Вы находитесь на странице: 1из 11

Article

Cite This: Ind. Eng. Chem. Res. 2019, 58, 2846−2856 pubs.acs.org/IECR

Acquiring Clean and Highly Dispersed Nickel Particles (ca. 2.8 nm)
by Growing Nickel-Based Nanosheets on Al2O3 as Efficient and
Stable Catalysts for Harvesting Cyclohexane Carboxylic Acid from
the Hydrogenation of Benzoic Acid
Weijie Lian,† Bo Chen,† Bingyu Xu, Song Zhang, Zhe Wan, Dan Zhao,* Ning Zhang,
and Chao Chen*
Key Laboratory of Jiangxi Province for Environment and Energy Catalysis, College of Chemistry, Nanchang University, Nanchang,
Jiangxi 330031, China
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV OF GOTHENBURG on October 22, 2019 at 13:33:31 (UTC).

ABSTRACT: By growing nickel-based nanosheets on Al2O3, a


nickel catalyst present as clean uniform nickel particles ca. 2.8
nm in size was feasibly produced from the two-dimensional
(2D) nanoprecursor predispersing strategy, and investigated
for the hydrogenation of benzoic acid to cyclohexane
carboxylic acid. A 98% yield of cyclohexane carboxylic acid
and a 15−40-fold enhanced turnover frequency (TOF) value
(31.4 s−1) were achieved with the catalyst. These values were
superior to those achieved (yield = 28%−40%, time of flight
TOF = 0.8−2.0 s−1) for traditional catalysts with larger nickel
particles (ca. 5.2−6.8 nm); interestingly, the superior activity
was re-examined, because of the well-maintained size of the
catalyst after continuous catalytic recycling. The kinetic
measurements further illuminated that the reaction activation energy could be clearly decreased by employing smaller nickel
particles, which could be responsible for the dramatic TOF enhancements. These results also suggest that the hydrogenation
with a nickel-based catalyst could be a structure-sensitive catalytic process.

1. INTRODUCTION application.22,23 Ni−Zr−B and Ni−Zr−B/γ-Al2O3 could


Direct hydrogenation of the benzene ring in a benzoic show ∼100% selectivity for CCA at 93% BA conversion, but
aromatic molecule was believed to be the most efficient way the performance was not reproducible in the subsequent
to synthesize cyclohexane derivatives,1,2 which has attracted reaction recycles, suggesting the poor stability of amorphous
considerable interest from both the pharmaceutical and nickel-based alloy catalysts for the reaction.24,25 To solve the
petrochemical industries, such as a safe drug intermediate problem, some researchers were interested in controllable
synthesis and diesel fuel production.3−6 In this field, the synthesis of nickel crystalline catalysts from some new
synthesis of cyclohexane carboxylic acid (CCA) from the preparation procedures. Bai et al.26 successfully prepared
hydrogenation of benzoic acid (BA) is significant, since CCA is highly dispersed nickel nanoparticles with an average size of ca.
a vital organic intermediate for important pharmaceuticals, 2.8 nm on SiO2; this unconventional preparation was believed
such as praziquantel, caprolactam, and ansatrienin.7−9 to be responsible for the excellent activity and stability
However, the application of this promising synthesis was observed for their catalysts. Xia et al.27 also produced active
suspended by the absence of proper hydrogenation catalysts and stable Ni/CSC catalysts from microwave treatment
for the reaction.10,11 procedures for the hydrogenation of BA to CCA. However,
Some precious metal catalysts, such as Pd/C,12 Pd/CN,13,14 relatively harsher reaction conditions were required, because of
Pd/MCN,15 Ir/γ-Al2O3,16 Ru/MC,17 Rh/CN,18 Rh/TiO2,19 Pt the larger (ca. 9.4 nm) nickel particles employed in the work of
nanowire,20 and RuPd/CN21 have been widely studied for the these researchers. These reports suggested that supported
chemical transformation. However, the high cost of these nickel particles could be a competitive candidate catalyst for
catalysts excludes their industrial potentials. As an alternative the efficient hydrogenation of benzoic acid, especially when the
route, recently, inexpensive metal catalysts, mostly focused on size of the nickel particles is well-controlled.
nickel-based catalysts, have been studied because of the
considerable activity of nickel exhibited for the reaction. Received: December 5, 2018
Early on, Raney nickel was used for the reaction and showed Revised: February 2, 2019
superior performance. However, the pyrophoric and environ- Accepted: February 7, 2019
mentally unfriendly nature of the catalyst limited its Published: February 7, 2019

© 2019 American Chemical Society 2846 DOI: 10.1021/acs.iecr.8b06037


Ind. Eng. Chem. Res. 2019, 58, 2846−2856
Industrial & Engineering Chemistry Research Article

