You are on page 1of 13

energies

Article
Direct Synthesis of NaBH4 Nanoparticles from
NaOCH3 for Hydrogen Storage
Ting Wang and Kondo-Francois Aguey-Zinsou *
MERLin, School of Chemical Engineering, The University of New South Wales, Sydney, NSW 2052, Australia;
ting.wang@unsw.edu.au
* Correspondence: f.aguey@unsw.edu.au; Tel.: +61-2-9385-7970

Received: 28 October 2019; Accepted: 18 November 2019; Published: 21 November 2019 

Abstract: Hydrogen is regarded as a promising energy carrier to substitute fossil fuels. However,
storing hydrogen with high density remains a challenge. NaBH4 is a potential hydrogen storage
material due to its high gravimetric hydrogen density (10.8 mass%), but the hydrogen kinetic and
thermodynamic properties of NaBH4 are poor against the application needs. Nanosizing is an
effective strategy to improve the hydrogen properties of NaBH4 . In this context, we report on the
direct synthesis of NaBH4 nanoparticles (~6–260 nm) from the NaOCH3 precursor. The hydrogen
desorption properties of such nanoparticles are reported as well as experimental conditions that lead
to the synthesis of (Na2 B12 H12 ) free NaBH4 nanoparticles.

Keywords: sodium borohydride; nanoparticles; synthesis; hydrogen storage

1. Introduction
Sodium borohydride (NaBH4 ) has attracted significant interest as a potential hydrogen storage
material, owing to its high hydrogen gravimetric capacity (10.8 mass%) [1,2]. However, the practical
application of NaBH4 is difficult to achieve, because of its stability both in terms of hydrogen kinetics
and thermodynamics. Typically, the dehydrogenation temperature of NaBH4 is above 500 ◦ C and the
kinetics of both the hydrogen release and uptake are too slow for practical use. Moreover, the inherent
phase segregation of NaBH4 and partial evaporation of Na upon hydrogen release lead to poor
hydrogen reversibility [3].
In order to enhance the hydrogen storage properties of NaBH4 , the approach of nanosizing, i.e.,
reduction of particle size below 100 nm, has proven to lead to some improvements in the thermodynamics
and kinetics. This is believed to result from several factors including an increase in the specific surface
area and shorter hydrogen diffusion paths [4,5]. The typical method to downsize borohydrides is
through their nanoconfinement in porous scaffolds [6,7], but the effective storage capacity achievable
by such an approach remains limited, owing to the difficulty of filling the nanopores and the “dead
volume and mass” of the host scaffold [4,8]. An alternative method is through the synthesis of isolated
nanoparticles and their stabilization as core-shell nanostructures [3,9]. However, to date, this approach
is significantly limited by the lack of synthetic methods for readily making NaBH4 nanoparticles of
controlled properties.
Thus far, the synthesis methods for preparing isolated nanosized NaBH4 particles are rather
limited [3,5]. A common method to synthesize nanoparticles is by growth in solution, but for NaBH4 ,
this is not without challenges. NaBH4 is a strong reducing agent, and thus, many of the common
reagents used for the synthesis and stabilization of conventional nanoparticles are out of scope.
In addition, the solvents that can be used to synthesize NaBH4 nanoparticles are rather limited owing
to the lack of solubility and/or reactivity of NaBH4 [2,3]. Oftentimes, solubilizing solvents form
adduct compounds with NaBH4 , and this brings additional complexity in terms of synthesizing

Energies 2019, 12, 4428; doi:10.3390/en12234428 www.mdpi.com/journal/energies


Energies 2019, 12, 4428 2 of 13

solvent free nanoparticles [10]. Via a wet chemistry approach, Christian et al. reported that NaBH4
nanoparticles could be precipitated in solution with a size ranging from 2 to 60 nm via anti-precipitation
with tetra-n-butylammonium bromide (TBAB) as a surfactant [1,5]. Solvent evaporation has also
been reported as an effective method for the synthesis of borohydride nanoparticles. For example,
Lai et al. have synthesized LiBH4 and NaBH4 nanoparticles with Ti doped by solvent evaporation [11]
and we have also successfully synthesized LiBH4 nanoparticles of controllable sizes with various
surfactants [5]. However, this approach leaves excessive amounts of surfactant around the nanoparticles.
This significantly limits the effective hydrogen storage capacity and contaminates the hydrogen release
with fragments of decomposing surfactants at elevated temperatures [5]. Removing such surfactants
could be an option, but these are needed to minimize the surface energy of nanoparticles during their
nucleation and growth.
A major drawback of current synthetic approaches is that they all rely on a source of NaBH4 as
the starting material to lead to the formation of nanoparticles, and this is an expensive approach given
the current cost of NaBH4 (95 USD$ kg−1 ) [2]. In this respect, the development of direct approaches for
making NaBH4 nanoparticles from common precursors is more desirable. However, this is not without
difficulties. As summarized in Table S1, the known synthetic methods of NaBH4 have drawbacks,
such as: a) low yields, b) high temperatures and/or hydrogen pressures required, and c) low purity.
Furthermore, direct production of nanoparticles from these synthetic methods is not evident, and to
date, only Lai et al. reported on a thermal decomposition of sodium trimethyl borate (NaB(OCH3 )4 ) to
make NaBH4 nanoparticles. In this case, the approach first involves the formation of NaB(OCH3 )4
through the reaction of NaH with trimethyl borate (B(OCH3 )3 ) via reaction (1). This is then followed by
the thermal decomposition of NaB(OCH3 )4 in tetrahydrofuran to precipitate NaBH4 (reaction (2)) [3]:

NaH + B(OCH3 )3 → NaBH(OCH3 )3 (1)

65−70 ◦ C
4NaBH(OCH3 )3 −−−−−−−→ NaBH4 + 3NaB(OCH3 )4 (2)

However, the method was found to be difficult to control, the yield of NaBH4 was low, and the
by-products of the reaction was difficult to remove [9].
In this work, we report on the development of two routes for making isolated NaBH4 nanoparticles
through the reaction of NaOCH3 with B2 H6 gas according to reaction (3):

3NaOCH3 + 2B2 H6 → 3NaBH4 + B(OCH3 )3 (3)

Considering that B(OCH3 )3 has a low boiling point (68 ◦ C) [12,13] and it is soluble in diethyl ether,
in which NaBH4 is not soluble [14–16], B(OCH3 )3 as a by-product can be removed by heating up or by
diethyl ether washing. Thus, we investigated both routes for making NaBH4 nanoparticles, and in
first instance, reaction (3) was carried out at 100 ◦ C to facilitate the evaporation of B(OCH3 )3 , while
B2 H6 was reacted with NaBH4 . To facilitate the dispersion of the reaction by-product, the feasibility of
reaction (3) in solvents was also investigated. The morphology, purity and hydrogen properties of the
NaBH4 nanoparticles obtained through these routes are reported. In particular, the main hydrogen
release temperature of the NaBH4 nanoparticles synthesized by both methods has been reduced by
more than 100 ◦ C in comparison to commercial NaBH4 .

