You are on page 1of 9

Highlighting work from the Department of Applied Chemistry at As featured in:

Kyushu University, Fukuoka, Japan.

Effects of ball-milling treatment on physicochemical properties


and solid base activity of hexagonal boron nitrides

Ball-milling is an established method for obtaining fine particles


with high surface area from bulk materials. This top-down
technique not only increases the surface area of hexagonal boron
nitride (h-BN) by comminution, but it also imparts the material
with chemical functionalities and altered morphology. The
ball-milled h-BN showed catalytic activity for the nitroaldol
reaction and the h-BN milled at 400 rpm was the most
active catalyst.
See Atsushi Takagaki et al.,
Catal. Sci. Technol., 2019, 9, 302.

rsc.li/catalysis
Registered charity number: 207890
Catalysis
Science &
Technology
View Article Online
PAPER View Journal | View Issue

Effects of ball-milling treatment on


Published on 08 November 2018. Downloaded by Cornell University Library on 9/2/2020 11:01:30 AM.

Cite this: Catal. Sci. Technol., 2019,


physicochemical properties and solid base activity
9, 302 of hexagonal boron nitrides
Shoichiro Namba,a Atsushi Takagaki, *bc Keiko Jimura, d
Shigenobu Hayashi, d

Ryuji Kikuchi a and S. Ted Oyama a

Hexagonal boron nitride (h-BN) was ball-milled at various rotation speeds (150–600 rpm) using a planetary
ball-mill. Ball-milling disrupted the layered structure of the h-BN, resulting in significant increases of sur-
face area. Ball-milling at 400 rpm gave the highest surface area of 412 m2 g−1 while higher rotation speeds
decreased the surface areas due to agglomeration. Moreover, ball-milling resulted in the emergence of
amino- and hydroxyl groups on the surface which were observed by Fourier transform infrared spectro-
scopy, and partial oxidation of the surface boron by the formation of B–OH groups was confirmed by
X-ray photoelectron spectroscopy. The appearance of trigonal B–O and tetrahedral B–O was observed by
boron-11 magic-angle spinning nuclear magnetic resonance spectroscopy. The number of base sites was
Received 10th May 2018, increased with the increase of rotation speeds of milling, corresponding to the formation of amino groups.
Accepted 8th November 2018
The ball-milled h-BN showed catalytic activity for the nitroaldol reaction between nitromethane and benz-
aldehyde in which the h-BN milled at 400 rpm exhibited the highest reaction rate and turnover frequency.
DOI: 10.1039/c8cy00940f
In addition, the ball-milled h-BN could convert glucose with the formation of fructose at 40 °C whereas
rsc.li/catalysis pristine h-BN showed no activity. The base sites were mainly responsible for the catalytic activity.

Introduction dehydrogenation (ODH) of light alkanes.17,18 The h-BN and


its nanotubes functioned as metal-free catalysts for ODH of
Ball-milling is an established method for obtaining fine parti- propane and afforded very high selectivity to propylene and
cles with high surface area from bulk materials. This top-down ethylene.17 A hydroxylated BN (ref. 19) and ball-milled h-BN
technique has been used for the synthesis of a variety of nano- (ref. 20) also catalyzed ODH of ethane with high selectivity to
materials such as zeolites,1,2 metal oxides,3,4 carbon materials,5 ethylene. Furthermore, carbon doped BN nanosheets
and metal–organic frameworks.6 Such ball-milled materials exhibited high activity for ODH of ethylbenzene to styrene.21
have broad applications as heterogeneous catalysts,7,8 electro- For ODH reactions, it is considered that the oxidized boron
catalysts,9 electrode materials,10 electronic devices,11 and hy- species formed on the edge sites of BN play an important role
drogen storage materials,12 and in drug delivery.13 in the activity.17–21
Hexagonal boron nitride (h-BN) has a similar crystal struc- Ball-milling not only increases the surface area of h-BN by
ture as graphite and is sometimes called “white graphite.” comminution, but also imparts it with chemical functionali-
The h-BN is attracting much attention because of its fascinat- ties and altered morphology. Ball-milling under a hydrogen
ing properties which are applicable for ultraviolet laser de- atmosphere afforded hydrogenated boron nitride,22 and in
vices,14 dielectric layers for graphene devices,15 and protective the presence of ammonia borane23 or aqueous NaOH solu-
coatings.16 The h-BN is also found to be active for oxidative tion24 produced exfoliated boron nitride nanosheets.
Recently, we have reported that the ball-milled h-BN func-
a
tioned as an efficient solid acid–base bifunctional catalyst
Department of Chemical System Engineering, School of Engineering, The
University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo, 113-8656, Japan
owing to the increase of surface area and the emergence of
b
Department of Applied Chemistry, Faculty of Engineering, Kyushu University, hydroxyl- and amino groups on the surface.25 The ball-milled
744 Motooka, Nishi-ku, Fukuoka, 819-0395, Japan. h-BN exhibited high activity for base-catalyzed reactions such
E-mail: atakagak@cstf.kyushu-u.ac.jp as the nitroaldol reaction and Knoevenagel condensation.
c
International Institute for Carbon-Neutral Energy Research (I2CNER), Kyushu
Solid acid–base bifunctionality is well documented for or-
University, Motooka, Nishi-ku, Fukuoka, 819-0395, Japan
d
Research Institute for Material and Chemical Measurement, National Institute of
ganic amine-tethered silicas,26–28 but rare for nitride-based
Advanced Industrial Science and Technology (AIST), Central 5, 1-1-1 Higashi, materials, therefore further study of the material is necessary.
Tsukuba, Ibaraki 305-8656, Japan The objective of this report is to investigate the effect of ball-