The conventional methods to prepare supported metal alternately, in turn, with deionized water and ethanol until
catalysts, such as incipient-wetness impregnation or deposi- the filtrate pH was neutral. The precipitate was redispersed in
tion−precipitation,28,29 were not efficient to manipulate the ethanol under stirring for 12 h, then was filtered and dried at
size of the supported particles. The steps involved in these 60 °C in air for 12 h to obtain the precursor powder of nickel-
preparations, such as assembling the metal precursors and the based nanosheets (NS) on Al2O3. Notably, although reductant
combination of metal precursors with supports, were largerly NaBH4 was employed in the above preparation of NS
unordered or uncontrollable processes, leading to the precursor, in a report of Zhang et al., it was found that the
difficulties in obtaining small uniform particles. Prereduction reduced nickel could suffer from an amorphous boron-
of metal precursors with the assistance of surfactants (as mediated oxidation process in aqueous solution with the
protection to prevent the excessive aggregation of metal atoms prolonged period, to result in the continuous and deep
present from reduction) before access of the support became a oxidation of nickel.38 In our case, the phenomenon would
mainstream technique for the controllable synthesis of metal appear to be similar to the observation from the literature,
particle catalysts.30−32 Although the strategy was highly given that a long period (2 h) was required to allow growth of
efficient to synthesize highly dispersed small metal nano- the nanosheet. In order to obtain the supported nickel catalysts
particles, the surfactant protector adhered to the surface of in metallic state as much as possible, the precursor powder was
small metal particles, which, in turn, would create difficulties in further reduced in 50 mL/min hydrogen flow at 400 °C for 3
the subsequent process of loading metal particles onto the h; the obtained catalyst was denoted as Ni/Al2O3-NS. Notably,
support. This adhesion could also prevent the access of to avoid excessive oxidation of nickel metal in a catalyst in air,
reactants to the metal surface, leading to the confined metal the characterizations and catalytic measurements of the catalyst
loading or depressed catalytic efficiency of the catalysts.33,34 should be performed as soon as possible after producing the
We noticed that two-dimensional (2D) nanomaterials were catalyst through the thermal reduction treatments.
more feasible and exhibited unique features for modulating For reference, two other Ni/Al2O3 catalysts were also
structural characteristics and electronic properties, and these prepared from traditional incipient-wetness impregnation
materials have been regarded as effcient catalyst precursors, (IWI) and deposition precipitation (DP) procedures. The
such LDH and LMH nanosheets.35−38 Considering that 2D amount of γ-Al2O3 support and nickel nitrate was the same as
nanomaterials could be more able to control the exposed active those in the above preparation of Ni/Al2O3-NS, and the last
sites than conventional preparation systems, in this work, a reduction step with hydrogen and the corresponding
new preparation of supported nickel nanoparticles was conditions to obtain the catalysts also were the same. The
designed by predispersing nickel precursors as 2D nanosheets two catalysts were denoted as Ni/Al2O3-IWI and Ni/Al2O3-
grown on Al2O3 for the catalytic hydrogenation of BA to CCA. DP, respectively. Briefly, the mixture of calcinated γ-Al2O3 and
The main interesting feature of the 2D nanoprecursor nickel nitrate solution was dried at 120 °C for 12 h. After
predispersing control was that the clean, highly dispersed washing with deionized water and ethanol, the powder was
nickel particles with a typical size of ca. 2.8 nm could be dried at 60 °C for 12 h to obtain the precursor of Ni/Al2O3-
feasibly obtained. These materials even endured a severe IWI. For Ni/Al2O3-DP, 3.6 g of nickel nitrate and 7.3 g of the
thermal-reduction treatment under surfactant-free conditions, Al2O3 support were mixed in 250 mL of deionized water. Then
subsequently leading to the attractive catalytic activity and 7.4 g of urea was added to the solution, and the suspension was
stability upon the more prominent interaction between the stirred at 80 °C for 16 h to complete the deposition−
size-controlled small nickel particles and Al2O3 support, in precipitation process. The precipitate was filtered and washed
comparison with samples prepared by traditional methods. with deionized water and ethanol. The obtained paste then was
These findings could provide an attractive reference to dried at 60 °C for 12 h to obtain the precursor.
synthesize efficient metal catalysts for catalytic hydrogenation 2.3. Characterizations. The actual nickel loading in the
applications. catalyst was quantified by an inductively coupled plasma−
optical emission spectrometry (ICP-OES) system (Agilent
2. EXPERIMENTAL SECTION Technologies, Model 5100 ICP-OES).
2.1. Materials. All reagents of analytical purity, including X-ray diffraction (XRD) data of the sample powder were
dodecane (Aldrich), γ-Al2O3, benzoic acid, ethanol, nickel collected at ambient temperature on a Puxi DX-3 diffrac-
nitrate (Ni(NO3)2·6H2O), urea, NaOH, NaBH4, and cyclo- tometer operated at 40 kV, 40 mA for Cu Kα (λ = 1.5418 Å),
hexane were purchased and used without further purification. with a scan speed of 0.125°/min and a step size of 0.02° in the
2.2. Preparation Procedures of Catalysts. Al2O3- 2θ range of 10°−80°.
supported nickel catalysts were prepared for the catalytic Transmission electron microscopy (TEM) was performed
hydrogenation of benzoic acid (BA) to cyclohexane carboxylic using a Model JEM-2100 TEM microscope (JEOL Co., Japan)
acid (CCA) in this work. The goal catalyst was prepared from operating at 200 kV to observe the morphology and size
a 2D nanoprecursor predispersing control by growing nickel- distribution of samples.
based nanosheets on Al2O3; details are given in the following The surface chemical property of the catalysts was further
description. 7.3 g of commercial γ-Al2O3 support calcined by analyzed by X-ray photoelectron spectroscopy (XPS). The
pseudo-boehmite (550 °C, 4 h) was immersed in an aqueous spectra were obtained at ambient temperature under ultrahigh
solution of nickel nitrate (3.6 g in 5 mL of deionized water) vacuum on an Axis Ultra DLD electron spectrometer. During
and dispersed under ultrasound for 30 min. The mixture was data processing, the binding energy was calibrated with
heated at 120 °C for 12 h and changed to a green paste. The reference to the C 1s peak of the contaminant carbon at
paste was moved into an ice bath and reduced by adding 25 284.6 eV.
mL of NaBH4/NaOH aqueous solution (containing 1.9 g of The H2 temperature-programmed reduction (H2-TPR)
NaBH4 and 0.2 g of NaOH) dropwise under vigorous stirring measurements were performed in a quartz U-tube reactor.
for 2 h. The precipitate then was filtered and washed Prior to these measurements, all the catalyst samples were
2847 DOI: 10.1021/acs.iecr.8b06037
Ind. Eng. Chem. Res. 2019, 58, 2846−2856
Industrial & Engineering Chemistry Research Article

