anode (negative): NiMH + OH ⇌ NiM + H2O + e
Cathode (positive): NiOOH + H2O + e ⇌ Ni(OH)2 + OH
Both forward reactions are discharge reactions at the anode (-ve electrode) and cathode (+ve electrode) which takes place in rechargeable NiMH batteries in highly alkaline solutions. Both electrodes involve the hydride in the electrochemical redox at both electrodes, which is similar to the sodium borohydride; thus, it could be utilised in the electroreduction of the BO2 into the BH4 (Section 5).

5. Conclusions and Future Outlook

Electrochemical reduction of sodium metaborate is still an expanding field of research and not many papers report the successful conversion of sodium metaborate into sodium borohydride. Research reports which mention the in situ electrochemical reduction of sodium metaborate into sodium borohydride are available in the literature, though they fail to investigate the conversion efficiencies of such techniques. In all the papers reviewed, electrochemical reduction takes place in an alkaline aqueous solution where water is the reductant but the same water molecules promote the hydrolysis of sodium borohydride releasing hydrogen gas when the pH reaches 12. Several patents report the successful electrochemical reduction of sodium metaborate into sodium borohydride, but research reports investigating the electrochemical reduction of sodium metaborate fail to reproduce and substantiate even trace amounts of sodium borohydride, as claimed by the patents. In alkaline aqueous solutions, there are a few reasons for the failure of such electrochemical conversion:
(a)
the metaborate anion BO2 is a negative ion which is repelled by the cathode surface. In addition, in the large presence of the OH ions, the HER in alkaline solutions (Equation (5)) is thermodynamically more favourable than the electro-reduction of the BO2 anion.
(b)
the BH4 is stabilized in alkaline solutions with pH of more than 12. Attempts to reduce the BO2 into BH4 require the presence of OH in concentrated solutions. However, in concentrated alkaline solutions with the presence of large amounts of OH, the effect of the HER becomes more prominent.
(c)
the presence of dissolved oxygen promotes the oxygen reduction reaction (OER, Equations (6)–(8). In addition, the presence of dissolved carbon dioxide (CO2) also promotes the electroreduction of CO2. In all the papers reviewed, there were no mention of eliminating dissolved O2 and CO2 by nitrogen or argon gas bubbling prior to the electroreduction of BO2. The analytical detection of BH4 using iodometric titraton must be abandoned altogether as it gives erroneously high concentrations, as high as eight times higher than the actual concentration of BH4 in the samples. More accurate methods of detection, such as based on the hydrolysis of the BH4 to release H2, and electrochemical methods of detection should be given priority.
Given are a few examples from the literature and suggestions to improve the efficiency of the electrochemical reduction of the BO2 into BH4. As reported in the literature, a pulsed potential of the cathode could be implemented in a divided electrochemical cell using a cationic exchange membrane (CEM). The initial positive pulse attracts the BO2 towards the electrode surface, and immediate negative pulse reduces the BO2 immediately into BH4. This method was adopted [59,60] for the in situ regeneration of BH4 from the BO2.
The second approach is to utilise electrode materials which could reduce the BO2 instead of relying on water in alkaline electrolyte as the reductant, as explained in Equations (14) and (15).
The only drawback is the presence of the smaller and more mobile OH in large concentrations compared to the BO2, which could decrease the conversion efficiency of the BO2 into BH4. Even at larger negative potentials, the presence of the HER in alkaline solutions should provide another competition for the electro-reduction of BO2 into BH4.
One possible solution of the problems of electro-reduction in alkaline aqueous solution is to perform the electro-reduction in organic solvents and molten salts such as ILs or deep eutectic solvents (DES) with the presence of such hydrogen carrier electrode materials, or with the presence of stronger chemical reductant such as hydrazine. Before attempting such procedures, preliminary investigation of the electro-reduction behaviour must be studied using cyclic voltammetry to ascertain the onset of the reduction wave and RDE to determine the D and n values in such medium.
The electro-reduction of BO2 could be the “next big thing” after the electro-reduction of dissolved carbon dioxide, as both are by-products formed from the release of energy and the recycling into their precursors is non-spontaneous with positive free energies.