302 | Catal. Sci. Technol., 2019, 9, 302–309 This journal is © The Royal Society of Chemistry 2019
View Article Online

Catalysis Science & Technology Paper

milling on the physicochemical properties of h-BN and its ac- etherate (CF3·(C2H5)2O), by setting the signal of NaBH4 spin-
tivity as a solid base. Two types of base-catalyzed reactions ning at 8 kHz to −42.00 ppm.
were carried out. One is the nitroaldol reaction of nitrometh- The oxidation state of elements and their compositions
ane and benzaldehyde because the reaction is one of the im- over the surface were analyzed by X-ray photoelectron spectro-
portant C–C-bond formation reactions and is appropriate to scopy (XPS, JPS-9200, JEOL). The samples were exposed to the
evaluate the solid acid–base bifunctionality as previously atmosphere for these measurements, but the oxidation state
reported.25 The coexistence of acidic OH groups facilitates of the components is not expected to change because the
the activation of the aldehyde.25,26,28 The other is the trans- functional groups B–OH and NH2 are unreactive at room tem-
Published on 08 November 2018. Downloaded by Cornell University Library on 9/2/2020 11:01:30 AM.

formation of glucose because it can be converted to perature. Some adsorption of H2O is likely, but a considerable
fructose via isomerization using a base or Lewis acid catalyst. amount would be desorbed by sample evacuation.
Conventional solid bases including hydrotalcite and MgO The base catalytic activity of ball-milled h-BN was deter-
were used for comparison because these solid bases can cata- mined through the nitroaldol reaction. The reaction was
lyze the nitroaldol reaction29 and the isomerization.30,31 conducted using an h-BN catalyst (50 mg) in 2 mL toluene so-
lution with benzaldehyde (0.5 mmol), nitromethane (1.25
mmol) and n-decane (0.1 mmol) as an internal standard. The
Experimental reaction was performed at 100 °C. It took 6 minutes to warm
the reaction mixture to this temperature. Aliquots were taken
The ball-milling method was used to exfoliate and disrupt using a syringe and analyzed by gas chromatography (GC-
layered pristine h-BN. A quantity of 0.8 g of h-BN (Wako) was FID; GC-2014, column DB-1MS, Shimadzu).
ball-milled at several rotation speeds (150, 300, 400, 500, 600 In addition, transformation of glucose was carried out
rpm) using a planetary ball-mill (Pulverisette 7, Fritsch, zirco- using 20 mg of the catalyst in 5 mL methanol with glucose
nia vessel with six zirconia balls (diameter 10 mm)). Ball- (0.17 mmol, 30 mg). The reaction was performed at 40 °C for
milling was conducted for 24 cycles in which each cycle was 15 min. The reactant and products were analyzed by high
carried out for 30 min with 5 min intervals. The rotational di- performance liquid chromatography (HPLC; LC-2000 plus,
rection was reversed between cycles. The samples prepared JACO) equipped with a differential refractive index detector
were denoted as BN 150–600 rpm, respectively. (RI-2031 plus, JASCO) using an Asahipak NH2P-50 column
The crystal structure of the h-BN was investigated by X-ray (flow rate: 1.0 mL min−1, eluent: CH3CN : H2O = 4 : 1). Reten-
diffraction (XRD, RINT-2700, Rigaku) with Cu Kα radiation (λ tion times of glucose and fructose were identified as 12.1
= 0.1548 nm) at a voltage of 40 kV and a current of 100 mA. min and 9.5 min.
Scans were obtained at a speed of 2° min−1 with a step width
of 0.02° for 2θ values of 10° to 60°. Results and discussion
The surface area was evaluated by the Brunauer–Emmett–
Teller (BET) method using nitrogen adsorption (BELLSORP Fig. 1 shows the XRD patterns of h-BN samples before and af-
mini-II, Microtrac-BEL). The sample was pretreated by evacu- ter ball-milling at several rotation speeds. Pristine h-BN
ation at 200 °C for 1 h before the measurement. The pore showed sharp peaks corresponding to the (100), (002), (101),
size distribution of the sample was obtained from analysis of
the adsorption branches of the isotherms using the Barrett,
Joyner, and Halenda (BJH) method.
The surface morphology of the h-BN was observed by
scanning electron microscopy (SEM, S-4700IJHitachi), and
JSM-7900F (JEOL)). The surface functional groups of the h-BN
were characterized by Fourier transform infrared spectro-
scopy (FTIR, FT/IR-6100, JASCO) measurements. The sample
was pressed into pellets with KBr for the measurements.
Boron-11 magic-angle spinning nuclear magnetic resonance
(11B MAS NMR) measurement was also carried out. The spec-
tra were recorded with a Bruker Avance III HD 600WB
spectrometer operating at a frequency of 192.63 MHz. A
Bruker MAS probe head was used with a zirconia rotor with
an outer diameter of 4 mm. The measurement was
performed at room temperature with a spinning rate of 12
kHz. The 11B spectrum was measured with a single pulse se-
quence combined with 1H high power decoupling during sig-
nal acquisition. The flip angle of the pulse and the recycle de-
lay were π/6 for solution and 30 s, respectively. The shift was
expressed with respect to neat boron trifluoride diethyl Fig. 1 X-ray diffraction patterns of pristine and ball-milled h-BN.