pretreated with an argon flow (30 mL/min, 1 h, 200 °C). The 0063), respectively. In addition to the phase feature of Al2O3
TPR profiles were obtained by passing carrier gas (10% H2 in support, two weak peaks could also be observed on the Ni/
Ar) through the catalyst samples (30 mL/min, ramping from Al2O3-IWI and Ni/Al2O3-DP samples, at ∼51.8° and ∼76.4°,
50 °C to 800 °C at 10 °C/min) from a Micromeritics Auto which could be attributed to the features of the metallic nickel
Chem II 2920 chemisorption analyzer. phase.39 In contrast, these Ni diffractions peaks could not be
Temperature-programmed desorption of H2 (H2-TPD) of resolved on the Ni/Al2O3-NS catalyst even in the local area, as
the catalyst was conducted in a quartz tube reactor shown by the enlarged illustration inset in Figure 1, suggesting
(Micromeritics AutoChem II 2920) that was equipped with that the nickel entities in Ni/Al2O3-NS catalyst could be
a thermal conductivity detection (TCD) system. Typically, 50 present as small particles (<ca.5.0 nm) or an amorphous form.
mg of sample was sealed in the reactor, pretreated with an Because the highest Ni diffraction intensity was shown for the
argon flow (30 mL/min, 1 h, 150 °C). The sample was Ni/Al2O3-IWI catalyst, the size order of nickel entities for the
reduced at 400 °C for 3 h in 50 mL/min H2 flow and cooled to three catalysts could be approximately estimated as Ni/Al2O3−
room temperature in H2 flow. After H2 chemisorption at 30 °C IWI > Ni/Al2O3-DP > Ni/Al2O3-NS.
for 1 h, by removing the free adsorbate with an argon stream at To confirm the estimation, TEM images for the three
the rate of 30 mL/min for 30 min, the program was performed catalysts were collected, as shown in Figure 2. For comparison,
in the temperature range of 30−500 °C with a heating rate of the morphologies of the precursors of the three catalysts were
10 °C/min to record the TCD signal. collected in the first row in Figure 2. In comparison with the
2.4. Catalytic Measurements. All experiments were blocklike images of precursors for the Ni/Al2O3-IWI and Ni/
perfomed in a Teflon reactor with 50 mL capacity contained Al2O3-DP catalysts, the precursor of the Ni/Al2O3-NS catalyst
in a stainless-steel autoclave equipped with a mechanical exhibited a unique flower edge around a block powder with a
engineering stirrer and an electric heating system. BA (61 mg), deep color; in fact, such a flower-edge appearance was initially
catalyst (50 mg) and cyclohexane (10 mL) were added into present on nickel-based nanosheets prepared without the
the reactor. The reaction system was filled with H2 five times presence of Al2O3 (shown by the illustration in Figure 2c-1),
then by evacuation to exclude air, and followed by increasing indicating that the precursor of Ni/Al2O3-NS catalyst was
the pressure to 1.0 MPa. The typical reaction was allowed at composed of the nickel-based nanosheets grown on Al2O3.
150 °C under the stirring rate of 400 rpm for 2.5 h. The Although such morphological differences disappeared after the
mixture after reaction was cooled to ambient temperature and thermal-reduction procedure for producing catalysts, the
filtered out of the catalyst; the filtrate was measured by gas gradually enlarged images of catalysts disclosed the size
chromatography (GC) (Agilent, Model GC-7820A) and the difference of nickel particles among the samples. As shown,
product were further ascertained by gas chromatography−mass the d-space values of typical particle from the three catalysts
spectroscopy (GC-MS) (Agilent, Model 7890B-5077MS). exhibited by the HRTEM images all matched well with the
0.203 nm value for crystalline Ni(111),40 indicating the
3. RESULTS AND DISCUSSION formation of metallic nickel particles after the thermal-
3.1. Physiochemical Properties of Catalysts. Nickel reduction treatment. The size distributions of particles
loadings in three catalysts, Ni/Al2O3-IWI, Ni/Al2O3−DP, and among the Ni/Al2O3-IWI, Ni/Al2O3-DP, and Ni/Al2O3-NS
Ni/Al2O3-NS, were ascertained by ICP-OES analysis as 10.0, catalysts were further determined as be 6.8 ± 1.5, 5.2 ± 1.2,
10.6, and 10.3 wt %, respectively, indicating comparable nickel and 2.8 ± 0.5 nm, respectively, which verified the size order
content among three catalysts. The entire structures of three estimation from XRD analysis.
catalysts were studied with the Al2O3 support as reference by In addition, the size order also indicated that growing nickel-
XRD measurements, and the corresponding diffraction based nanosheets on Al2O3 as a precursor could be more
patterns are given in Figure 1. All samples exhibited similar beneficial to obtaining small nickel nanoparticles, in compar-
peak at 2θ values of 19.5°, 37.6°, 45.8°, and 66.8°; these peaks ison with the two traditional methods, especially when a
matched well with the crystalline planes (111), (311), (400), considerable nickel loading (∼10.0 wt % for three catalysts)
and (440) of γ-Al2O3 (Powder Diffraction File (PDF) No. 29- was achieved. Notably, in most preparations to produce metal
nanoparticles smaller than ca. 5.0 nm, the protecting
surfactants were always employed to prevent the aggregation
of reduced metal to bigger particles; the performance of the
obtained metal particles would inevitably suffer from the
influence of adhered surfactants on the metal surface. In
contrast, there no surfactants were used in our preparations,
especially for the preparation of ca. 2.8 nm nickel particles
from the 2D nanoprecursor predispersing strategy. Thus, the
properties of metal catalysts on the relatively clean metal
surface could be better to be recognized from the character-
izations.
The redox states of catalysts were measured by XPS and H2-
TPR spectra, as given in Figure 3. Two main peaks located at
∼851.8 and 855.4 eV with a shakeup peak at 861.4 eV were
resolved in the Ni 2p3/2 spectra (Figure 3A) for the three
catalysts, indicating the co-occurrence of Ni(0) and Ni(II)
distributions on the surface of the three catalysts.41−43 The
Figure 1. X-ray diffractions of catalysts: o, Al2O3; a, Ni/Al2O3-IWI; b, distributions were further resolved as the atomic ratio values of
Ni/Al2O3-DP; and c, Ni/Al2O3-NS. Ni(0)/Ni(II), which were collected in Table 1. As shown, with
2848 DOI: 10.1021/acs.iecr.8b06037
Ind. Eng. Chem. Res. 2019, 58, 2846−2856
Industrial & Engineering Chemistry Research Article

Figure 2. TEM data of samples: (a) Ni/Al2O3-IWI, (b) Ni/Al2O3-DP, and (c) Ni/Al2O3-NS. Using the labeling format of X = a, b, or c, the
following images are shown: X-1, images of precursors; X-2, images of catalysts; X-3, local images of catalysts; and X-4, HRTEM images of particles
in catalysts. Size distributions of particles in catalysts are shown in panels X-5 (where X = a, b, or c).

Figure 3. (A) XPS (Ni 2p3/2) spectra and (B) H2-TPR profiles of catalysts: a, Ni/Al2O3-IWI; b, Ni/Al2O3-DP; and c, Ni/Al2O3-NS.

Table 1. Physicochemical Parameter of Catalysts


nickel contenta particle sizeb metal dispersiond H desorption amountd TOF activation energy, Eae
catalyst (wt %) (nm) Ni(0)/Ni(II)c (%) (μmol gcat−1) (s −1) (kJ mol−1)
Ni/Al2O3-IWI 10.0 6.8 ± 1.5 0.09 6.9 113.9 0.8 56.9
Ni/Al2O3-DP 10.6 5.2 ± 1.2 0.11 6.6 116.1 2.0 52.7
Ni/Al2O3-NS 10.3 2.8 ± 0.5 0.05 12.2 201.0 31.4 41.9
a
Based on ICP-OES measurements. bStatistic results derived from TEM tests. cSurface atomic ratio determined by XPS measurements. dMeasured
from H2-TPD characterizations. eCalculated from kinetic measurements.

the decreasing order of size among the three catalysts, the obvious peak shifts toward higher binding energy occurred
Ni(0)/Ni(II) ratio also decreased, which could be observed with the decrease in particle size among catalysts, suggesting a
because the smaller the nickel particle size was, the higher the stronger interaction between the Al2O3 support and nickel
surface energy or more active the particle surface was, particles would occur when smaller nickel particles adhered to
facilitating the surface oxidation of particles.44 To eliminate the oxide support.45,46 The trend was further estimated with
the surface oxidation influence on catalytic performance, H2-TPR profiles from Figure 3B. Two main reduction peaks
before catalytic measurements, all catalysts were freshly made occurred for the three catalysts, which could be roughly
with the thermal reduction treatment. In addition, more distinguished according to the reduction temperature. One set
2849 DOI: 10.1021/acs.iecr.8b06037
Ind. Eng. Chem. Res. 2019, 58, 2846−2856
Industrial & Engineering Chemistry Research Article