2. Materials and Methods


Except the solid-gas synthesis of NaBH4 nanoparticles, all the other experiments were performed
under an inert atmosphere in an argon filled LC-Technology glove box (O2 and H2 O < 1 ppm).
Energies 2019, 12, 4428 3 of 13

2.1. Materials
NaOCH3 (95%), NaBH4 (98%), ZnCl2 (≥98%), tetra-n-butylammonium bromide (TBAB, >98%),
methanol (anhydrous, 99.8%), and octadecylamine (ODA, ≥99%) were purchased from Sigma-Aldrich
and hexane (99%, HPLC grade) was purchased from Honeywell. NaOCH3 was purified by solubilization
in a large amount of anhydrous methanol and recrystallization of the centrifuged solution was done by
evaporating the methanol. ZnCl2 was further dried at 120 ◦ C overnight on a Schlenk line under vacuum
(0.01 MPa). Hexane was dried with activated molecular sieve (4 Å) prior to use. The other chemicals
were used as received.

2.2. Synthesis of the Diborane (B2 H6 ) Precursor


NaZn2 (BH4 )5 as the precursor of B2 H6 was synthesized by ball milling NaBH4 and ZnCl2 (2:1 molar
ratio) in a stainless-steel vial (25 mL) containing a single stainless-steel ball (15 mm diameter). A Retsch
MM301 mill operated at a frequency of 20Hz was used for this. The milling process was done twice
for 10 min.

2.3. Synthesis of the NaBH4 Nanoparticles via Solid-Gas Reaction


The reaction was undertaken in an in-house built Sievert apparatus with a customized sample
holder which had two compartments separated by a stainless-steel mesh (20 µm porosity). 100 mg
of purified NaOCH3 was placed on the top of the mesh and variable amounts (0.2, 0.3, 0.4, 1.0 g) of
NaZn2 (BH4 )5 were placed at the bottom of the sample holder. The reaction was carried out at 100 ◦ C
overnight. The materials synthesized by this solid-gas reaction (~72 mg) are denoted NaBH4 -SGX,
where X is the molar ratio of NaOCH3 and the B2 H6 generated, X = 1.2, 1.8, 2.4, 5.9.

2.4. Synthesis of the NaBH4 Nanoparticles via Suspension in Hexane and Reaction with B2 H6
NaOCH3 nanoparticles were first synthesized by a solvent evaporation method. To this aim,
0.2 M of purified NaOCH3 and 0.01 M TBAB acting as surfactant were dissolved in 10 mL methanol
at room temperature. The solution was then placed in a Biotage V-10 evaporator and the methanol
was evaporated at a controlled rate to trigger the nucleation and growth of NaOCH3 nanoparticles.
The evaporation was carried out at 50 ◦ C, at a centrifugal speed of 3000 rpm. The resulting NaOCH3
powered was then further dried overnight on a Schlenck line at 40 ◦ C and under a vacuum level of
0.01 MPa to remove any methanol residue.
70 mg of as-synthesized NaOCH3 nanoparticles were then suspended in 10 mL hexane with
0.01 M octadecylamine acting as surfactant in a round flask. Variable amount of NaZn2 (BH4 )5 (0.5, 1,
and 2 g) were placed in a separate round flask on a hot plate at 100 ◦ C to generate B2 H6 . It is important
to mention that the amount of B2 H6 is much larger than that used in the solid-gas reaction. This is to
compensate for the weak airtightness of the rubber seal on the glassware. Both round-bottomed flasks
were sealed by a rubber septum and connected by a cannula to allow for the B2 H6 generated from the
decomposition of NaZn2 (BH4 )5 to react with the NaOCH3 nanoparticles in suspension. After reaction
overnight, the resulting material was separated from the hexane by centrifugation, washed three
times with diethyl ether to remove the by-product B(OCH3 )3 , and dried on a Schlenck line at room
temperature under vacuum at 0.01 MPa. The materials synthesized for this suspension gas reaction
(~42 mg) are denoted NaBH4 -SUGX, where X is the molar ratio of NaOCH3 and the B2 H6 generated,
X = 4.2, 8.5, 16.9.

2.5. Characterization
The morphologies of the materials were determined by Transmission Electron Microscopy (TEM)
on a Philips CM200 operated at 200 kV. The materials were dispersed in cyclohexane followed by
ultrasonication and then dropped onto a carbon coated copper grid. The copper grid was then
Energies 2019, 12, 4428 4 of 13

transferred to the TEM in a quick manner as to minimize air exposure. The mean particle size and
distribution were collected from a minimum of 300 particles from a range of TEM images.
Crystalline phases were determined by X-Ray Diffraction (XRD) on a Philips X’pert Multipurpose
XRD system operated at 40 mA and 45 kV with a monochromated Cu Kα radiation (λ = 1.541 Å) over
a 2θ range from 10 to 70◦ . The materials were protected against oxidation from air by a Kapton foil.
Fourier Transform Infrared Spectroscopy (FTIR) was conducted on a Bruker Vertex 70 V.
The materials were mixed with KBr in the glove box and loaded in an air-tight chamber fitted
on a Harrick-Scientific Praying Mantis Diffuse Reflectance Infrared Fourier Transform Spectroscopy
(DRIFT) accessory. The spectra were collected at room temperature from 600 to 3400 cm−1 over
124 scans with a resolution of 1 cm−1 .
The hydrogen desorption properties of the materials were determined by Mass Spectrometry (MS)
with an Omnistar instrument from Pfeiffer. The thermal stability of the surfactants was investigated
by simultaneous Thermogravimetric Analysis/Differential Scanning Calorimetry (TGA/DSC) with a
TGA/DSC1 from Mettler Toledo coupled to the MS. The characterizations were conducted under an
argon flow at 20 mL min−1 from 20 to 700 ◦ C with a heating rate of 10 ◦ C min−1 and alumina crucibles
(70 mg) were used.
The carbon content of the synthesized materials (NaBH4 -SG1.8 and NaBH4 -SUG8.5) that may
result from remaining traces of NaOCH3 or B(OCH3 )3 was determined with a total carbon analyzer
(Multi N/C2100, Analytic Jena, Germany). The materials were dissolved in 10 mL of Milli-Q water
and the resulting solution was filtered before total carbon analysis. Based on this analysis, both
NaBH4 -SG1.8 and NaBH4 -SUG8.5 (which correspond to the optimum reaction conditions) were found
to contain ~2.0 % of carbon per weight of material.