Boron and carbon are in the same row and are neighbours in the periodic table, and both oxides are difficult to reduce. There are hundreds of papers reporting the photocatalytic and photo-electrocatalytic reduction of dissolved CO2 into short chain molecules, such as methane, methanol and formic acid, using semiconductor materials such as titania TiO2 and ZnO, but there is yet to be a single paper on the same subject devoted to the reduction of BO2. Nevertheless, both are important processes in the drive towards decreasing global warming and climate change, and the utilisation of hydrogen energy to replace fossil fuels.

Author Contributions

Conceptualization: W.J.B. and A.H.; methodology: W.J.B. and M.S.; validation: S.T.S. and W.J.B.; formal analysis: S.A. and N.S.J.; investigation, W.J.B.; resources: W.J.B.; data curation: S.T.S.; writing—original draft preparation: W.J.B.; writing—review and editing: W.J.B. and S.T.S.; visualization: S.A. and N.S.J.; supervision: W.J.B.; project administration: W.J.B.; funding acquisition: W.J.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Malaya, Kuala Lumpur, Malaysia, grant number PPSI-2020-CLUSTER-IDIG04.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schlesinger, H.; Brown, H.C.; Finholt, A. The preparation of sodium borohydride by the high temperature reaction of sodium hydride with borate esters. J. Am. Chem. Soc. 1953, 75, 205–209. [Google Scholar] [CrossRef]
  2. Demirci, Ü.B. Sodium borohydride for the near-future energy: A “rough diamond’’ for Turkey. Turk. J. Chem. 2018, 42, 193–220. [Google Scholar] [CrossRef]
  3. Kojima, Y.; Haga, T. Recycling process of sodium metaborate to sodium borohydride. Int. J. Hydrogen Energy 2003, 28, 989–993. [Google Scholar] [CrossRef]
  4. Li, Z.; Liu, B.; Zhu, J.; Morigasaki, N.; Suda, S. NaBH4 formation mechanism by reaction of sodium borate with Mg and H2. J. Alloys Compd. 2007, 437, 311–316. [Google Scholar] [CrossRef]
  5. Liu, B.H.; Li, Z.P.; Zhu, J.K.; Morigasaki, N.; Suda, S. Sodium borohydride synthesis by reaction of Na2O contained sodium borate with al and hydrogen. Energy Fuels 2007, 21, 1707–1711. [Google Scholar] [CrossRef]
  6. 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. Hydrogen Energy 2013, 38, 2804–2809. [Google Scholar] [CrossRef]
  7. Figen, A.K.; Pişkin, S. Microwave assisted green chemistry approach of sodium metaborate dihydrate (NabO2·2H2O) synthesis and use as raw material for sodium borohydride (NaBH4) thermochemical production. Int. J. Hydrogen Energy 2013, 38, 3702–3709. [Google Scholar] [CrossRef]
  8. Ou, T.; Panizza, M.; Barbucci, A. Thermochemical recycling of hydrolyzed NaBH4. Part II: Systematical study of parameters dependencies. Int. J. Hydrogen Energy 2013, 38, 15940–15945. [Google Scholar] [CrossRef]
  9. 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. Hydrogen Energy 2009, 34, 1717–1725. [Google Scholar] [CrossRef]
  10. Kong, L.; Cui, X.; Jin, H.; Wu, J.; Du, H.; Xiong, T. Mechanochemical synthesis of sodium borohydride by recycling sodium metaborate. Energy Fuels 2009, 23, 5049–5054. [Google Scholar] [CrossRef]
  11. 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. Hydrogen Energy 2010, 35, 4027–4040. [Google Scholar] [CrossRef]
  12. 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]