This journal is © The Royal Society of Chemistry 2019 Catal. Sci. Technol., 2019, 9, 302–309 | 303
View Article Online

Paper Catalysis Science & Technology

(102) and (004) reflections, indicating that the h-BN before dians, and θ is the diffraction angle of the (002) plane in de-
ball-milling had high crystallinity. No apparent change was grees (13.3). The mean crystallite sizes were 45 nm for pris-
found for BN 150 rpm, suggesting that the slow rotation tine h-BN, 39 nm for BN 150 rpm, 10 nm for BN 300 rpm, 15
speed did not engender much impact energy. In contrast, sig- nm for BN 400 rpm, 29 nm for BN 500 rpm, and 31 nm for
nificant weakening of all diffraction peaks was observed for BN 600 rpm. The broad shoulders at ∼25 and ∼42 degrees
BN 300 rpm and BN 400 rpm, indicating that ball milling at were observed for the ball-milled samples, which are likely at-
such rotation speeds disrupted both the layer and in-plane tributed to the amorphous phase.
structures of h-BN. At rotation speeds higher than 400 rpm, Fig. 2 shows the surface areas and pore size distributions
Published on 08 November 2018. Downloaded by Cornell University Library on 9/2/2020 11:01:30 AM.

the peaks corresponding to the in-plane modes ((100) and of pristine and ball-milled h-BN samples. The BET surface
(102)) continuously weakened whereas those of the layer area of pristine h-BN was quite low, 3 m2 g−1. The surface area
structures ((002) and (004)) strengthened. The increase of the was increased to 26 m2 g−1 for BN 150 rpm. A substantial in-
peaks corresponding to the layered structures is likely due to crease of the surface areas were found for BN 300 and BN 400
agglomeration of the fine particles. Agglomeration is a well- rpm, which were 384 and 412 m2 g−1, respectively. The surface
known phenomenon for long time milling.32,33 No peaks for areas were decreased to 211 m2 g−1 for BN 500 rpm and 93
other materials such as B2O3 and BIJOH)3 were found, which m2 g−1 for BN 600 rpm. These variations were in good agree-
is important because these boron compounds would exhibit ment with the results of XRD. The increase of surface areas
activity in acid–base reactions. To verify the crystallinity of was attributed to the disruption of the structure of h-BN, and
the samples, the peak width at half maximum intensity of the decrease of the surface areas by ball-milling at high rota-
the (002) plane was calculated. The values of FWHM (degree) tion speeds was likely due to agglomeration of the small parti-
were 0.18 for pristine h-BN, 0.21 for BN 150 rpm, 0.8 for BN cles. Both micropores and mesopores were monotonically in-
300 rpm, 0.54 for BN 400 rpm, 0.28 for BN 500 rpm, and 0.27 creased with the increase of the BET surface areas.
for BN 600 rpm. The mean crystallite sizes of the samples
were calculated using the Scherrer equation, D = Kλ/β cos θ
where K is the Scherrer constant (0.9), λ is the wavelength of
X-ray radiation (Cu Kα, 0.15418 nm), β is the FWHM in ra-

Fig. 3 SEM images of pristine h-BN (a and b), BN 400 rpm (c and d)
and BN 600 rpm (e and f).

Fig. 2 (a) BET surface areas and (b) pore size distributions of pristine
and ball-milled h-BN. Fig. 4 FTIR spectra of pristine and ball-milled h-BN.