of peaks appearing below 300 °C could be attributed to the all three catalysts (curves a, b, and c) showed obvious
reduction of nickel particles; the other set initiated from 350 conversion and high selectivity (>98%) for cyclohexane
°C correlated with nickel species on the Ni/Al2O3 inter- carboxylic acid (CCA). In comparison with the low conversion
face.42,47,48 Interestingly, the two sets of peaks showed different (26%−40%) on Ni/Al2O3-IWI and Ni/Al2O3-DP, benzoic acid
change trends in temperature shift with particle size differences could be almost completely converted to CCA with a yield of
on the three catalysts. The lowest temperature of ∼140 °C for 98% on Ni/Al2O3-NS within 2.5 h. In addition, the initial
the reduction of nickel particles was observed for the Ni/ benzoic acid conversion within 0.5 h on Ni/Al2O3-NS was
Al2O3-NS catalyst with the smallest nickel particles (ca. 2.8 clearly faster than that with the other two catalysts. A control
nm), which could be attributed to the higher reactivity of experiment involving filtering Ni/Al2O3-NS out of reaction
smaller nickel particles with H2.49,50 However, the highest solutions after 1 h was also performed, as shown by the dotted
temperature of ∼512 °C for the reduction of interfacial nickel curve in Figure 5A. The reaction stopped once the catalyst was
species was also observed for the catalyst, suggesting a stronger removed, indicating that the supported nickel particles were
oxidation influence from Al2O3 on smaller nickel particles.51−54 responsible for the efficient conversion of benzoic acid. Here,
Therefore, both XPS and H2-TPR characterizations implied note that 0.24 wt % of Na residue was found on the Ni/Al2O3-
that a more prominent Ni−Al2O3 interaction occurred for the NS catalyst, because of the Na-contained reagent used in the
Ni/Al2O3-NS catalyst, in comparison with other two catalysts preparation of the catalyst, but Na was not detected in Ni/
with bigger nickel particles. Al2O3-IWI or Ni/Al2O3-DP. Researchers have noted that Na+
Before catalytic hydrogenation measurements, the H2- could play an important role in the catalytic performance of
adsorption−desorption features of freshly prepared catalysts nickel-based catalysts.56 To evaluate this influence, Ni/Al2O3-
were determined by H2-TPD profiles, as shown in Figure 4. IWI and Ni/Al2O3-DP samples with deliberately introduced
Na+ at contents of 0.26 and 0.28 wt %, respectively, together
with a Ni/Al2O3-NS sample with 0.56 wt % of Na residue were
also employed for the reaction. Their catalytic performance
curves are labeled as a′, b′, and c′ in Figure 5A. The curves of
the three catalysts with additional introduced Na almost
overlapped with the curves of their respective counterpart with
low Na loadings; these results indicated that the content of
Na+ present in Ni/Al2O3 catalysts was not a governing factor
for their catalytic performance in the hydrogenation of benzoic
acid to CCA.
To estimate the catalytic stability of the catalysts, the recycle
uses of the catalysts for the reaction are compared in Figures
5B−D. As shown, the yield of cyclohexane carboxylic acid
significantly decreased over repeated used of Ni/Al2O3-IWI
and Ni/Al2O3-DP catalysts, demonstrated by their lowered
yield (from ∼30% to ∼12%); In contrast, a slight yield
Figure 4. H2-TPD profiles of catalysts: a, Ni/Al2O3-IWI; b, Ni/Al2O3- decrease from 98% to 91% after five catalytic cycles was
DP; and c, Ni/Al2O3-NS. observed for Ni/Al2O3-NS; the trend was parallel with the
benzoic acid conversion change during catalytic cycles, which
Similar H2-desorptions exhibiting a main single peak at could be attributed to the gradual weight loss from the catalyst-
approximately 100 °C were observed for the three catalysts, solution separation process between the different catalytic
indicating almost identical H2 activation feature among the cycles. The above comparison indicated that the Ni/Al2O3-NS
three catalysts. Thus, the area resolved from the desorption catalyst with smaller nickel particles was more active and stable
peak could be used to estimate and compare the number of Ni than the other two catalysts with bigger nickel particles.
active sites among three catalysts. According to a standard Because the reactant benzoic acid was a type of acid, such a
estimation method with known Pt/Al2O3 catalyst, as described feature could play an important role in the stability of nickel
in ref 55, H desorption (or active site) numbers were catalysts, the reaction solutions from catalytic recycling
determined and are collected in Table 1. The H-desorption processes for the three catalysts were all carefully collected
number on Ni/Al2O3-NS was approximately twice the values and analyzed by ICP-OES. There was no evident amount of
for the Ni/Al2O3-IWI and Ni/Al2O3-DP catalysts, suggesting metal species (including nickel and aluminum) found in any of
that the active surface area value or active site number on Ni/ the solutions. In addition, we also measured the nickel loading
Al2O3-NS could be twice as high as those on the other two on recycled catalysts; the nickel loading deviations were
catalysts. This finding was reasonable, since the Ni dispersion estimated within 0.4 wt %, according to the basic nickel loading
on Ni/Al2O3-NS was almost two times higher than that on the of ∼10.0 wt % for the three used catalysts. These results
other two catalysts because the size ratio of nickel particles was suggested that the leakage of metal ions from the particles was
∼50% between Ni/Al2O3-NS and the other two catalysts. not obvious for the catalysts studied in this work.
3.2. Catalytic Performance of Catalysts. The catalytic To illuminate the possible change of catalysts during
performance values for the hydrogenation of benzoic acid to catalytic recycles, three catalysts were reobserved by TEM
cyclohexane carboxylic acid for the three catalysts are collected measurements after repeat uses, as shown in Figure 6. In
in Figure 5. Figure 5A shows the dependence of benzoic acid comparison with the original size of the fresh samples, the
conversion on reaction time for the catalysts. With the nickel particles on Ni/Al2O3-IWI and Ni/Al2O3-DP catalysts
insignificant benzoic acid conversion observed for the reaction all increased in size by a factor of ∼3, on average, after five
system containing bare Al2O3 as the blank reference (curve o), catalytic cycles. In contrast, a slight size increase for nickel
2850 DOI: 10.1021/acs.iecr.8b06037
Ind. Eng. Chem. Res. 2019, 58, 2846−2856
Industrial & Engineering Chemistry Research Article

Figure 5. (A): Dependence of benzoic acid conversion on reaction time over catalysts: o, Al2O3; a, Ni/Al2O3-IWI; b, Ni/Al2O3-DP; c, Ni/Al2O3-
NS (the control experiment on the catalyst was also performed); a′, 0.26%Na−Ni/Al2O3-IWI; b′, 0.28%Na−Ni/Al2O3-DP; and c′, 0.56% Na−Ni/
Al2O3-NS. (B−D) Variations of yield of cyclohexane carboxylic acid during catalytic recycles over different catalysts, (B) for Ni/Al2O3-IWI (panel
(B)), for Ni/Al2O3-DP (panel (C)), and for Ni/Al2O3-NS (panel (D)).