  13. Chen, W.; Ouyang, L.; Liu, J.; Yao, X.; Wang, H.; Liu, Z.; 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]
  14. 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. Hydrogen Energy 2017, 42, 13127–13135. [Google Scholar] [CrossRef]
  15. Zhong, H.; Ouyang, L.Z.; Ye, J.S.; Liu, J.W.; Wang, H.; Yao, X.D.; Zhu, M. An one-step approach towards hydrogen production and storage through regeneration of NaBH4. Energy Storage Mater. 2017, 7, 222–228. [Google Scholar] [CrossRef]
  16. Zhong, H.; Ouyang, L.; Zeng, M.; Liu, J.; Wang, H.; Shao, H.; Felderhoff, M.; Zhu, M. Realizing facile regeneration of spent NaBH4 with Mg–Al alloy. J. Mater. Chem. A 2019, 7, 10723–10728. [Google Scholar] [CrossRef] [Green Version]
  17. Qin, C.; Ouyang, L.; Wang, H.; Liu, J.; Shao, H.; Zhu, M. Regulation of high-efficient regeneration of sodium borohydride by magnesium-aluminum alloy. Int. J. Hydrogen Energy 2019, 44, 29108–29115. [Google Scholar] [CrossRef]
  18. Le, T.T.; Pistidda, C.; Puszkiel, J.; Milanese, C.; Garroni, S.; Emmler, T.; Capurso, G.; Gizer, G.; Klassen, T.; Dornheim, M. Efficient synthesis of alkali borohydrides from mechanochemical reduction of borates using magnesium–aluminum-based waste. Metals 2019, 9, 1061. [Google Scholar] [CrossRef] [Green Version]
  19. Zhu, Y.; Ouyang, L.; Zhong, H.; Liu, J.; Wang, H.; Shao, H.; Huang, Z.; Zhu, M. Closing the loop for hydrogen storage: Facile regeneration of NaBH4 from its hydrolytic product. Angew. Chem. Int. Ed. 2020, 132, 8701–8707. [Google Scholar] [CrossRef]
  20. Nunes, H.X.; Silva, D.L.; Rangel, C.M.; Pinto, A.M. Rehydrogenation of sodium borates to close the NaBH4-H2 cycle: A review. Energies 2021, 14, 3567. [Google Scholar] [CrossRef]
  21. Šljukić, B.; Santos, D.M.; Sequeira, C.A.; Banks, C.E. Analytical monitoring of sodium borohydride. Anal. Methods 2013, 5, 829–839. [Google Scholar] [CrossRef]
  22. Santos, D.; Sequeira, C. On the electrosynthesis of sodium borohydride. Int. J. Hydrogen Energy 2010, 35, 9851–9861. [Google Scholar] [CrossRef]
  23. Çelikkan, H.; Şahin, M.; Aksu, M.L.; Veziroğlu, T.N. The investigation of the electrooxidation of sodium borohydride on various metal electrodes in aqueous basic solutions. Int. J. Hydrogen Energy 2007, 32, 588–593. [Google Scholar] [CrossRef]
  24. Amendola, S.; Onnerud, P.; Kelly, M.T.; Binder, M. Inexpensive, in situ monitoring of borohydride concentrations. Talanta 1999, 49, 267–270. [Google Scholar] [CrossRef]
  25. Elamin, B.; Means, G.E. Spectrophotometric determination of sodium borohydride with 2,4,6-trinitrobenzenesulfonic acid. Anal. Chim. Acta. 1979, 107, 405–409. [Google Scholar] [CrossRef]
  26. Gyenge, E.; Oloman, C. Electrosynthesis attempts of tetrahydridoborates. J. Appl. Electrochem. 1998, 28, 1147–1151. [Google Scholar] [CrossRef]
  27. Park, E.H.; Jeong, S.U.; Jung, U.H.; Kim, S.H.; Lee, J.; Nam, S.W.; Lim, T.H.; Park, Y.J.; Yu, Y.H. Recycling of sodium metaborate to borax. Int. J. Hydrogen Energy 2007, 32, 2982–2987. [Google Scholar] [CrossRef]
  28. Mirkin, M.V.; Bard, A.J. Voltammetric method for the determination of borohydride concentration in alkaline aqueous solutions. Anal. Chem. 1991, 63, 532–533. [Google Scholar] [CrossRef]
  29. Mirkin, M.V.; Yang, H.; Bard, A.J. Borohydride oxidation at a gold electrode. J. Electrochem. Soc. 1992, 139, 2212. [Google Scholar] [CrossRef]