304 | Catal. Sci. Technol., 2019, 9, 302–309 This journal is © The Royal Society of Chemistry 2019
View Article Online

Catalysis Science & Technology Paper

Fig. 3 shows SEM images of pristine h-BN, BN 400 rpm, culated to be 55% for trigonal B–N, 30% for trigonal B–O,
and BN 600 rpm. Pristine h-BN consisted of plate-like parti- and 15% for tetrahedral B–O.
cles with 1–10 μm in size. In contrast, BN 400 rpm was com- Fig. 6 shows the XPS spectra of h-BN samples. Fig. 6(a)
posed of small particles with less than one μm in size. BN displays the B1s spectra. Pristine h-BN had a single peak cen-
600 rpm had the same morphology as BN 400 rpm. tered at 191.0 eV.43,44 The sample BN 150 rpm showed almost
Fig. 4 shows FTIR spectra for h-BN before and after ball-
milling. A strong absorption centered at 1376 cm−1 and a
weak and sharp band at 819 cm−1 were observed for pristine
Published on 08 November 2018. Downloaded by Cornell University Library on 9/2/2020 11:01:30 AM.

h-BN. These peaks correspond to B–N stretching and B–N


bending vibrations, respectively.34,35 No differences in the
spectra of pristine h-BN and BN 150 rpm were observed. In
contrast, apparent changes were found for the samples ball-
milled at 300 rpm and higher rotation. The bands attributed
to B–N vibrations, ∼1376 cm−1 and ∼819 cm−1 decreased and
broadened with the increase of ball milling rotation speed.
Simultaneously, the bands at 1079 cm−1 and 925 cm−1
appeared for samples BN 300–600 rpm which can be ascribed
to B–O vibrations.36,37 Also, the bands around 3400 cm−1 and
3150 cm−1 were observed which can be assigned to O–H and
N–H stretching vibrations, respectively.25,36,38 These bands
became intense with the increase of milling rotation speed.
These indicate that hydroxyl- and amino groups were formed
by ball-milling via cleavage of B–N bonds.39
Fig. 5 shows the 11B MAS NMR spectra of ball-milled h-BN
(BN 400 rpm). Three signals are observed; the first one cen-
tered at about 24 ppm with an isotropic chemical shift (δiso)
of 29.6 ppm, the second one ranging from 18 to 8 ppm with
an δiso of 18.8 ppm, and the third narrow one with an δiso of
0.8 ppm. The signal with the shift of 29.6 ppm is attributed
to trigonal boron connected with nitrogen, B–N,40 which is
dominant in h-BN. The second one with an δiso of 18.8 ppm
is ascribed to trigonal boron bonded with oxygen, B–O.41 The
appearance of this signal was in good agreement with the re-
sults of FTIR. The signal at 0.8 ppm is attributed to tetrahe-
dral boron, which might be due to the formation of B–OIJH)
on the basal plane of h-BN or amorphous boron oxide. The
appearance of the tetrahedral boron was also reported for the
ball-milled h-BN with LiOH.42 The distribution of B was cal-

Fig. 6 XPS spectra of pristine and ball-milled h-BN. (a) B1s, (b) N1s
11
Fig. 5 B MAS NMR spectra of ball-milled h-BN (BN 400 rpm). and (c) O1s.

This journal is © The Royal Society of Chemistry 2019 Catal. Sci. Technol., 2019, 9, 302–309 | 305
View Article Online

Paper Catalysis Science & Technology

Table 1 Surface compositions of pristine and ball-milled h-BN calculated


from XPS

Sample N/B O/B BB–O/(BB–N + BB–O) O/BB–O


Pristine h-BN 0.97 0 0 —
BN 150 rpm 0.92 0.01 0.0 —
BN 300 rpm 0.84 0.13 0.10 1.26
BN 400 rpm 0.78 0.34 0.26 1.31
BN 500 rpm 0.61 0.55 0.40 1.38
Published on 08 November 2018. Downloaded by Cornell University Library on 9/2/2020 11:01:30 AM.