particles from ca. 2.8 nm to ca. 4.2 nm was exhibited by Ni/ measurements, which could be an important factor to
Al2O3-NS. accelerate the reaction. To illuminate the influence among
In accordance with the observations, the XRD patterns of the catalysts, the initial reaction rate (within 20 min) was
the used catalysts in Figure 7 showed more distinct Ni normalized to the active sites to obtain the turnover frequency
crystalline diffractions, particularly on Ni/Al2O3-IWI and Ni/ (TOF) for each individual catalyst; the TOF values are listed
Al2O3-DP catalysts. These results clearly demonstrate that the in Table 1. Interestingly, a dramatic enhancement in TOF (an
small nickel particles derived from the predispersed nickel- increase of ∼15-fold to 40-fold) on Ni/Al2O3-NS (31.4 s−1),
based nanosheets could adhere more tightly on the Al2O3 compared to Ni/Al2O3-IWI (0.8 s−1) and Ni/Al2O3-DP (2.0
surface than the larger nickel particles obtained from s−1), was observed. The active site number on Ni/Al2O3-NS
traditional incipient-wetness impregnation (IWI) and deposi- was just two times that of the other two catalysts (denoted by
tion precipitation (DP) procedures, i.e., the precursor-2D- the Ni dispersion in Table 1), indicating that the Ni sites on
dispersing strategy not only controlled the nickel particles to smaller particles of the Ni/Al2O3-NS catalyst were much more
produce smaller sizes, but also induced a stronger interaction efficient than those on the Ni/Al2O3-IWI and Ni/Al2O3-DP
between the nickel particles and Al2O3 support, in comparison catalysts. This result also suggested that the kinetic factor for
with traditional preparations. The synthetic effect could the entire reaction system should be further considered. Thus,
efficiently prevent the movement, aggregation, and growth of in Figure 8, the relationships between ln Ct and reaction time t
metal particles during thermal catalytic reactions and was at three temperatures (120, 140, and 150 °C (or 393, 413, and
responsible for the superior activity and stability observed for 423 K, respectively) are given. The relationships approach
the Ni/Al2O3-NS catalyst. This deduction was also supported linear dependences on the three catalysts (Figure 8A−C),
by the previous XPS and H2-TPR measurements, in which the suggesting that the conversions of benzoic acid all followed
stronger interaction between nickel particles and the Al2O3 pseudo-first-order kinetics for the three catalysts. With these
support in Ni/Al2O3-NS, compared to the other two catalysts, plots, the dependences of ln k (where k represents the reaction
has been illuminated. Furthermore, the above results indicated rate constant) on the reciprocal of reaction temperature (T−1)
that the well-controlled nickel particles with small size such as are given in Figure 8D; the reaction activation energy (Ea) for
3.0−4.0 nm, were a key factor in accelerating the hydro- Ni/Al2O3-IWI, Ni/Al2O3-DP, and Ni/Al2O3-NS were esti-
genation of the benzene ring of benzoic acid. mated as 56.9, 52.7, and 41.9 kJ mol−1, respectively. With the
In this study, one may argue that the smaller nickel particles Ea values (listed in the last column of Table 1), according to
would possess more active sites, as described from H2-TPD the Arrhenius equation, the influence of Ea on reaction rate (r)
2851 DOI: 10.1021/acs.iecr.8b06037
Ind. Eng. Chem. Res. 2019, 58, 2846−2856
Industrial & Engineering Chemistry Research Article

Figure 6. TEM images of catalysts (X-1) and particle size distribution in catalysts (X-2) after catalytic recycles (X = a, b, or c): a, Ni/Al2O3-IWI; b,
Ni/Al2O3-DP; and c, Ni/Al2O3-NS.

and TOF was determined for the catalysts, in which R is the 12, respectively. Correspondingly, the experimental ratios with
universal gas constant (R = 8.314 J mol−1 K−1), N is active site the TOF values listed in Table 1 were determined as
number of catalyst (which is roughly estimated by the TOF(NS)/TOF(IWI) = 42 and TOF(NS)/TOF(DP) = 16,
desorption amount H, listed in Table 1), and T is temperature respectively. The experimental ratios matched well with the

ij k yz
(chosen to be 423 K). theoretical ratios resolved from Ea, which indicated that the

lnjjj 1 zzz = a2
kinetic for the hydrogenation of benzoic acid to cyclohexane
jk z
E − Ea1
k 2{
carboxylic acid with Al2O3-supported nickel catalysts would be
RT (1) intensively accelerated by lowering the reaction activation
energy as a result of employing smaller nickel particles, i.e., the
r1 k reaction could present structure-sensitive catalytic behavior on
= 1
r2 k2 (2) oxide-supported nickel catalysts, since the reaction activity was
dependent directly on the size of the nickel particles.57 In fact,
TOF1 kN such an influence of metal particle size on its catalytic activity
= 1 2 has long been a hot topic for heterogeneous catalysis systems.
TOF2 k 2N1 (3)
Baranova et al. found that smaller platinum particles loaded on
Based on eqs 1−3, the TOF ratio between Ni/Al2O3-NS and zirconium-based supports would be more favorable for the
other two reference catalysts would be theoretically calculated catalytic oxidation of ethylene and CO due to the lowered
as TOF(NS)/TOF(IWI) = 40 and TOF(NS)/TOF(DP) = activation energies, compared to larger particles.58 In an earlier
2852 DOI: 10.1021/acs.iecr.8b06037
Ind. Eng. Chem. Res. 2019, 58, 2846−2856
Industrial & Engineering Chemistry Research Article

be differentiated as geometrical and electronic factors. In view


of the geometrical factor, the exposed atoms on the corner or
edge of a metal particle would be more active than those on
the flat surface, because of the more unsaturated coordination
state of corner or edge atoms. The smaller the size of a metal
particle, the higher the percentage of such more highly active
sites, which leads to the activity enhancement.60,61 Considering
the electronic aspect, metal nanoparticles with smaller size
would show enhanced electron density near the Fermi level
(EF), narrowed d-band, and higher-lying d-band center.62
These electronic features would be more favorable to facilitate
the interaction of the metal with reactant to accelerate the
reaction. Therefore, for supported metal catalysts, particularly
those with small metal nanoparticles, the influence from the
metal/support interface and the nature of the metal particle
Figure 7. X-ray diffraction (XRD) spectra of catalysts after catalytic could conjunctively govern their catalytic activity. In our case,
recycles: o, Al2O3; a, Ni/Al2O3-IWI; b, Ni/Al2O3-DP; and c, Ni/ these factors could coexist for the structure-sensitive catalytic
Al2O3-NS. behavior observed for the Ni/Al2O3 catalysts in the catalytic
hydrogenation of benzoic acid to cyclohexane carboxylic acid.
Beyond all doubt, more-detailed investigations or more
report, such enhanced catalytic activity by using smaller metal contrast experiments were required to clearly clarify the
particles was believed to be correlated with the enlarged concrete contribution of these factors; given the topic of the
metal−support interface, since more-active species could form current work, such investigations could be expected in the
along the interface, to accelerate the reaction.59 Besides the future.
metal−support interaction, the feature change of the metal Generally, it can be concluded that the 2D-nanoprecursor-
particle itself with size could also play an important role in predispersing strategy (growing 2D nickel-based nanosheets on
governing its catalytic performance. Generally, the effects could Al2O3 as a catalyst precursor) employed for designing Ni/

Figure 8. Kinetic measurements. (A−C) Relation lines between ln Ct and reaction time (t) over different catalysts: for Ni/Al2O3-IWI (panel (A)),
for Ni/Al2O3-DP (panel (B)), and for Ni/Al2O3-NS (panel (C)). (D) Dependences of ln k on the reciprocal of reaction temperature (T−1) over
catalysts: a, Ni/Al2O3-IWI; b, Ni/Al2O3-DP; and c, Ni/Al2O3-NS.