3. Results and Discussion

3.1. NaBH4 Nanoparticles Synthesized via Solid-Gas Interaction

3.1.1. Morphology of the As-Synthesized Materials


The morphology of the as-synthesized material was characterized by TEM (Figure 1). By exposing
the NaOCH3 particles (Figure S1) to increasing amounts of B2 H6 , (from molar ratio 1.8 to 5.9),
the morphology of the material synthesized evolved from large undefined particles of ~262 nm to
spherical like particles of 6 nm. This reduction in size is probably due to a fragmentation of the
NaOCH3 particles with B2 H6 (reaction 3), and the subsequent release of gaseous B(OCH3 )3 at 100 ◦ C.
This phenomenon has also been found to occur in reactions involving some gas release [17,18].

3.1.2. Composition of the As-Synthesized Materials


XRD and FTIR characterizations were performed to confirm the compositions of the as-synthesized
material. From XRD analysis (Figure 2 and Figure S2), it is apparent that when the molar ratio of
NaOCH3 and B2 H6 is above 1.8, all the NaOCH3 is reacted because only diffraction peaks related to
thermodynamically stable phase of NaBH4 were observed. FTIR analysis further confirmed this trend
(Figure 3 and Figure S3). As the molar ratio evolved from 1.2 to 1.8, the vibrational peaks corresponding
to NaOCH3 (Figure S3) disappeared, and this is in agreement with the XRD results (Figure 2). The peak
at ~43 ◦ is assigned to the stainless steel sample holder.
A typical FTIR spectrum of NaBH4 would display vibrational peaks associated with the BH4 −
group in the region of 2410–2200 cm−1 corresponding to the stretching mode of BH4 − , whereas the
bending modes of BH4 − are to be found around 1120 cm−1 [19,20].
For NaBH4 -SG1.2, the additional peaks observed at 2582, 1194, 1077, and 994 cm−1 were assigned
to the C-O stretching modes associated with NaOCH3 [21–29]. Such vibrations were not observed at
higher molar ratio of NaOCH3 and B2 H6 and this confirmed that above a ratio of 1.2, NaOCH3 was fully
reacted. However, when the molar ratio reached 2.4 and 5.9, a new vibration appeared at 2490 cm−1
Energies 2019, 12, 4428 5 of 13

(Figure 3). This vibration was assigned to the stretching mode of B12 H12 2− [11,30]. Accordingly, traces
of Na2 B12 H12 have been generated via the reaction of the newly formed NaBH4 with the excess of
B2 H6 . Thus far, there have been no reports regarding the direct formation of Na2 B12 H12 via the reaction
between NaBH4 and B2 H6 . Only the formation of NaB3 H8 has been reported to occur according
to (4) [31,32]:
NaBH4 + B2 H6 → NaB3 H8 + H2 (4)
Energies 2019, 12, x FOR PEER REVIEW 5 of 14

Figure 1. Transmission Electron


Figure 1. Transmission Microscopy
Electron Microscopy(TEM)
(TEM)images
images ofof(a),
(a,b)
(b) NaBH 4 -SG1.2;
NaBH4-SG1.2; (c), (c,d) NaBH
(d) NaBH 4- 4 -SG1.8;

(e,f) NaBHSG1.8;
4 (e),
-SG2.4; (f) NaBH
and 4-SG2.4;
(g,h) NaBHand (g),
4 (h)
-SG5.9.NaBH 4-SG5.9.
The inset The inset
diagrams diagrams
are are
the the corresponding
corresponding mean
mean particle
particle
sizes and size sizes and size distribution.
distribution.
synthesized material. From XRD analysis (Figures 2 and S2), it is apparent that when the molar ratio
of NaOCH3 and B2H6 is above 1.8, all the NaOCH3 is reacted because only diffraction peaks related
to thermodynamically stable phase of NaBH4 were observed. FTIR analysis further confirmed this
trend (Figures 3 and S3). As the molar ratio evolved from 1.2 to 1.8, the vibrational peaks
corresponding to NaOCH3 (Figure S3) disappeared, and this is in agreement with the XRD results
Energies 2019, 12,
(Figure 2). 4428
The peak at ~43 ° is assigned to the stainless steel sample holder. 6 of 13

Energies 2019, 12, x FOR PEER REVIEW 7 of 14

nanoparticles. This was also confirmed by the presence of C-H stretching modes from B(OCH3)3 in
all the synthesized materials (Figure 3).
Figure 2. XRD patterns the as-synthesized NaBH4 via solid-gas interaction.
Figure 2. XRD patterns the as-synthesized NaBH4 via solid-gas interaction.

A typical FTIR spectrum of NaBH4 would display vibrational peaks associated with the BH4-
group in the region of 2410–2200 cm-1 corresponding to the stretching mode of BH4-, whereas the
bending modes of BH4- are to be found around 1120 cm-1 [19,20].
For NaBH4-SG1.2, the additional peaks observed at 2582, 1194, 1077, and 994 cm-1 were assigned
to the C-O stretching modes associated with NaOCH3 [21–29]. Such vibrations were not observed at
higher molar ratio of NaOCH3 and B2H6 and this confirmed that above a ratio of 1.2, NaOCH3 was
fully reacted. However, when the molar ratio reached 2.4 and 5.9, a new vibration appeared at 2490
cm-1 (Figure 3). This vibration was assigned to the stretching mode of B12H122- [11,30]. Accordingly,
traces of Na2B12H12 have been generated via the reaction of the newly formed NaBH4 with the excess
of B2H6. Thus far, there have been no reports regarding the direct formation of Na2B12H12 via the
reaction between NaBH4 and B2H6. Only the formation of NaB3H8 has been reported to occur
according to (4) [31,32]:
NaBH4 + B2H6 → NaB3H8 + H2 (4)
Chen et al. reported that NaB3H8 can decompose to NaBH4, Na2B12H12, Na2B10H10 at 150 °C [33].
Given that our synthesis was performed at 100 °C, one possibility is that the Na2B12H12 phase observed
by FTIR is the result of the decomposition of NaB3H8.
Finally, the broad peaks between 1420–1260 cm-1 that were observed in all the spectra of the as-
synthesized NaBH4 nanoparticles were assigned to the stretching modes of B-O bond [26]. These
vibrations areFigure
most likely
3. FTIRthe resultthe
spectra of remaining traces
as-synthesized of B(OCH
NaBH 3)3 in the as-synthesized NaBH4
via solid-gas interaction.
4

Figure 3. FTIR spectra the as-synthesized NaBH4 via solid-gas interaction.