  30. Sanlı, E.; Çelikkan, H.; Uysal, B.Z.; Aksu, M.L. Anodic behavior of Ag metal electrode in direct borohydride fuel cells. Int. J. Hydrogen Energy 2006, 31, 1920–1924. [Google Scholar] [CrossRef]
  31. Sanli, E.; Uysal, B.Z.; Aksu, M.L. The oxidation of NaBH4 on electrochemicaly treated silver electrodes. Int. J. Hydrogen Energy 2008, 33, 2097–2104. [Google Scholar] [CrossRef]
  32. Santos, D.; Sequeira, C. Cyclic voltammetry investigation of borohydride oxidation at a gold electrode. Electrochim. Acta 2010, 55, 6775–6781. [Google Scholar] [CrossRef]
  33. Chatenet, M.; Molina-Concha, M.; Diard, J.-P. First insights into the borohydride oxidation reaction mechanism on gold by electrochemical impedance spectroscopy. Electrochim. Acta 2009, 54, 1687–1693. [Google Scholar] [CrossRef]
  34. Gyenge, E. Electrooxidation of borohydride on platinum and gold electrodes: Implications for direct borohydride fuel cells. Electrochim. Acta 2004, 49, 965–978. [Google Scholar] [CrossRef]
  35. Gyenge, E.; Atwan, M.; Northwood, D. Electrocatalysis of borohydride oxidation on colloidal Pt and Pt-alloys (Pt-Ir, Pt-Ni, and Pt-Au) and application for direct borohydride fuel cell anodes. J. Electrochem. Soc. 2005, 153, A150–A158. [Google Scholar] [CrossRef]
  36. Liu, B.H.; Li, Z.P.; Suda, S. Electrocatalysts for the anodic oxidation of borohydrides. Electrochim. Acta 2004, 49, 3097–3105. [Google Scholar] [CrossRef]
  37. Liu, B.H.; Li, Z.P.; Suda, S. Anodic oxidation of alkali borohydrides catalyzed by nickel. J. Electrochem. Soc. 2003, 150, A398. [Google Scholar] [CrossRef]
  38. Okinaka, Y. An electrochemical study of electroless gold-deposition reaction. J. Electrochem. Soc. 1973, 120, 739. [Google Scholar] [CrossRef]
  39. Dong, H.; Feng, R.; Ai, X.; Cao, Y.; Yang, H.; Cha, C. Electrooxidation mechanisms and discharge characteristics of borohydride on different catalytic metal surfaces. J. Phys. Chem. B 2005, 109, 10896–10901. [Google Scholar] [CrossRef]
  40. Colominas, S.; McLafferty, J.; Macdonald, D.D. Electrochemical studies of sodium borohydride in alkaline aqueous solutions using a gold electrode. Electrochim. Acta 2009, 54, 3575–3579. [Google Scholar] [CrossRef]
  41. Chatenet, M.; Micoud, F.; Roche, I.; Chainet, E. Kinetics of borohydride oxidation and oxygen reduction in sodium hydroxide electrolyte: Part 1. BH4 electrooxidation on Au and Ag catalysts. Electrochim. Acta 2006, 51, 5459–5467. [Google Scholar] [CrossRef]
  42. Gardiner, J.A.; Collat, J.W. Polarography of the tetrahydroborate ion. The effect of hydrolysis on the system. Inorg. Chem. 1965, 4, 1208–1212. [Google Scholar] [CrossRef]
  43. Wang, K.; Lu, J.; Zhuang, L. Direct determination of diffusion coefficient for borohydride anions in alkaline solutions using chronoamperometry with spherical au electrodes. J. Electroanal. Chem. 2005, 585, 191–196. [Google Scholar] [CrossRef]
  44. Chatenet, M.; Molina-Concha, M.; El-Kissi, N.; Parrour, G.; Diard, J.-P. Direct rotating ring-disk measurement of the sodium borohydride diffusion coefficient in sodium hydroxide solutions. Electrochim. Acta 2009, 54, 4426–4435. [Google Scholar] [CrossRef]
  45. Gardiner, J.A.; Collat, J.W. Kinetics of the stepwise hydrolysis of tetrahydroborate ion. J. Am. Chem. Soc. 1965, 87, 1692–1700. [Google Scholar] [CrossRef]
  46. Pecsok, R.L. Polarographic studies on the oxidation and hydrolysis of sodium borohydride1. J. Am. Chem. Soc. 1953, 75, 2862–2864. [Google Scholar] [CrossRef]