2853 DOI: 10.1021/acs.iecr.8b06037


Ind. Eng. Chem. Res. 2019, 58, 2846−2856
Industrial & Engineering Chemistry Research Article

Al2O3-NS catalyst was advanced not only in controlling (7) Moore, B. S.; Cho, H.; Casati, R.; Kennedy, E.; Reynolds, K. A.;
supported nickel as clean and small nanoparticles, but also in Mocek, U.; Beale, J. M.; Floss, H. G. Biosynthetic studies on
offering a positive metal−support interaction, which, in turn, ansatrienin A. Formation of the cyclohexanecarboxylic acid moiety. J.
leads to dramatic enhancements in both activity and stability, Am. Chem. Soc. 1993, 115, 5254.
(8) Shinkai, H.; Nishikawa, M.; Sato, Y.; Toi, K.; Kumashiro, I.;
compared to the catalysts prepared from traditional methods.
Seto, Y.; Fukuma, M.; Dan, K.; Toyoshima, S. N-(Cyclohexylcarbon-
yl)-D-phenylalanines and Related Compounds. A New Class of Oral
4. CONCLUSIONS Hypoglycemic Agents. 2. J. Med. Chem. 1989, 32, 1436.
Clean, small nickel nanoparticles with a uniform average size of (9) Shinkai, H.; Toi, K.; Kumashiro, I.; Seto, Y.; Fukuma, M.; Dan,
ca. 2.8 nm indicated on Al2O3 were feasibly obtained by K.; Toyoshima, S. N-acylphenylalanines and related compounds. A
growing nickel-based nanosheets on Al2O3 as a precursor. The new class of oral hypoglycemic agents. J. Med. Chem. 1988, 31, 2092.
2D nanoprecursor predispersing strategy was notable for not (10) Zhang, P.; Wu, T. B.; Hou, M. Q.; Ma, J.; Liu, H. Z.; Jiang, T.;
only achieving control of the production of small nanoparticles, Wang, W. T.; Wu, C. Y.; Han, B. X. The Hydrogenation of Aromatic
but also inducing a more positive interaction between the Compounds under Mild Conditions by Using a Solid Lewis Acid and
Supported Palladium Catalyst. ChemCatChem 2014, 6, 3323.
metal and support, compared to traditional methods, which, in (11) Grootendorst, E. J.; Pestman, R.; Koster, R. M.; Ponec, V.
turn, resulted in favorable catalytic activity and stability for the Selective Reduction of Acetic Acid to Acetaldehyde on Iron Oxides. J.
hydrogenation of benzoic acid to cyclohexane carboxylic acid Catal. 1994, 148, 261.
with the derived catalyst. The feasibility and the feature of the (12) Anderson, J. A.; McKenna, F. M.; Linares-Solano, A.; Wells, R.
strategy to produce active metal catalyst with high dispersion P. K. Use of Water as a Solvent in Directing Hydrogenation Reactions
under conditions such as the absence of surfactant, of Aromatic Acids over Pd/carbon Nanofibre Catalysts. Catal. Lett.
considerable metal loading, and thermal endurance made the 2007, 119, 16.
method a promising system to synthesize efficient supported (13) Xu, X.; Tang, M. H.; Li, M. M.; Li, H. R.; Wang, Y.
transition-metal catalysts for advanced heterogeneous catalysis, Hydrogenation of Benzoic Acid and Derivatives over Pd Nano-
such as scaled-up catalytic hydrogenation applications. particles Supported on N-Doped Carbon Derived from Glucosamine


Hydrochloride. ACS Catal. 2014, 4, 3132.
(14) Nie, R. F.; Jiang, H. Z.; Lu, X. H.; Zhou, D.; Xia, Q. H. Highly
AUTHOR INFORMATION
active electron-deficient Pd clusters on N-doped active carbon for
Corresponding Authors aromatic ring hydrogenation. Catal. Sci. Technol. 2016, 6, 1913.
*Tel.: +86-15879176996. E-mail: zhaodan@ncu.edu.cn (D. (15) Jiang, H. Z.; Yu, X. L.; Nie, R. F.; Lu, X. H.; Zhou, D.; Xia, Q.
Zhao). H. Selective hydrogenation of aromatic carboxylic acids over basic N-
*Tel.: +86-15179167359. E-mail: chaochen@ncu.edu.cn (C. doped mesoporous carbon supported palladium catalysts. Appl. Catal.,
Chen). A 2016, 520, 73.
(16) Tang, M. H.; Mao, S. J.; Li, X. F.; Chen, C. H.; Li, M. M.;
ORCID Wang, Y. Highly effective Ir-based catalysts for benzoic acid
Dan Zhao: 0000-0001-5618-9795 hydrogenation: experiment and theory guided catalyst rational design.
Chao Chen: 0000-0002-0913-0544 Green Chem. 2017, 19, 1766.
Author Contributions (17) Jiang, Z. L.; Lan, G. J.; Liu, X. Y.; Tang, H. D.; Li, Y. F. Solid
† state synthesis of Ru−MC with highly dispersed semi-embedded
These authors contributed equally.
ruthenium nanoparticles in a porous carbon framework for benzoic
Notes acid hydrogenation. Catal. Sci. Technol. 2016, 6, 7259.
The authors declare no competing financial interest. (18) Cao, Y.; Tang, M.; Li, M.; Deng, J.; Xu, F.; Xie, L.; Wang, Y. In

■ ACKNOWLEDGMENTS
This work was financially supported by the National Natural
Situ Synthesis of Chitin-Derived Rh/N−C Cataylsts: Efficient
Hydrogenation of Benzoic Acid and Derivatives. ACS Sustainable
Chem. Eng. 2017, 5, 9894.
(19) Nakanishi, K.; Yagi, R.; Imamura, K.; Tanaka, A.; Hashimoto,
Science Foundation of China (NSFC) (Nos. 21003071 and K.; Kominami, H. Ring hydrogenation of aromatic compounds in
21563018) and Doctoral Fund of Ministry of Education of aqueous suspensions of an Rh-loaded TiO2 photocatalyst without use
China (No. 20093601120007).


of H2 gas. Catal. Sci. Technol. 2018, 8, 139.
(20) Guo, Z. Q.; Hu, L.; Yu, H. H.; Cao, X. Q.; Gu, H. W.
REFERENCES Controlled hydrogenation of aromatic compounds by platinum
(1) Zhang, X. H.; Zhang, Q.; Wang, T. J.; Li, B. S.; Xu, Y.; Ma, L. L. nanowire catalysts. RSC Adv. 2012, 2, 3477.
Efficient upgrading process for production of low quality fuel from (21) Tang, M. M.; Mao, S. J.; Li, M. M.; Wei, Z. Z.; Xu, F.; Li, H. R.;
bio-oil. Fuel 2016, 179, 312. Wang, Y. RuPd Alloy Nanoparticles Supported on N-Doped Carbon
(2) Zhang, X. H.; Tang, W. W.; Zhang, Q.; Li, Y. P.; Chen, L. G.; as an Efficient and Stable Catalyst for Benzoic Acid Hydrogenation.
Xu, Y.; Wang, C. G.; Ma, L. L. Production of hydrocarbon fuels from ACS Catal. 2015, 5, 3100.
heavy fraction of bio-oil through hydrodeoxygenative upgrading with (22) Mészáros, S.; Halász, J.; Kónya, Z.; Sipos, P.; Pálinkó, I. Search
Ru-based catalyst. Fuel 2018, 215, 825. for a Raney-Ni type catalyst efficient in the transformation of excess
(3) Stanislaus, A.; Cooper, B. H. Aromatic Hydrogenation Catalysis: glycerol into more valuable products. Catal. Commun. 2014, 43, 116.
A Review. Catal. Rev.: Sci. Eng. 1994, 36, 75. (23) Pojer, P. M. Deuterated” Raney nickel: deuteration (reduction)
(4) Wang, Y.; Yao, J.; Li, H.; Su, D.; Antonietti, M. Highly Selective of alkenes, carbonyl compounds and aromatic rings. Proton-
Hydrogenation of Phenol and Derivatives over a Pd@Carbon Nitride deuterium exchange of “activated” aliphatic and aromatic ring
Catalyst in Aqueous Media. J. Am. Chem. Soc. 2011, 133, 2362. hydrogens. Tetrahedron Lett. 1984, 25, 2507.
(5) Glorius, F. Asymmetric hydrogenation of aromatic compounds. (24) Bai, G. Y.; Wen, X.; Zhao, Z.; Li, F.; Dong, H. X.; Qiu, M.
Org. Biomol. Chem. 2005, 3, 4171. Chemoselective Hydrogenation of Benzoic Acid over Ni−Zr−B−
(6) Ma, Y. F.; Xu, G. Y.; Wang, H.; Wang, Y. X.; Zhang, Y.; Fu, Y. PEG(800) Nanoscale Amorphous Alloy in Water. Ind. Eng. Chem. Res.
Cobalt Nanocluster Supported on ZrREnOx for the Selective 2013, 52, 2266.
Hydrogenation of Biomass Derived Aromatic Aldehydes and Ketones (25) Wen, X.; Cao, Y. Y.; Qiao, X. L.; Niu, L. B.; Huo, L.; Bai, G. Y.
in Water. ACS Catal. 2018, 8, 1268. Significant effect of base on the improvement of selectivity in the