Chen et al. reported that NaB3 H8 can decompose to NaBH4 , Na2 B12 H12 , Na2 B10 H10 at 150 ◦ C [33].
Given
3.2. that
NaBH our synthesis was performed at 100 ◦ C, one possibility is that the Na2 B12 H12 phase observed
4 Nanoparticles Synthesized from NaOCH3 Nanoparticles Reacted in Suspension
by FTIR is the result of the decomposition of NaB3 H8 .
Although NaBH4 nanoparticles have been successfully obtained by the solid-gas method, the
Finally, the broad peaks between 1420–1260 cm−1 that were observed in all the spectra of the
approach did not lead to nanoparticles of well-defined morphology. In order to better control the
as-synthesized NaBH
particle size and 4 nanoparticles
synthesize were assigned
isolated NaBH to the stretching modes of B-O bond [26]. These
4 nanoparticles, we investigated the possibility to first
vibrations
synthesize NaOCH3 nanoparticles by solvent evaporationofmethod,
are most likely the result of remaining traces B(OCH3and )3 inthen
the as-synthesized
suspended theseNaBH4
nanoparticles. This was also confirmed by the presence of C-H stretching modes from B(OCH
NaOCH3 nanoparticles in hexane with octadecylamine as a stabilizer to retain the morphology during3 )3 in all
thereaction
synthesized
with Bmaterials (Figureat
2H6. After reaction 3).room temperature, the as-synthesized NaBH4 nanoparticles were
collected by centrifugation and washed by diethyl ether to remove the solubilized B(OCH3)3 residue
3.2.[14–16].
NaBH4Figure
Nanoparticles
4a and 4bSynthesized from NaOCH
shows the typical 3 Nanoparticles
TEM images Reacted inNaOCH
of the as-synthesized Suspension
3 nanoparticles

obtained by solvent evaporation method, and thus, such an approach was found to be effective in
Although NaBH4 nanoparticles have been successfully obtained by the solid-gas method,
leading to nanoparticles of NaOCH3 with a mean particle size of 16 nm.
the approach did not lead to nanoparticles of well-defined morphology. In order to better control
3.2.1. Morphology of the As-synthesized NaBH4
After the reaction of the suspended NaOCH3 nanoparticles with B2H6, analysis by TEM of the
resulting material revealed polydispersed particles with a mean size of 14 nm (NaBH4-SUG4.2, Figure
4c and 4d) and 12 nm (NaBH4-SUG8.5, Figure 4e, 4f, and NaBH4-SUG16.9, Figure 4g, 4h). This is
slightly smaller than the particle size of NaOCH3 (Figure 4a and 4b), and this reduction in particle
Energies 2019, 12, 4428 7 of 13

the particle size and synthesize isolated NaBH4 nanoparticles, we investigated the possibility to first
synthesize NaOCH3 nanoparticles by solvent evaporation method, and then suspended these NaOCH3
nanoparticles in hexane with octadecylamine as a stabilizer to retain the morphology during reaction
with B2 H6 . After reaction at room temperature, the as-synthesized NaBH4 nanoparticles were collected
by centrifugation and washed by diethyl ether to remove the solubilized B(OCH3 )3 residue [14–16].
Figure 4a,b shows the typical TEM images of the as-synthesized NaOCH3 nanoparticles obtained
Energies 2019, 12, x FOR PEER REVIEW 8 of 14
by solvent evaporation method, and thus, such an approach was found to be effective in leading to
nanoparticles of NaOCH3 with a mean particle size of 16 nm.

Figure 4. Figure
TEM 4.image
TEM image of the
of the as-synthesized (a),
as-synthesized (b) NaOCH
(a,b) NaOCH3 nanoparticles;(c) (d) NaBH4-SUG4.2; (e)
3 nanoparticles; (c,d) NaBH4 -SUG4.2;
(f) NaBH4-SUG8.5; and (g) (h) NaBH4-SUG16.9. The inset diagrams are the corresponding mean
(e,f) NaBH4 -SUG8.5; and (g,h) NaBH4 -SUG16.9. The inset diagrams are the corresponding mean
particle sizes and size distribution.
particle sizes and size distribution.
Energies 2019, 12, 4428 8 of 13

3.2.1. Morphology of the As-Synthesized NaBH4


After the reaction of the suspended NaOCH3 nanoparticles with B2 H6 , analysis by TEM of
the resulting material revealed polydispersed particles with a mean size of 14 nm (NaBH4 -SUG4.2,
Figure 4c,d) and 12 nm (NaBH4 -SUG8.5, Figure 4e,f, and NaBH4 -SUG16.9, Figure 4g,h). This is slightly
smaller than the particle size of NaOCH3 (Figure 4a,b), and this reduction in particle size could be the
result of elemental loss in the form of B(OCH3 )3 that is released during the synthesis (reaction 3).
Energies 2019, 12, x FOR PEER REVIEW 9 of 14
3.2.2. Composition of the As-Synthesized Materials
3.2.2. Composition of the As-synthesized materials.
Figure 5 shows the XRD patterns of the as-synthesized NaBH4 obtained from the suspension-gas
Figure
reaction. 5 shows
Different the cubic
to the XRD phase
patterns of the as-synthesized
of NaBH NaBH4 obtained from the suspension-
4 synthesized by solid-gas reaction, the as-synthesized
gas reaction.
NaBH Different to the cubic phase of NaBH4 synthesized by solid-gas reaction, ◦ the as- ◦
4 exhibited peaks of both tetragonal and cubic NaBH4 . The peaks of NaOCH3 at ~11.5 , 30.7 ,
synthesized
◦ NaBH 4 exhibited peaks of both tetragonal and cubic NaBH4. The peaks of NaOCH3 at
and 31.9 in NaBH4 -SUG4.2 disappeared when the molar ratios of NaOCH3 and B2 H6 were 8.5 and
~11.5°,
16.9. This30.7°, and 31.9°
indicated that in
all NaBH 4-SUG4.2 disappeared when the molar ratios of NaOCH3 and B2H6
the NaOCH 3 was fully reacted with B2 H6 to produce NaBH4 .
were 8.5 and 16.9. This indicated that all the NaOCH3 was fully reacted with B2H6 to produce NaBH4.

Figure5.5.XRD
Figure XRDpatterns
patternsof
ofthe
theas-synthesised
as-synthesisedNaBH
NaBH44via
viasuspension-gas
suspension-gasinteraction.
interaction.