BN 600 rpm 0.62 0.49 0.30 1.62

the same spectrum as pristine h-BN. At rotation speeds of


300 rpm and higher, an additional peak centered at 193.1 eV Fig. 7 Reaction rate for nitroaldol reaction using ball-milled h-BN.
appeared which is ascribed to oxidized boron, B3+.43,45–47 The
oxidized boron was related to the formation of B–OH groups,
and the relative composition of oxidized boron to h-BN struc- 150 rpm. For the other samples, the number of base sites
tural boron BB–O/(BB–N + BB–O) increased with the increase of was 0.31 mmol g−1 in BN 300 rpm, 0.52 mmol g−1 in BN 400
rotation speed. The ratios were 0 for BN 150 rpm, 0.10 for BN rpm, 0.83 mmol g−1 in BN 500 rpm and 0.74 mmol g−1 in BN
300 rpm, 0.26 for BN 400 rpm, 0.40 for BN 500 rpm, and 0.30 600 rpm. The increase of the number of base sites is in good
for BN 600 rpm. The samples ball-milled at high rotation agreement with the FTIR results in which the intensity of the
speeds had a layered structure with low surface areas as band corresponding to the N–H bond was increased. Table 2
shown in Fig. 1 and 2, which was likely due to agglomeration. summarizes the reaction rates, number of base sites, and
It can be considered that the inner part of the agglomerated turnover frequencies (TOFs) as well as the surface areas. The
particles was hard to be oxidized, resulting in the decrease of TOF increased from 2.2 h−1 for BN 300 rpm to 4.4 h−1 for BN
the ratio of oxidized boron to structural boron (BB–O/(BB–N + 400 rpm and then decreased to 2.6 h−1 for BN 500 rpm and
BB–O)) for BN 600 rpm. Fig. 6(b) shows the N1 s XPS spectra. 1.6 h−1 for BN 600 rpm. BN 400 rpm showed the highest reac-
A single peak centered at 398.8 eV was observed for all sam- tion rate and the highest TOF among the samples tested.
ples. The peak weakened and broadened with the increase of Given that TOFs for structure-sensitive reactions can vary by
rotation speed. The broad spectra of the ball-milled h-BN 106 and more,48 such differences are considered to be small
samples at high rotation speeds may involve amino groups in and indicated that the base sites are catalyzing the reaction.
addition to structural nitrogen, but difficult to be separated. It was found that the reaction rates were not directly corre-
Fig. 6(c) shows the O1s XPS spectra. A single peak centered at lated with surface areas (see Fig. 2 and 7). The surface area of
533.5 eV appeared for BN 300 rpm. The peak strengthened BN 300 rpm was close to that of BN 400 rpm, but the reaction
with the increase of rotation speed. rate over BN 300 rpm was lower than that over BN 400 rpm.
Table 1 lists the results of N/B, O/B atomic ratios on the Also, the reaction rates over BN 400 rpm and BN 500 rpm
h-BN surface calculated from XPS. The N/B ratio for pristine were similar, but the surface area of BN 500 rpm was half of
h-BN was close to one. As the rotation speed became high, that of BN 400 rpm. It is concluded that the number of base
the O/B ratio was increased from 0 to 0.55 whereas the N/B sites does not track the surface areas because of differences
ratio was decreased. The ratio of oxygen to oxidized boron, O/ in exposed surfaces. The base sites are primarily responsible
BB–O, was 1.26 for BN 300 rpm, 1.31 for BN 400 rpm, 1.38 for for the catalytic activity. In addition, the dependence of the
BN 500 rpm, and 1.62 for BN 600 rpm. The O/BB–O ratios were surface areas on the catalytic activity was estimated. The areal
between one and two, suggesting that the average structure is reaction rates should be a constant if the active sites were
a mixture of single and germinal hydroxyl groups, B–OH and uniformly formed on the basal plane of h-BN. However, the
BIJOH)2. The increase of the O/BB–O ratios indicates the incre- areal reaction rates were increased with the increase of the
ment on the proportion of germinal hydroxyl groups via suc- rotation speed. The areal reaction rate was 0 μmol m−2 h−1
cessive cleavage of the B–N bond. for BN 150 rpm, 1.8 μmol m−2 h−1 for BN 300 rpm, 5.6 μmol
Fig. 7 shows the results of the nitroaldol reaction of nitro- m−2 h−1 for BN 400 rpm, 10 μmol m−2 h−1 for BN 500 rpm,
methane and benzaldehyde using ball-milled h-BN. No activ- and 13 μmol m−2 h−1 for BN 600 rpm. This indicates that
ity was found for BN 150 rpm. A small degree of activity for most of the active sites were formed on the edge sites of ball-
the nitroaldol reaction was found with BN 300 rpm for which milled h-BN. The catalytic activities of ball-milled h-BN cata-
the reaction rate was 0.69 mmol g−1 h−1. The reaction rate lysts were compared with those of calcined MgO, calcined
was much increased for BN 400 rpm to 2.3 mmol g−1 h−1. A hydrotalcite (HT), and n-butylamine. The reaction rate was
similar activity was found for BN 500 rpm, and then the activ- 0.3 mmol g−1 h−1 for MgO, 1.8 mmol g−1 h−1 for HT, and 4.9
ity decreased for BN 600 rpm. The number of base sites in mmol g−1 h−1 for n-butylamine. The TOFs of MgO, HT, and
ball-milled samples was determined by titration using n-butylamine were much lower than that of BN 400 rpm.
benzoic acid. No or negligible base sites were found in BN Table 3 shows the results of the nitroaldol reaction using

306 | Catal. Sci. Technol., 2019, 9, 302–309 This journal is © The Royal Society of Chemistry 2019
View Article Online

Catalysis Science & Technology Paper

Table 2 Results of nitroaldol reaction of nitromethane and benzaldehyde using ball-milled h-BNa

Sample SBET/m2 g−1 Base sites/mmol g−1 Reaction rate/mmol g−1 hour−1 TOF/h−1
BN 150 rpm 26 0 0.0 0
BN 300 rpm 384 0.31 0.69 2.2
BN 400 rpm 412 0.52 2.3 4.4
BN 500 rpm 211 0.83 2.2 2.6
BN 600 rpm 93 0.74 1.2 1.6
Calcined MgOb 20 0.31 0.3 1.0
Calcined HTc
Published on 08 November 2018. Downloaded by Cornell University Library on 9/2/2020 11:01:30 AM.