2854 DOI: 10.1021/acs.iecr.8b06037


Ind. Eng. Chem. Res. 2019, 58, 2846−2856
Industrial & Engineering Chemistry Research Article

hydrogenation of benzoic acid over NiZrB amorphous alloy supported (42) Prakash, M. G.; Mahalakshmy, R.; Krishnamurthy, K. R.;
on γ-Al2O3. Catal. Sci. Technol. 2015, 5, 3281. Viswanathan, B. Selective hydrogenation of cinnamaldehyde on nickel
(26) Zhang, H. L.; Gao, X. J.; Ma, Y. Y.; Han, X.; Niu, L. B.; Bai, G. nanoparticles supported on titania: role of catalyst preparation
Y. A highly dispersed and stable Ni/mSiO2-AE nanocatalyst for methods. Catal. Sci. Technol. 2015, 5, 3313.
benzoic acid hydrogenation. Catal. Sci. Technol. 2017, 7, 5993. (43) Liu, L. J.; Lou, H.; Chen, M. Selective hydrogenation of furfural
(27) Lu, X. H.; Shen, Y.; He, J.; Jing, R.; Tao, P. P.; Hu, A.; Nie, R. to tetrahydrofurfuryl alcohol over Ni/CNTs and bimetallic Cu-Ni/
F.; Zhou, D.; Xia, Q. H. Selective hydrogenation of benzoic acid to CNTs catalysts. Int. J. Hydrogen Energy 2016, 41, 14721.
cyclohexane carboxylic acid over microwave-activated Ni/carbon (44) Guo, H. J.; Zhang, H. R.; Zhang, L. Q.; Wang, C.; Peng, F.;
catalysts. Mol. Catal. 2018, 444, 53. Huang, Q. L.; Xiong, L.; Huang, C.; Ouyang, X. P.; Chen, X. D.; Qiu,
(28) Torres, C. C.; Alderete, J. B.; Mella, C.; Pawelec, B. Maleic X. Q. Selective Hydrogenation of Furfural to Furfuryl Alcohol over
anhydride hydrogenation to succinic anhydride over mesoporous Ni/ Acid-Activated Attapulgite-Supported NiCoB Amorphous Alloy
TiO2 catalysts: Effects of Ni loading and temperature. J. Mol. Catal. A: Catalyst. Ind. Eng. Chem. Res. 2018, 57, 498.
Chem. 2016, 423, 441. (45) Xu, M.; He, S.; Chen, H.; Cui, G. Q.; Zheng, L. R.; Wang, B.;
(29) Menchavez, R. N.; Morra, M. J.; He, B. B. Co-Production of Wei, M. TiO2−x-Modified Ni Nanocatalyst with Tunable Metal−
Ethanol and 1,2-Propanediol via Glycerol Hydrogenolysis Using Ni/ Support Interaction for Water−Gas Shift Reaction. ACS Catal. 2017,
Ce−Mg Catalysts: Effects of Catalyst Preparation and Reaction 7, 7600.
Conditions. Catalysts 2017, 7, 290. (46) Wang, R.; Li, Y. H.; Shi, R. H.; Yang, M. M. Effect of metal−
(30) Karousis, N.; Tsotsou, G. E.; Evangelista, F.; Rudolf, P.; support interaction on the catalytic performance of Ni/Al2O3 for
Ragoussis, N.; Tagmatarchis, N. Carbon Nanotubes Decorated with selective hydrogenation of isoprene. J. Mol. Catal. A: Chem. 2011, 344,
Palladium Nanoparticles: Synthesis, Characterization, and Catalytic 122.
Activity. J. Phys. Chem. C 2008, 112, 13463. (47) Chary, K. V. R.; Rao, P. V. R.; Vishwanathan, V. Synthesis and
(31) Prabhuram, J.; Wang, X.; Hui, C. L.; Hsing, I. M. Synthesis and high performance of ceria supported nickel catalysts for hydro-
Characterization of Surfactant-Stabilized Pt/C Nanocatalysts for Fuel dechlorination reaction. Catal. Commun. 2006, 7, 974.
Cell Applications. J. Phys. Chem. B 2003, 107, 11057. (48) Zhang, X. H.; Tang, W. W.; Zhang, Q.; Wang, T. J.; Ma, L. L.
(32) Singha, R. K.; Yadav, A.; Agrawal, A.; Shukla, A.; Adak, S.; Hydrodeoxygenation of lignin-derived phenoic compounds to hydro-
Sasaki, T.; Bal, R. Synthesis of highly coke resistant Ni nanoparticles carbon fuel over supported Ni-based catalysts. Appl. Energy 2018, 227,
supported MgO/ZnO catalyst for reforming of methane with carbon 73.
dioxide. Appl. Catal., B 2016, 191, 165. (49) Bathla, A.; Pal, B. Catalytic Selective Hydrogenation and Cross
(33) Manikandan, D.; Divakar, D.; Rupa, A. V.; Revathi, S.; Preethi, Coupling Reaction Using Polyvinylpyrrolidone-Capped Nickel Nano-
M. E. L.; Sivakumar, T. Synthesis of platinum nanoparticles in
particles. ChemistrySelect 2018, 3, 4738.
montmorillonite and their catalytic behaviour. Appl. Clay Sci. 2007, (50) Vargas, H.; Morales, J.; Bokhimi, X.; Klimova, T. E. Effect of
37, 193.
the preparation method on the hydrogenation activity of Ni/SBA-15
(34) Wolfbeisser, A.; Sophiphun, O.; Bernardi, J.; Wittayakun, J.;
catalysts: Comparison of EDTA complexation and DPU. Catal. Today
Föttinger, K.; Rupprechter, G. Methane dry reforming over ceria-
2018, 305, 133.
zirconia supported Ni catalysts. Catal. Today 2016, 277, 234.
(51) Zhao, F. Z.; Gong, M.; Cao, K.; Zhang, Y. H.; Li, J. L.; Chen, R.
(35) Zhao, Y. F.; Li, Z. H.; Li, M. Z.; Liu, J. J.; Liu, X. W.;
Atomic Layer Deposition of Ni on Cu Nanoparticles for Methanol
Waterhouse, G. I. N.; Wang, Y. S.; Zhao, J. Q.; Gao, W.; Zhang, Z. S.;
Long, R.; Zhang, Q. H.; Gu, L.; Liu, X.; Wen, X. D.; Ma, D.; Wu, L. Synthesis from CO2 Hydrogenation. ChemCatChem 2017, 9, 3772.
(52) Tan, Y. C.; Abu Bakar, N. H. H.; Tan, W. L.; Abu Bakar, M.