The
TheFTIR
FTIRspectra
spectraof ofthe
theNaBH
NaBH44synthesized
synthesizedby bysuspension-gas
suspension-gasreaction
reactionalso
alsoshowed
showedthe thetypical
typical
stretching modes and bending modes of the BH − groups in the range of 2400–2220–1cm−1 and at
stretching modes and bending modes of the BH4 4groups in the range of 2400–2220 cm and at ~1120
-

~1120 cm−1 , respectively


cm–1, respectively (Figure(Figure 6). Additional
6). Additional vibrational vibrational peaks assigned
peaks assigned to C-H wereto C-H were observed
observed between
between 2970–2680 cm −1 (stretching modes), ~1470 cm −1 (bending modes) and 780–710 cm −1 (stretching
2970–2680 cm (stretching modes), ~1470 cm (bending modes) and 780–710 cm (stretching modes)
–1 –1 –1

modes)
[22]. [22].
Compared
Comparedwith withthe
theFTIR
FTIRspectrum
spectrumof ofpristine
pristineoctadecylamine
octadecylamine(Figure(FigureS4),
S4),thethestretching
stretchingmodes
modes
of the primary NH shifted from 3331, 3254 cm −1–1to ~3280, 3135 cm−1 upon stabilization of the NaBH
of the primary NH2 shifted from 3331, 3254 cm to ~3280, 3135 cm upon stabilization of the NaBH4 4
2 –1

nanoparticles.
nanoparticles. In In addition,
addition, thethe multiple
multiple peaks
peaks of of NH
NH22 rocking
rockingandandwagging
waggingmodes modesatat940–750
940–750cm cm-1−1
in
in pristineoctadecylamine
octadecylamine[34–36] [34–36]appeared
appearedas as aa weak
weak peak −1 This indicates that the
pristine peak atat ~880
~880 cmcm–1.. This indicates that the
interaction
interactionof ofNaBH
NaBH44and andoctadecylamine
octadecylaminewas wasthrough
throughits itshead
headgroup
group(NH (NH22).).
FTIR
FTIRanalysis
analysis also proved
also provedthe presence
the presenceof the ofunreacted NaOCHNaOCH
the unreacted 3 in NaBH 3 in4 -SUG4.2. For example,
NaBH4-SUG4.2. For
the peaks at around 2582, 1194, 1082, and 994 cm −1 in NaBH–1-SUG4.2 were assigned to the stretching
example, the peaks at around 2582, 1194, 1082, and 994 cm in NaBH4-SUG4.2 were assigned to the
4
modes of C-O
stretching modes in NaOCH
of C-O in 3 (Figure
NaOCH6)3 (Figure
[21–24,29]. These peaks
6) [21–24,29]. Thesewere not were
peaks observed in NaBHin
not observed 4 -SUG8.5
NaBH4-
and NaBH
SUG8.5 and4 -SUG16.9 by FTIR,by
NaBH4-SUG16.9 and thisand
FTIR, provesthis that
proves at NaOCH 3 and B32 H
that at NaOCH 6 molar
and B2H6 molarratios ratios
higherhigher
than
8.5, NaOCH was fully reacted. However, a significant difference of the suspension-gas
than 8.5, NaOCH3 was fully reacted. However, a significant difference of the suspension-gas reaction
3 reaction as
as compared to the solid-gas reaction carried at 100 °C is the absence of any vibrations related to the
formation NaB12H12 in NaBH4-SUG16.9, despite the excess of B2H6. This indicates that NaB12H12
cannot be formed at room temperature. This may be due to the high energy required for its formation,
e.g., ΔfH0 = −1086.196 kJ mol–1 for its cubic phase, ΔfH0 = −1086.381 kJ mol–1 for its monoclinic phase
[37].
Energies 2019, 12, 4428 9 of 13

compared to the solid-gas reaction carried at 100 ◦ C is the absence of any vibrations related to the
formation NaB12 H12 in NaBH4 -SUG16.9, despite the excess of B2 H6 . This indicates that NaB12 H12
cannot be formed at room temperature. This may be due to the high energy required for its formation,
e.g., ∆f H0 = −1086.196 kJ mol−1 for its cubic phase, ∆f H0 = −1086.381 kJ mol−1 for its monoclinic
Energies[37].
phase 2019, 12, x FOR PEER REVIEW 10 of 14

Figure6.6. FTIR
Figure FTIRspectra
spectraof
ofthe
theas-synthesized
as-synthesizedNaBH
NaBH44 via
via suspension-gas
suspension-gas interaction.
interaction.
3.3. Hydrogen Desorption Properties
3.3. Hydrogen Desorption Properties
The hydrogen desorption properties of the as-synthesized NaBH4 were characterized by TGA/DSC
The hydrogen desorption properties of the as-synthesized NaBH4 were characterized by
and MS. Figures 7 and 8 show the TGA, DSC, and MS of NaBH4 -SG1.8 and NaBH4 -SUG8.5, which
TGA/DSC and MS. Figure 7 and Figure 8 show the TGA, DSC, and MS of NaBH4-SG1.8 and NaBH4-
correspond to the optimum NaOCH3 and B2 H6 molar ratios for both reaction paths according to the
SUG8.5, which correspond to the optimum NaOCH3 and B2H6 molar ratios for both reaction paths
XRD and
Energies FTIR
2019, 12, xanalyses.
FOR PEER REVIEW 11 of 14
according to the XRD and FTIR analyses.
The TGA analysis (Figure 7a) of NaBH4-SG1.8 first showed a mass loss of 3.5% between 280–440
°C. This is correlated to hydrogen released as observed by simultaneous MS analysis (Figure 7b).
Within this temperature range, no obvious peak was observed by DSC. There is also no obvious
evidence of other gases released from NaBH4-SG1.8.
For NaBH4-SUG8.5, the mass loss of 7.1 % from 175–240 °C resulted from the decomposition of
octadecylamine (Figures 8 and S5) [5,38,39], as confirmed by the endothermic peaks observed at 215
°C (Figure 8a) and octadecylamine fragments observed by MS (Figure 8b). It should be noted that in
this case, the mass loss observed by TGA is much larger than the carbon amount determined by the
total carbon analysis, because octadecylamine was removed by filtration before the carbon analysis
[40]. At low temperatures, i.e., 170 °C, the small peak of hydrogen release observed by MS was
assigned to the recombination of the protic hydrogen (NHδ+) in the octadecylamine with the hydridic
hydrogen (BHδ-) from NaBH4. At higher temperatures, i.e., between 240 and 505 °C, the additional
mass loss of 2.1% was assigned to hydrogen release as indicated by MS (Figure 8b).
Figure 7. (a)
Figure 7. (a) TGA/DSC
TGA/DSC under Ar of
under Ar of NaBH
NaBH44-SG1.8;
-SG1.8;(b)
(b)H
H22desorption
desorptionfrom
fromNaBH
NaBH 4 -SG1.8
4-SG1.8
recordedbyby
recorded
mass
mass spectrometry and selected mass corresponding to the release of B22H6 6and B(OCH3)33. 3 .
spectrometry and selected mass corresponding to the release of B H and B(OCH )