109 2.03 1.8 0.9


n-Butylamined — 13.7 4.9 0.4
a
100 °C. b
Pretreated at 700 °C for 2 h. c Hydrotalcite (Mg/Al = 3). Pretreated at 500 °C for 2 h. d 0.05 mmol.

Table 3 Results of nitroaldol reaction of nitromethane and benzalde- formation of β-nitrostyrene by nucleophilic attack of
hyde and using ball-milled h-BN for 12 h
deprotonated nitromethane. Kinetic isotope effect experi-
Nitroaldol reactiona ments using CD3NO2 revealed that the abstraction of a pro-
Sample Conv./% Selectivity to β-nitrostyrene/% ton from nitromethane by base sites of the BN catalyst was
the rate-determining step. This is in good agreement with the
BN 150 rpm 0 0
BN 300 rpm 10 37 present study because the base sites are mainly responsible
BN 400 rpm 68 77 for the catalytic activity.
BN 500 rpm 20 50 The catalytic activity of ball-milled h-BN was further inves-
BN 600 rpm 15 29 tigated by transformation of glucose. Fig. 8 shows the results
a
Reaction conditions: benzaldehyde (0.5 mmol), nitromethane (1.25 of transformation of glucose using ball-milled h-BN catalysts
mmol), catalyst (50 mg), toluene (2 mL), 100 °C, 12 h. at 40 °C for 15 min. The ball-milled samples exhibited high
conversion with the formation of fructose whereas pristine
h-BN did not. Conventional solid bases including MgO and
ball-milled h-BN catalysts for 12 h. BN 400 rpm showed the hydrotalcite showed moderate glucose conversion with no
highest conversion (68%) with the highest selectivity to formation of fructose, suggesting that the apparent glucose
β-nitrostyrene (77%). conversion was probably due to the adsorption of glucose on
Our previous study has investigated the acid–base the solids at 40 °C. The sample BN 400 rpm showed the
bifunctionality of the ball-milled h-BN for the nitroaldol reac- highest conversion among the samples tested, which is in
tion,25,26,49 and the proposed reaction sequence is shown in good agreement with the results of the nitroaldol reaction.
Scheme 1. The reaction involves (1) the abstraction of a pro- The base-catalyzed aldose-ketose isomerization is named the
ton from nitromethane by an amino group of the BN catalyst, Lobry de Bruyn–van Ekenstein transformation in which de-
(2) formation of an imine intermediate by nucleophilic attack protonation of the α-carbonyl carbon of glucose with a base
of an amino group of the catalyst on the aldehyde group of is the first step. It can be considered that a hydroxyl group of
benzaldehyde with the aid of polarization of the carbonyl the BN catalyst would help the polarization of the carbonyl
group by a hydroxyl group of the BN catalyst,50,51 and (3) the group, which accelerates the isomerization as same as the

Fig. 8 Results of glucose transformation in methanol at 40 °C using


Scheme 1 Proposed reaction sequence of nitroaldol reaction on the ball-milled h-BN. (Reaction conditions: glucose (0.17 mmol, 30 mg),
ball-milled h-BN catalyst.20 catalyst (20 mg), methanol (5 mL), 40 °C, 15 min.)

This journal is © The Royal Society of Chemistry 2019 Catal. Sci. Technol., 2019, 9, 302–309 | 307
View Article Online

Paper Catalysis Science & Technology

nitroaldol reaction. The yield of fructose of the sample BN 8 F. Yang, Y. Cao, Z. Chen, X. He, L. Hou and Y. Li, New J.
400 rpm was slightly lower than that of the sample BN 600 Chem., 2018, 42, 2718–2725.
rpm, probably due to the further transformation of fructose 9 X. Ma, J. Wen, S. Zhang, H. Yuan, K. Li, F. Yan, X. Zhang
such as retro-aldol reaction. and Y. Chen, ACS Sustainable Chem. Eng., 2017, 5,
10266–10274.
Conclusions 10 Z. Du, S. N. Ellis, R. A. Dunlap and M. N. Obrovac,
J. Electrochem. Soc., 2016, 163, A13–A18.
Hexagonal boron nitride solid base catalysts were prepared 11 B. D. Keller, N. Ferralis and J. C. Grossman, Nano Lett.,
Published on 08 November 2018. Downloaded by Cornell University Library on 9/2/2020 11:01:30 AM.

by simple ball-milling at several rotation speeds of a commer- 2016, 16, 2951–2957.