  47. Cheng, H.; Scott, K. Determination of kinetic parameters for borohydride oxidation on a rotating au disk electrode. Electrochim. Acta 2006, 51, 3429–3433. [Google Scholar] [CrossRef]
  48. Atwan, M.H.; Macdonald, C.L.; Northwood, D.O.; Gyenge, E.L. Colloidal au and au-alloy catalysts for direct borohydride fuel cells: Electrocatalysis and fuel cell performance. J. Power Sources 2006, 158, 36–44. [Google Scholar] [CrossRef]
  49. Cooper, H. Electrolytic Process for the Production of Alkali Metal Borohydrides. U.S. Patent 3734842, 22 May 1973. [Google Scholar]
  50. Sharifian, H.; Dutcher, J.S. Production of Quaternary Ammonium and Quaternary Phosphonium Borohydrides. U.S. Patent 3734842, 27 February 1990. [Google Scholar]
  51. Hale, C.H.; Sharifian, H. Production of Metal Borohydrides and Organic Onium Borohydrides. U.S. Patent 4931154, 5 June 1990. [Google Scholar]
  52. McLafferty, J.; Colominas, S.; Macdonald, D. Attempts to cathodically reduce boron oxides to borohydride in aqueous solution. Electrochim. Acta 2010, 56, 108–114. [Google Scholar] [CrossRef]
  53. Amendola, S. Electroconversion Cell. U.S. Patent 6,497,973 B1, 24 December 2002. [Google Scholar]
  54. Kawai, M.; Ito, M. Japanese Patent 2003-247088, 2003.
  55. Wang, J.; Sun, Y.; Liang, Z. Preliminary study on electrochemical reduction of sodium metaborate to produce sodium borohydride. J. Taiyuan Univ. Tech. 2006, 37, 539. [Google Scholar]
  56. Sanli, A.E.; Kayacan, İ.; Uysal, B.Z.; Aksu, M.L. Recovery of borohydride from metaborate solution using a silver catalyst for application of direct rechargable borohydride/peroxide fuel cells. J. Power Sources 2010, 195, 2604–2607. [Google Scholar] [CrossRef]
  57. Shen, Y.; Sun, T.; Jia, J. A novel desulphurization process of coal water slurry via sodium metaborate electroreduction in the alkaline system. Fuel 2012, 96, 250–256. [Google Scholar] [CrossRef]
  58. Zhu, Q.-Y.; Zhu, C.-G.; Wang, F.-W.; Wei, Y.-J. Preparation of sodium borohydride by copper electrolysis. Asian J. Chem. 2013, 25, 7749–7752. [Google Scholar] [CrossRef]
  59. Shu, C.; Sun, T.; Jia, J.; Lou, Z. A novel desulfurization process of gasoline via sodium metaborate electroreduction with pulse voltage using a boron-doped diamond thin film electrode. Fuel 2013, 113, 187–195. [Google Scholar] [CrossRef]
  60. Shu, C.; Lai, F.; Zhu, F.; Luo, D. Optimal extractive and reductive desulfurization process using sodium borohydride in situ generated via sodium metaborate electroreduction in ionic liquid. Chem. Eng. Process.-Process Intensif. 2020, 150, 107869. [Google Scholar] [CrossRef]
Figure 1. Cyclic voltammetry 1 V s−1 on Au electrode in 2 M NaOH + 0.09 M NaBH4 solution at 25 °C. Reproduced from ref. [32] with permission from the publisher.
Figure 1. Cyclic voltammetry 1 V s−1 on Au electrode in 2 M NaOH + 0.09 M NaBH4 solution at 25 °C. Reproduced from ref. [32] with permission from the publisher.
Energies 16 00015 g001
Figure 2. Cyclic voltammograms of (A) 2 M NaOH and (B) 1 mM NaBH4 in 2 M NaOH at 6 mm Au disk at 0.100 V s−1. Reproduced from ref. [40] with permission from the publisher.
Figure 2. Cyclic voltammograms of (A) 2 M NaOH and (B) 1 mM NaBH4 in 2 M NaOH at 6 mm Au disk at 0.100 V s−1. Reproduced from ref. [40] with permission from the publisher.