The endothermic peaks observed for NaBH4-SG1.8 (Figure 7a) at 484 °C and NaBH4-SUG8.5 at
472 °C (Figure 8a) were assigned to the melting of the as-synthesized NaBH4 nanoparticles. Compare
to the melting point of commercial NaBH4 (506 °C, Figure S6a), this indicates that the reduction in
particle size influences the melting behavior of NaBH4 particles, and this result is in agreement with
previous observations made for almost all free nanoparticles [41–43]. Furthermore, from the
to the melting point of commercial NaBH4 (506 °C, Figure S6a), this indicates that the reduction in
particle size influences the melting behavior of NaBH4 particles, and this result is in agreement with
previous observations made for almost all free nanoparticles [41–43]. Furthermore, from the
TGA/DSC/MS analysis, it can be observed that the temperatures at which the main hydrogen release
start have been significantly reduced to 452 °C for NaBH4-SG1.8 and 384 °C for NaBH4-SUG8.5 in
Energies 2019, 12, 4428
comparison to commercial NaBH4 at 586 °C (Figure S6). This is believed to be attributed to10the of 13

reduction in particle size of the NaBH4 nanoparticles synthesized.

Figure 8. (a) TGA/DSC under Ar of NaBH4 -SUG8.5; (b) H2 desorption from NaBH4 -SUG8.5
Figure 8. (a) TGA/DSC under Ar of NaBH4-SUG8.5; (b) H2 desorption from NaBH4-SUG8.5 recorded
recorded by mass spectrometry and selected mass corresponding to the release of B2 H6 , B(OCH3 )3 ,
by mass spectrometry and selected mass corresponding to the release of B2H6, B(OCH3)3, and
and octadecylamine.
octadecylamine.

The TGA analysis (Figure 7a) of NaBH4 -SG1.8 first showed a mass loss of 3.5% between 280–440 ◦ C.
At temperatures >500 °C, the large mass losses of 17.7% for NaBH4-SG1.8 (Figure 7a) and 24.6%
This is correlated to hydrogen released as observed by simultaneous MS analysis (Figure 7b). Within
for NaBH4-SUG8.5 (Figure 8a) are believed to be the result of additional hydrogen release in
this temperature range, no obvious peak was observed by DSC. There is also no obvious evidence of
conjunction with Na evaporation owing to the low melting point of Na (97.5 °C) and a high vapor
other gases released from NaBH4 -SG1.8.
pressure of 10–3 MPa at 500 °C [44,45].
For NaBH4 -SUG8.5, the mass loss of 7.1 % from 175–240 ◦ C resulted from the decomposition of
octadecylamine
4. Conclusions (Figure 8 and Figure S5) [5,38,39], as confirmed by the endothermic peaks observed at
215 ◦ C (Figure 8a) and octadecylamine fragments observed by MS (Figure 8b). It should be noted that
We
in this have
case, synthesized
the NaBH4 nanoparticles
mass loss observed with larger
by TGA is much different sizes
than theranging
carbon from
amount6 todetermined
260 nm through
by the
the direct
total reaction
carbon of NaOCH
analysis, with B2H6. The particle
because 3octadecylamine size decreased
was removed as the
by filtration amount
before of B2H6analysis
the carbon increased
[40].
during the solid-gas reaction. However, the formation of Na2B12H12 was found to occur when
At low temperatures, i.e., 170 ◦ C, the small peak of hydrogen release observed by MS was assigned to
the recombination of the protic hydrogen (NHδ+ ) in the octadecylamine with the hydridic hydrogen
(BHδ- ) from NaBH4 . At higher temperatures, i.e., between 240 and 505 ◦ C, the additional mass loss of
2.1% was assigned to hydrogen release as indicated by MS (Figure 8b).
The endothermic peaks observed for NaBH4 -SG1.8 (Figure 7a) at 484 ◦ C and NaBH4 -SUG8.5 at

472 C (Figure 8a) were assigned to the melting of the as-synthesized NaBH4 nanoparticles. Compare to
the melting point of commercial NaBH4 (506 ◦ C, Figure S6a), this indicates that the reduction in particle
size influences the melting behavior of NaBH4 particles, and this result is in agreement with previous
observations made for almost all free nanoparticles [41–43]. Furthermore, from the TGA/DSC/MS
analysis, it can be observed that the temperatures at which the main hydrogen release start have been
significantly reduced to 452 ◦ C for NaBH4 -SG1.8 and 384 ◦ C for NaBH4 -SUG8.5 in comparison to
commercial NaBH4 at 586 ◦ C (Figure S6). This is believed to be attributed to the reduction in particle
size of the NaBH4 nanoparticles synthesized.
At temperatures >500 ◦ C, the large mass losses of 17.7% for NaBH4 -SG1.8 (Figure 7a) and 24.6% for
NaBH4 -SUG8.5 (Figure 8a) are believed to be the result of additional hydrogen release in conjunction
with Na evaporation owing to the low melting point of Na (97.5 ◦ C) and a high vapor pressure of
10−3 MPa at 500 ◦ C [44,45].

4. Conclusions
We have synthesized NaBH4 nanoparticles with different sizes ranging from 6 to 260 nm through
the direct reaction of NaOCH3 with B2 H6 . The particle size decreased as the amount of B2 H6 increased
during the solid-gas reaction. However, the formation of Na2 B12 H12 was found to occur when
NaOCH3 was exposed to excessive amounts of B2 H6 during the reaction at 100 ◦ C. In contrast, reacting
NaOCH3 nanoparticles in suspension with B2 H6 at room temperature did not lead to the formation of
Na2 B12 H12 , and the purity of the resulting NaBH4 was ~98.0%. The start of hydrogen release from our
Energies 2019, 12, 4428 11 of 13

NaBH4 nanoparticles is much lower than that of commercial NaBH4 (by at least 100 ◦ C) and this was
attributed to the nanosize nature of the materials synthesized. These results not only demonstrate the
possibility to synthesize NaBH4 nanoparticle via a direct synthesis route with monodispersed particles,
in particular through the reaction of NaOCH3 suspended in solution, but also the possibility to reduce
the temperature at which hydrogen is released. Additional modification through a core-shell approach,
for example, to confine NaBH4 and its elements upon hydrogen release, could lead to better hydrogen
properties in particular in terms of hydrogen reversibility by avoiding elemental segregation. This will
be at the core of future work.