cial low-surface area boron nitride. XRD revealed that ball- 12 S. Milosevic, S. Kurko, L. Pasquini, L. Matovic, R. Vujasin, N.
milling at a rotation speed of 300 rpm or higher disrupted Novakovic and J. G. Novakovic, J. Power Sources, 2016, 307,
the layer structure of pristine h-BN. Nitrogen-adsorption mea- 481–488.
surements indicated that the surface areas were drastically 13 C. Orellana-Tavra, E. F. Baxter, T. Tian, T. D. Bennett,
increased from 3 m2 g−1 for pristine h-BN to ca. 400 m2 g−1 N. K. H. Slater, A. K. Cheetham and D. Fairen-Jimenez,
for h-BN 300 rpm and h-BN 400 rpm. At a rotation speed of Chem. Commun., 2015, 51, 13878–13881.
400 rpm and higher the surface areas decreased likely 14 K. Watanabe, T. Takashi and K. Hisao, Nat. Mater., 2004, 3,
due to agglomeration as indicated by SEM and XRD. FTIR 404–409.
and XPS measurements showed the formation of amino- and 15 L. Song, L. Ci, H. Lu, P. B. Sorokin, C. Jin, J. Ni, A. G.
hydroxyl groups on the surface of the ball-milled h-BN. The Kvashnin, D. G. Kvashnin, J. Lou, B. I. Yakobson and P. M.
number of base sites increased with the increase of ball- Ajayan, Nano Lett., 2010, 10, 3209–3215.
milling speeds. The ball-milled h-BN exhibited catalytic activ- 16 Z. Liu, Y. Gonf, W. Zhou, L. Ma, J. Yu, J. C. Idrobo, J. Jung,
ity in the nitroaldol reaction between nitromethane and benz- A. H. MacDonald, R. Vajtai, J. Lou and P. M. Ajayan, Nat.
aldehyde, in which h-BN 400 rpm was the most active cata- Commun., 2013, 4, 3541.
lyst. The base sites were responsible for the activity. 17 J. T. Grant, C. A. Carrero, F. Goeltl, J. Venegas, P. Mueller,
S. P. Burt, S. E. Specht, W. P. McDermott, A. Chieregato and
Conflicts of interest I. Hermans, Science, 2016, 354, 1570–1573.
18 Y. Fang and X. Wang, Angew. Chem., Int. Ed., 2017, 56,
There are no conflicts to declare.
15506–15518.
19 L. Shi, B. Yan, D. Shao, F. Jiang, D. Wang and A.-H. Lu,
Acknowledgements Chin. J. Catal., 2017, 38, 389–395.
This work was supported by a Grant-in-Aid for Challenging 20 Y. Honda, A. Takagaki, R. Kikuchi and S. T. Oyama, Chem.
Exploratory Research (No. 16K14474) and Science Research Lett., 2018, 47, 1090–1093.
(B) (No. 18H01785) of JSPS, Japan. A part of this work was 21 F. Guo, P. Yang, Z. Pan, X.-N. Cao, Z. Xie and X. Wang,
supported by the AIST Nanocharacterization Facility (ANCF) Angew. Chem., Int. Ed., 2017, 56, 8231–8235.
platform as a program of the Nanotechnology Platform of the 22 P. Wang, S. Orimo, T. Matsushima, H. Fujii and G. Majer,
Ministry of Education, Culture, Sports, Science and Technol- Appl. Phys. Lett., 2002, 80, 318–320.
ogy (MEXT), Japan. 23 L. Liu, Z. Xiong, D. Hu, G. Wu, B. Liu and P. Chen, Chem.
Lett., 2013, 42, 1415–1416.
Notes and references 24 D. Lee, B. Lee. K. H. Park, H. J. Ryu, S. Jeon and S. H. Hong,
Nano Lett., 2015, 15, 1238–1244.
1 K. Yamamoto, S. E. Borjas Garcia, F. Saito and A. 25 S. Torii, K. Jimura, S. Hayashi, R. Kikuchi and A. Takagaki,
Muramatsu, Chem. Lett., 2006, 35, 570–571. J. Catal., 2017, 355, 176–184.
2 T. Kurniawan, O. Muraza, K. Miyake, A. S. Hakeem, Y. 26 J. D. Bass, A. Solovyov, A. J. Pascall and A. Katz, J. Am. Chem.
Hirota, A. M. Al-Amer and N. Nishiyama, Ind. Eng. Chem. Soc., 2016, 128, 3737–3747.
Res., 2017, 56, 4258–4266. 27 V. E. Collier, N. C. Ellbracht, G. I. Lindy, E. G. Moschetta
3 X. Cheng, E. Fabbri, B. Kim, M. Nachtegaal and T. J. and C. W. Jones, ACS Catal., 2016, 6, 460–468.
Schmidt, J. Mater. Chem. A, 2017, 5, 13130–13137. 28 K. Motokura, Bull. Chem. Soc. Jpn., 2017, 90, 137–147.
4 H. Horie, A. Iwase and A. Kudo, ACS Appl. Mater. Interfaces, 29 K. Akutu, H. Kabashima, T. Seki and H. Hattori, Appl. Catal.,
2015, 7, 14638–14643. A, 2003, 247, 65–74.
5 I.-Y. Jeon, H.-J. Choi, S.-M. Jung, J.-M. Seo, M.-J. Kim, L. Dai 30 C. Moreau, R. Durand, A. Roux and D. Tichit, Appl. Catal., A,
and J.-B. Baek, J. Am. Chem. Soc., 2013, 135, 1386–1393. 2000, 193, 257–264.
6 T. D. Bennett and A. K. Cheetham, Acc. Chem. Res., 2014, 47, 31 A. Takagaki, M. Ohara, S. Nishimura and K. Ebitani, Chem.
1555–1562. Commun., 2009, 6276–6278.
7 C. Ciotonea, R. Averlant, G. Rochard, A.-S. Mamede, J.-M. 32 L. Opoczky, Powder Technol., 1977, 17, 1–7.
Giraudon, H. Alamdari, J.-F. Lamonier and S. Royer, 33 S. Fadda, A. Cincotti, A. Concas, M. Pisu and G. Cao, Powder
ChemCatChem, 2017, 9, 2366–2376. Technol., 2009, 194, 207–216.