Z.; Tung, C. H.; Zhang, T. R. Reductive Transformation of Layered-
Double-Hydroxide Nanosheets to Fe-Based Heterostructures for Hydrogenation of Liquid Styrene by Alumina Supported Nickel
Efficient Visible-Light Photocatalytic Hydrogenation of CO. Adv. Catalysts: Comparison between Classical and Non-Classical Methods.
Mater. 2018, 30, 1803127. IOP Conf. Ser.: Mater. Sci. Eng. 2016, 133, 012017.
(36) Chen, G. B.; Gao, R.; Zhao, Y. F.; Li, Z. H.; Waterhouse, G. I. (53) Zhou, M. H.; Tian, L. F.; Niu, L.; Li, C.; Xiao, G. M.; Xiao, R.
N.; Shi, R.; Zhao, J. Q.; Zhang, M. T.; Shang, L.; Sheng, G. Y.; Zhang, Upgrading of liquid fuel from fast pyrolysis of biomass over modified
X. P.; Wen, X. D.; Wu, L. Z.; Tung, C. H.; Zhang, T. R. Alumina- Ni/CNT catalysts. Fuel Process. Technol. 2014, 126, 12.
Supported CoFe Alloy Catalysts Derived from Layered-Double- (54) Zhou, M. H.; Ye, J.; Liu, P.; Xu, J. M.; Jiang, J. C. Water-
Hydroxide Nanosheets for Efficient Photothermal CO2 Hydro- Assisted Selective Hydrodeoxygenation of Guaiacol to Cyclohexanol
genation to Hydrocarbons. Adv. Mater. 2018, 30, 1704663. over Supported Ni and Co Bimetallic Catalysts. ACS Sustainable
(37) Li, Z. H.; Liu, J. J.; Zhao, Y. F.; Waterhouse, G. I. N.; Chen, G. Chem. Eng. 2017, 5, 8824.
B.; Shi, R.; Zhang, X.; Liu, X. W.; Wei, Y. M.; Wen, X. D.; Wu, L. Z.; (55) Zhang, X. H.; Fang, X. Z.; Feng, X. H.; Li, X.; Liu, W. M.; Xu,
Tung, C. H.; Zhang, T. R. Co-Based Catalysts Derived from Layered- X. L.; Zhang, N.; Gao, Z. X.; Wang, X.; Zhou, W. F. Ni/Ln2Zr2O7 (Ln
Double-Hydroxide Nanosheets for the Photothermal Production of = La, Pr, Sm and Y) catalysts for methane steam reforming: the effects
Light Olefins. Adv. Mater. 2018, 30, 1800527. of A site replacement. Catal. Sci. Technol. 2017, 7, 2729.
(38) Fan, H.; Huang, X.; Shang, L.; Cao, Y. T.; Zhao, Y. F.; Wu, L. (56) Ang, M. L.; Oemar, U.; Saw, E. T.; Mo, L.; Kathiraser, Y.; Chia,
Z.; Tung, C. H.; Yin, Y. D.; Zhang, T. R. Controllable Synthesis of B. H.; Kawi, S. Highly Active Ni/xNa/CeO2 Catalyst for the Water−
Ultrathin Transition-Metal Hydroxide Nanosheets and their Extended Gas Shift Reaction: Effect of Sodium on Methane Suppression. ACS
Composite Nanostructures for Enhanced Catalytic Activity in the Catal. 2014, 4, 3237.
Heck Reaction. Angew. Chem., Int. Ed. 2016, 55, 2167. (57) Li, Y. P.; Wen, J.; Ali, A. M.; Duan, M.; Zhu, W.; Zhang, H.;
(39) Younis, M. N.; Malaibari, Z. O.; Ahmad, M. W.; Ahmed, M. S. Chen, C.; Li, Y. Size structure−catalytic performance correlation of
Hydrogen production through steam reforming of diesel over highly supported Ni/MCF-17 catalysts for COx-free hydrogen production.
efficient promoted Ni/γ-Al2O3 catalysts containing lanthanide series Chem. Commun. 2018, 54, 6364.
(La, Ce, Eu, Pr, and Gd) promoters. Energy Fuels 2018, 32, 7054. (58) Isaifan, R. J.; Ntais, S.; Couillard, M.; Baranova, E. A. Size-
(40) Gao, J.; Jiang, Q.; Liu, Y. F.; Liu, W.; Chu, W.; Su, D. S. dependent activity of Pt/yttria-stabilized zirconia catalyst for ethylene
Probing the enhanced catalytic activity of carbon nanotube supported and carbon monoxide oxidation in oxygen-free gas environment. J.
Ni-LaOx hybrids for the CO2 reduction reaction. Nanoscale 2018, 10, Catal. 2015, 324, 32.
14207. (59) Cargnello, M.; Doan-Nguyen, V. V. T.; Gordon, T. R.; Diaz, R.
(41) Bao, F.; Tan, F.; Wang, W.; Qiao, X.; Chen, J. Facile E.; Stach, E. A.; Gorte, R. J.; Fornasiero, P.; Murray, C. B. Control of
preparation of Ag/Ni(OH)2 composites with enhanced catalytic metal nanocrystal size reveals metal-support interface role for ceria
activity for reduction of 4-nitrophenol. RSC Adv. 2017, 7, 14283. catalysts. Science 2013, 341, 771.

2855 DOI: 10.1021/acs.iecr.8b06037


Ind. Eng. Chem. Res. 2019, 58, 2846−2856
Industrial & Engineering Chemistry Research Article

(60) Yang, F. F.; Liu, D.; Zhao, Y. T.; Wang, H.; Han, J. Y.; Ge, Q.
F.; Zhu, X. L. Size Dependence of Vapor Phase Hydrodeoxygenation
of m-Cresol on Ni/SiO2 Catalysts. ACS Catal. 2018, 8, 1672.
(61) Pushkarev, V. V.; An, K.; Alayoglu, S.; Beaumont, S. K.;
Somorjai, G. A. Hydrogenation of benzene and toluene over size
controlled Pt/SBA-15 catalysts: Elucidation of the Pt particle size
effect on reaction kinetics. J. Catal. 2012, 292, 64.
(62) Zhang, G. R.; Zhao, D.; Feng, Y. Y.; Zhang, B. S.; Su, D. S.; Liu,
G.; Xu, B. Q. Catalytic Pt-on-Au nanostructures: why Pt becomes
more active on smaller Au particles. ACS Nano 2012, 6, 2226.

2856 DOI: 10.1021/acs.iecr.8b06037


Ind. Eng. Chem. Res. 2019, 58, 2846−2856

Вам также может понравиться