Supplementary Materials: The following are available online at http://www.mdpi.com/1996-1073/12/23/4428/s1.


Author Contributions: All Authors contributed equally. Conceptualization, T.W. and K.-F.A.-Z.; Methodology,
T.W. and K.-F.A.-Z.; Validation, T.W.; Formal Analysis, T.W.; Investigation, T.W. and K.-F.A.-Z.; Resources,
K.-F.A.-Z.; Data Curation, T.W.; Writing-Original Draft Preparation, T.W.; Writing-Review & Editing, K.-F.A.-Z.;
Supervision, K.-F.A.-Z.; Project Administration, K.-F.A.-Z.; Funding Acquisition, K.-F.A.-Z.
Funding: This research was funded by UNSW Internal Research Grant program and Office of Naval Research
(Award No: ONRG - NICOP - N62909-16-1-2155).
Acknowledgments: We appreciate the use of instruments in the Mark Wainwright Analytical Centre at UNSW.
The China Scholarship Council is gratefully acknowledged for the Scholarship granted to Ting Wang.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Christian, M.L.; Aguey-Zinsou, K.-F. Core–Shell strategy leading to high reversible hydrogen storage capacity
for NaBH4 . ACS Nano 2012, 6, 7739–7751. [CrossRef] [PubMed]
2. Lai, Q.; Sun, Y.; Wang, T.; Modi, P.; Cazorla, C.; Demirci, U.B.; Ares Fernandez, J.R.; Leardini, F.;
Aguey-Zinsou, K.F. How to design hydrogen storage materials? Fundamentals, synthesis, and storage tanks.
Adv. Sust. Sys. 2019, 3, 1900043. [CrossRef]
3. Lai, Q.; Wang, T.; Sun, Y.; Aguey-Zinsou, K.F. Rational design of nanosized light elements for hydrogen
storage: Classes, synthesis, characterization, and properties. Adv. Mater. Technol. 2018, 3, 1700298. [CrossRef]
4. De Jongh, P.E.; Adelhelm, P. Nanosizing and nanoconfinement: New strategies towards meeting hydrogen
storage goals. ChemSusChem 2010, 3, 1332–1348. [CrossRef]
5. Wang, T.; Aguey-Zinsou, K.F. Controlling the growth of LiBH4 nanoparticles for hydrogen storage. Energy
Technol. 2019, 7, 1801159. [CrossRef]
6. Lai, Q.; Yang, Y.; Aguey-Zinsou, K.-F. Nanoconfinement of borohydrides in hollow carbon spheres: Melt
infiltration versus solvent impregnation for enhanced hydrogen storage. Int. J. Hydrogen Energy 2019, 44,
23225–23238. [CrossRef]
7. Lai, Q.; Christian, M.; Aguey-Zinsou, K.-F. Nanoconfinement of borohydrides in CuS hollow nanospheres:
A new strategy compared to carbon nanotubes. Int. J. Hydrogen Energy 2014, 39, 9339–9349. [CrossRef]
8. Clémençon, D.; Davoisne, C.; Chotard, J.-N.; Janot, R. Enhancement of the hydrogen release of Mg(BH4 )2 by
concomitant effects of nano-confinement and catalysis. Int. J. Hydrogen Energy 2019, 44, 4253–4262. [CrossRef]
9. Christian, M.L. Core-Shell Borohydrides for Reversible Hydrogen Storage. Ph.D. Thesis, University of New
South Wales, Sydney, Australia, 2013.
10. Lai, Q.W.; Paskevicius, M.; Sheppard, D.A.; Buckley, C.E.; Thornton, A.W.; Hill, M.R.; Gu, Q.F.; Mao, J.F.;
Huang, Z.G.; Liu, H.K.; et al. Hydrogen storage materials for mobile and stationary applications: Current
state of the art. ChemSusChem 2015, 8, 2789–2825. [CrossRef]
11. Lai, Q.; Milanese, C.; Aguey-Zinsou, K.-F. Stabilization of nanosized borohydrides for hydrogen storage:
Suppressing the melting with TiCl3 doping. ACS Appl. Energy Mater. 2018, 1, 421–430. [CrossRef]
12. Mangold, H.; Ettlinger, M.; Kerner, D.; Kleinschmit, P. Boron Oxide-Silicon Dioxide Mixed Oxide. U.S. Patent
US6242373B1, 5 June 2001.
13. Murphy, R.J.; Dickinson, D.J.; Turner, P. Treatment of Wood and Wood-Based Materials. U.S. patent 5330847,
19 July 1994.
14. Lewis, R.J., Sr.; Lewis, R.A.; Hawley, G.G. Hawley’s Condensed Chemical Dictionary, 16th ed.; John Wiley &
Sons: Hoboken, NJ, USA, 2016.
Energies 2019, 12, 4428 12 of 13