308 | Catal. Sci. Technol., 2019, 9, 302–309 This journal is © The Royal Society of Chemistry 2019
View Article Online

Catalysis Science & Technology Paper

34 R. Geick, C. H. Perry and G. Rupprecht, Phys. Rev., 44 W. Lei, D. Portehault, R. Dimova and M. Antonietti, J. Am.
1966, 146, 543–547. Chem. Soc., 2011, 133, 7121–7127.
35 C. Zhi, Y. Bando, C. Tang, D. Golberg, R. Xie and T. 45 J. P. L. Perez, B. W. MacMahon, J. Yu, S. Schneider, J. A.
Sekigushi, Appl. Phys. Lett., 2005, 86, 213110. Boatz, T. W. Hawkins, P. D. McCrary, L. A. Flores, R. D.
36 T. Sainsbury, A. Satti, P. May. Z. Wang, I. McGovern, Y. K. Rogers and S. L. Anderson, ACS Appl. Mater. Interfaces,
Gun'ko and J. Coleman, J. Am. Chem. Soc., 2012, 134, 2014, 6, 8513–8525.
18758–18771. 46 Q. Weng, D. G. Kvashnin, X. Wang, O. Cretu, Y. Yang, M.
37 X.-F. Wu, Z.-H. Zhao, Y. Sun, H. Li, Y.-J. Wang, C.-X. Zhang, Zhou, C. Zhang, D.-M. Tang, P. B. Sorokin, Y. Bando and D.
Published on 08 November 2018. Downloaded by Cornell University Library on 9/2/2020 11:01:30 AM.

X.-D. Gong, Y.-D. Wang, X.-Y. Yang and Y. Liu, J. Inorg. Golberg, Adv. Mater., 2017, 29, 1700695.
Organomet. Polym., 2017, 27, 1142–1147. 47 D. Saha, G. Orkoulas, S. Yohannan, H. C. Ho, E. Cakmak, J.
38 C. Huang, C. Chen, X. Ye, W. Ye, J. Hu, C. Xu and X. Qiu, Chen and S. Ozcan, ACS Appl. Mater. Interfaces, 2017, 9,
J. Mater. Chem. A, 2013, 1, 12192–12197. 14506–14517.
39 M. I. Baraton, T. Merle, P. Quintard and V. Lorenzelli, 48 M. Boudart, Proc. 6th Int. Cong. Catal, ed. G. C. Bond, P. B.
Langmuir, 1993, 9, 1486–1491. Wells and F. C. Thompson, Chemical Society London, 1976,
40 P. S. Marchetti, D. Kwon, W. R. Shmidt, L. V. Interrante and pp. 1–9.
G. E. Maciel, Chem. Mater., 1991, 3, 482–486. 49 K. Motokura, M. Tada and Y. Iwasawa, Angew. Chem., Int.
41 D. Zielniok, C. Cramer and H. Eckert, Chem. Mater., Ed., 2008, 47, 9230–9235.
2007, 19, 3162–3170. 50 S. L. Hruby and B. H. Shanks, J. Catal., 2009, 263,
42 F. Zhang, K. Németh, J. Bareño, F. Dogan, I. D. Bloom and 181–188.
L. L. Shaw, RSC Adv., 2016, 6, 27901–27914. 51 G. Sartori, G. Bigi, R. Maggi, R. Sartorio, D. J. Macquarrie,
43 L. Fu, G. Chen. N. Jiang, J. Yu, C.-T. Lin and A. Yu, J. Mater. M. Lenarda, L. Storaro, S. Coluccia and G. Martra, J. Catal.,
Chem., 2016, 4, 19107–19115. 2004, 222, 410–418.

This journal is © The Royal Society of Chemistry 2019 Catal. Sci. Technol., 2019, 9, 302–309 | 309

You might also like