Energies 16 00015 g002
Figure 3. Cyclic voltammetry (CV) at the gold disk (0.124 cm2) of the RRDE in (A) supporting electrolyte (1 M NaOH) for Ω = 0) and in (B) 0.1 M NaOH + 10−3 M NaBH4 at 25 °C for several revolution rates (Ω) of the RDE. Reproduced from ref. [44] with permission from the publisher.
Figure 3. Cyclic voltammetry (CV) at the gold disk (0.124 cm2) of the RRDE in (A) supporting electrolyte (1 M NaOH) for Ω = 0) and in (B) 0.1 M NaOH + 10−3 M NaBH4 at 25 °C for several revolution rates (Ω) of the RDE. Reproduced from ref. [44] with permission from the publisher.
Energies 16 00015 g003
Figure 4. Electrochemical cell for bulk electrolysis of NaBO2 and open circuit potential (OCP) measurements in alkaline solution at 25 °C. The NaBO2 is reduced at the cadmium cathode while a high impedance voltmeter prevents current to flow into the circuit between the indicator and reference electrodes for the OCP monitoring. Reproduced from ref. [22] with permission from the publisher.
Figure 4. Electrochemical cell for bulk electrolysis of NaBO2 and open circuit potential (OCP) measurements in alkaline solution at 25 °C. The NaBO2 is reduced at the cadmium cathode while a high impedance voltmeter prevents current to flow into the circuit between the indicator and reference electrodes for the OCP monitoring. Reproduced from ref. [22] with permission from the publisher.
Energies 16 00015 g004
Figure 5. (A) Cyclic voltammogram of a BDD electrode in 0.1 M NaOH and 0.1 M NaOH + 0.15 M NaBO2 at 0.100 V s−1; and (B) 11 boron NMR spectrogram of electrolytes before and after electroreduction, reaction conditions: 1.5 V forward pulse voltage, 0.3 V reverse pulse voltage, 1.5 s forward pulse duration, 0.5 s reverse pulse duration, 0.15 M NaBO2 concentration, 0.1 M NaOH, 1.5 h of electrolytic time. Reproduced from ref. [59] with permission from the publisher.
Figure 5. (A) Cyclic voltammogram of a BDD electrode in 0.1 M NaOH and 0.1 M NaOH + 0.15 M NaBO2 at 0.100 V s−1; and (B) 11 boron NMR spectrogram of electrolytes before and after electroreduction, reaction conditions: 1.5 V forward pulse voltage, 0.3 V reverse pulse voltage, 1.5 s forward pulse duration, 0.5 s reverse pulse duration, 0.15 M NaBO2 concentration, 0.1 M NaOH, 1.5 h of electrolytic time. Reproduced from ref. [59] with permission from the publisher.
Energies 16 00015 g005
Figure 6. Left: The schematic diagram of the novel desulfurization process. Right: Cyclic voltammogram of IL and IL + NaBO2 on a glassy carbon electrode at 0.100 V s−1. The green arrows show the movement of the molecule. Reproduced from ref. [60] with permission from the publisher.
Figure 6. Left: The schematic diagram of the novel desulfurization process. Right: Cyclic voltammogram of IL and IL + NaBO2 on a glassy carbon electrode at 0.100 V s−1. The green arrows show the movement of the molecule. Reproduced from ref. [60] with permission from the publisher.
Energies 16 00015 g006
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

Basirun, W.J.; Shah, S.T.; Shalauddin, M.; Akhter, S.; Jamaludin, N.S.; Hayyan, A. A Review of Electrochemical Reduction of Sodium Metaborate. Energies 2023, 16, 15. https://doi.org/10.3390/en16010015

AMA Style

Basirun WJ, Shah ST, Shalauddin M, Akhter S, Jamaludin NS, Hayyan A. A Review of Electrochemical Reduction of Sodium Metaborate. Energies. 2023; 16(1):15. https://doi.org/10.3390/en16010015

Chicago/Turabian Style

Basirun, Wan Jefrey, Syed Tawab Shah, Md. Shalauddin, Shamima Akhter, Nazzatush Shimar Jamaludin, and Adeeb Hayyan. 2023. "A Review of Electrochemical Reduction of Sodium Metaborate" Energies 16, no. 1: 15. https://doi.org/10.3390/en16010015

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