15. Richardson, M.L.; Gangolli, S. The Dictionary of Substances and Their Effects, 2nd ed.; Royal Society of Chemistry:
Cambridge, UK, 1992; Volume 2.
16. Kanth, J.V.; Brown, H.C. Improved procedures for the generation of diborane from sodium borohydride and
boron trifluoride. Inorg. Chem. 2000, 39, 1795–1802. [CrossRef] [PubMed]
17. Wang, T.; Zhang, X.; Zhang, F.; Wang, W.; Liang, Y.; Tang, Y. Uniform ultrasmall manganese monoxide
nanoparticle/carbon nanocomposite as a high-performance anode for lithium storage. Electrochim. Acta 2016,
196, 634–641. [CrossRef]
18. Sun, B.; Chen, Z.; Kim, H.-S.; Ahn, H.; Wang, G. MnO/C core–shell nanorods as high capacity anode materials
for lithium-ion batteries. J. Power Sour. 2011, 196, 3346–3349. [CrossRef]
19. Huang, J.; Ouyang, L.; Gu, Q.; Yu, X.; Zhu, M. Metal-Borohydride-Modified Zr(BH4 )4 ·8NH3 : Low-Temperature
dehydrogenation yielding highly pure hydrogen. Chem. Eur. J. 2015, 21, 14931–14936. [CrossRef]
20. Zheng, J.; Xiao, X.; Zhang, L.; Li, S.; Ge, H.; Chen, L. Facile synthesis of bowl-like 3D Mg (BH4 )2 –NaBH4 –
fluorographene composite with unexpected superior dehydrogenation performances. J. Mater. Chem. A
2017, 5, 9723–9732. [CrossRef]
21. Xue, Z. Raman Spectroscopy of Carboxylic Acid and Water Aggregates; Logos Verlag: Berlin, Germany; GmbH:
Berlin, Germany, 2011.
22. Ramesh, S.; Leen, K.H.; Kumutha, K.; Arof, A.K. FTIR studies of PVC/PMMA blend based polymer
electrolytes. Spectrochim. Acta Part A 2007, 66, 1237–1242. [CrossRef]
23. Dai, H.-L. Spectroscopy, Structure, and Energy Transfer of Transient Radicals in Combustion; University of
Pennsylvania: Philadelphia, PA, USA, 2007.
24. Gómez-Ordóñez, E.; Rupérez, P. FTIR-ATR spectroscopy as a tool for polysaccharide identification in edible
brown and red seaweeds. Food Hydrocolloids 2011, 25, 1514–1520. [CrossRef]
25. Cornelius, C.J.; Marand, E. Hybrid inorganic–organic materials based on a 6FDA–6FpDA–DABA polyimide
and silica: Physical characterization studies. Polymer 2002, 43, 2385–2400. [CrossRef]
26. Gautam, C.; Yadav, A.K.; Mishra, V.K.; Vikram, K. Synthesis, IR and Raman spectroscopic studies of (Ba, Sr)
TiO3 borosilicate glasses with addition of La2 O3 . Open J. Inorg. Non Met. Mater. 2012, 2, 47–54.
27. Lopes, J.d.O.; Garcia, R.A.; Souza, N.D.D. Infrared spectroscopy of the surface of thermally-modified teak
juvenile wood. Maderas. Ciencia Tecnología 2018, 20, 737–746. [CrossRef]
28. Mahmood, Z.; Azam, M.; Mushtaq, A.; Kausar, R.; Kausar, S.; Gilani, S.R. Comparative vapour phase
FTIR spectra and vibrational assignment of manganese pentacarbonyls derivatives of the type XMn (CO)5 :
(where X = Br, Cl, I, H, D, CH3 , CD3 , CF3 ). Spectrochim. Acta Part A 2006, 65, 445–452. [CrossRef] [PubMed]
29. Liu, K.-Z.; Jackson, M.; Sowa, M.G.; Ju, H.; Dixon, I.M.; Mantsch, H.H. Modification of the extracellular
matrix following myocardial infarction monitored by FTIR spectroscopy. BBA Mol. Basis Dis. 1996, 1315,
73–77. [CrossRef]
30. Genady, A.R. Labeled undecahydro-closo-dodecaborates based on azo dyes for boron neutron capture
therapy: Synthesis, characterization, and visualization in cells. Acta Chim. Slov. 2012, 59, 1.
31. Nelson, M.A.; Kodama, G. Deuterated sodium octahydrotriborate (1-). Inorg. Chem. 1979, 18, 3276–3278.
[CrossRef]
32. Gaines, D.F.; Schaeffer, R.; Tebbe, F. Convenient preparations of solutions containing the triborohydride ion.
Inorg. Chem. 1963, 2, 526–528. [CrossRef]
33. Chen, W.; Wu, G.; He, T.; Li, Z.; Guo, Z.; Liu, H.; Huang, Z.; Chen, P. An improved synthesis of unsolvated
NaB3 H8 and its application in preparing Na2 B12 H12 . Int. J. Hydrogen Energy 2016, 41, 15471–15476. [CrossRef]
34. Mirghani, M.; Man, Y.C.; Jinap, S.; Baharin, B.; Bakar, J. FTIR spectroscopic determination of soap in refined
vegetable oils. J. Am. Oil Chem. Soc. 2002, 79, 111–116. [CrossRef]
35. Ramalingam, M.; Sundaraganesan, N.; Saleem, H.; Swaminathan, J. Experimental (FTIR and FT-raman) and
ab initio and DFT study of vibrational frequencies of 5-amino-2-nitrobenzoic acid. Spectrochim. Acta Part A
2008, 71, 23–30. [CrossRef]
36. Gibson, N.; Shenderova, O.; Luo, T.; Moseenkov, S.; Bondar, V.; Puzyr, A.; Purtov, K.; Fitzgerald, Z.; Brenner, D.
Colloidal stability of modified nanodiamond particles. Diam. Relat. Mater. 2009, 18, 620–626. [CrossRef]
37. Caputo, R.; Garroni, S.; Olid, D.; Teixidor, F.; Suriñach, S.; Baro, M.D. Can Na2 [B12 H12 ] be a decomposition
product of NaBH4 ? Phys. Chem. Chem. Phys. 2010, 12, 15093–15100. [CrossRef]
Energies 2019, 12, 4428 13 of 13

38. Purkayastha, D.; Sarma, B.; Bhattacharjee, C. Surfactant-assisted low-temperature synthesis of monodispersed
phase pure cubic CoO solid nanoparallelepipeds via thermal decomposition of cobalt (II) acetylacetonate.
Mater Lett. 2013, 107, 71–74. [CrossRef]
39. Züttel, A.; Rentsch, S.; Fischer, P.; Wenger, P.; Sudan, P.; Mauron, P.; Emmenegger, C. Hydrogen storage
properties of LiBH4 . J. Alloys Compd. 2003, 356, 515–520. [CrossRef]
40. Fan, B.; Wei, G.; Zhang, Z.; Qiao, N. Characterization of a supramolecular complex based on octadecylamine
and β-cyclodextrin and its corrosion inhibition properties in condensate water. Corros. Sci. 2014, 83, 75–85.
[CrossRef]
41. Jiang, H.; Moon, K.-S.; Dong, H.; Hua, F.; Wong, C. Size-Dependent melting properties of tin nanoparticles.
Chem. Phys. Lett. 2006, 429, 492–496. [CrossRef]
42. Nanda, K. Size-Dependent melting of nanoparticles: Hundred years of thermodynamic model. Pramana
2009, 72, 617–628. [CrossRef]
43. Olson, E.; Efremov, M.Y.; Zhang, M.; Zhang, Z.; Allen, L. Size-Dependent melting of Bi nanoparticles. J. Appl.
Phys. 2005, 97, 034304. [CrossRef]
44. Honig, R.E. Vapor Pressure Data for the More Common Elements; David Sarnoff Research Center: Princeton, NJ,
USA, 1957.
45. Ngene, P.; van den Berg, R.; Verkuijlen, M.H.; de Jong, K.P.; de Jongh, P.E. Reversibility of the hydrogen
desorption from NaBH4 by confinement in nanoporous carbon. Energy Environ. Sci. 2011, 4, 4108–4115.
[CrossRef]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like