Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Article

Cite This: Ind. Eng. Chem. Res. 2018, 57, 1506−1515 pubs.acs.org/IECR

Citric Acid-Assisted Synthesis of Nanoparticle Copper Catalyst


Supported on an Oxide System for the Reduction of Furfural to
Furfuryl Alcohol in the Vapor Phase
N. J. Venkatesha*,† and S. Ramesh‡

Chemistry Research Center, Bangalore Institute of Technology, K.R Road V.V Pura, Bengaluru, Karnataka 560004, India

Institute of condensed Matter and NanoscienceMolecules, Solids and Reactivity (IMCN/MOST) Universite catholique de
Louvain, Place Louis Pasteur, 1, Box L4.01.09, 1348 Louvain la-Neuve, Belgium
*
S Supporting Information

ABSTRACT: The simplest method of synthesizing metallic Cu


(111) nanoparticles on the zinc−aluminum mixed oxides was
found to be easily obtained by solid state grinding of all the three
metal nitrates and organic compounds in suitable proportions.
Organic compounds such as citric acid, formic acid, and
hydrazine were used, which act as reducing agents for Cu. The
catalysts were characterized by various physicochemical methods
such as X-ray diffraction, X-ray fluorescence, temperature
programmed reduction, high-resolution transmission electron
spectroscopy, X-ray photoelectron spectroscopy, field emission
scanning electron microscopy and N2-physisorption. The use of
citric acid resulted in Cu(111) nanoparticles with higher surface
area and better dispersion on the surface of mixed oxides. The
catalysts were used for the selective hydrogenation of furfural to
furfuryl alcohol under vapor phase conditions. Among the various catalysts studied, citric acid-assisted catalysis showed better
conversion with higher selectivity of desired product and deactivation free catalytic activity on stream more than 30 h.

1. INTRODUCTION mental pollution because of its high toxic nature and its higher
In the present situation, there is a growing awareness for the temperature with pressurized conditions; these are the main
use of renewable feed stock and the development of new disadvantage of this catalyst.5,6,10 Presently there are several
catalytic processes for the preparation of fuel and fine research groups working on this reaction using various
chemicals. Furfural is one of the biomass derived starting metallics, such as Pt, Pd, Ru, Cu, Ni, Cu, Zn, Ca, Al, Na,
materials obtained from the rice husk and corn core, this was bimetallic, Pd−Cu, Cu−Co, Pt−Sn, and ultrafine amorphous
used for the production of various industrially important alloys Ni−P, Ni−B, Ni−P−B, Co−B, Fe(NiFe)O4, and Ni−
chemicals.1,2 Furfural [FAL] can used to produce furfuryl Fe−B. Cu−Al layered double hydroxide (LDH) and Cu on
alcohol [FOL] as an intermediate to the manufacture of lysine, different supports Cr2O3, Al2O3, SiO2, SBA-15, and CaO11−18
ascorbic acid, numerous lubricants,3,4 tetrahydrofurfuryl alco- used as catalysts for FAL hydrogenation has been reported,19
hol, 2-methylfuran, and tetrahydrofuran. The FOL was The reported catalytic systems involve the preparation of the
produced through the hydrogenation reaction using both catalyst in metal oxide form, and before the reaction occurs, the
vapor and liquid phase reactors. Among the various products, catalyst is to be reduced at harsh conditions with molecular
FOL finds a variety of applications in the chemical industry; it hydrogen for several hours. This process consumes much
was mainly used for the production of various synthetic resins, hydrogen and energy; the former is found to be the future
rubbers, fibers, and chemical resistant thermostatic resins which alternative energy.20−22
are used for strengthening ceramics, molds for metal castings, In the present study a catalyst is prepared by the simplest
impregnating solutions, and carbon binders. It is also used as a method and does not require further reduction before reaction,
solvent for furan resins, pigments, varnishes, and as rocket fuel. making it economically viable for sustainable development. It
Also important, it is an intermediate for the manufacture of shows smaller metallic Cu particles, and their maximum (111)
lysine, vitamin C, lubricants, dispersing agents, and plasti-
cizers.5−9 Received: November 13, 2017
In an industrial process copper chromite has been used Revised: December 22, 2017
commercially for the selective hydrogenation of furfural to Accepted: January 15, 2018
furfuryl alcohol. The presence of Cr6+ causes severe environ- Published: January 15, 2018

© 2018 American Chemical Society 1506 DOI: 10.1021/acs.iecr.7b04701


Ind. Eng. Chem. Res. 2018, 57, 1506−1515
Industrial & Engineering Chemistry Research Article

plane orientation on the surface showed deactivation free that, it was heated from 50 to 800 °C with a rate of 10 °C/min.
catalytic activity of the CZA-based catalysts for the FAL The amount of hydrogen consumed as a function of
hydrogenation to FOL under reduction-free conditions. temperature was monitored by a thermal conductivity detector
Various organic compounds as hydrogen sources such as (TCD) detector.
hydrazine, citric acid, and formic acid were used for in situ The XPS analyses were performed on a SSX 100/206
reduction of Cu, and the the prepared catalysts were photoelectron spectrometer from Surface Science Instruments
characterized by temperature programmed reduction (TPR), (USA) equipped with a monochromatized microfocused Al X-
high-resolution transmission electron spectroscopy (HRTEM), ray source (powered at 20 mA and 10 kV). The samples
scanning transmission electron microscopy (STEM), X-ray powder pressed in small stainless steel troughs of 4 mm
photoelectron spectroscopy (XPS), X-ray fluorescence (XRF), diameter were placed on an aluminum conductive carousel.
Brunauer−Emmett−Teller (BET) surface area analysis, and The pressure in the analysis chamber was around 10−6 Pa, and
powder X-ray diffraction (pXRD). the pass energy was set at 150 eV. The C-(C,H) component of
the C 1s peak of carbon was fixed to 284.8 eV to set the binding
2. EXPERIMENTAL SECTION energy scale. Data treatment was performed with the Casa XPS
2.1. Materials. Furfural and furfuryl alcohol was procured program (Casa Software Ltd., UK), molar fractions were
from Sigma-Aldrich, India. Al(NO3)3·9H2O, Cu (NO3)2·3H2O, calculated using peak areas normalized on the basis of
and Zn(NO3)2·6H2O were purchased from SD fine chemicals. acquisition parameters and sensitivity factors provided by the
Hydrazine hydrate, citric acid, and formic acid were obtained manufacturer.
from Merck, India, and used without further purification. TEM images of the fresh and used catalysts were obtained by
2.2. Catalyst Preparation. The catalysts were prepared by using FEI TITAN3 80−300 kV aberration corrected trans-
taking 50 wt % of Cu(NO3)2, 20 wt % of Zn(NO3)2, and 30 wt mission electron microscope having both HRTEM and STEM-
% of Al(NO3)2. To this mixture, 70 wt % organic compounds Z contrast imaging facility.
was added and the mixture was ground well using mortar and Dispersion of metallic Cu atoms were determined by pulsed
pestle. The sample was then dried at 120 °C for 3 h and titration of N2O using BELCAT-2 TPX instrument Japan, the
calcined under inert gas atmosphere for 4 h at 300 °C (with a catalysts of 200 mg were placed in the quartz tube (reduced
ramp rate of 5°/min). After that, the sample was cooled to form). Helium gas was used as carrier gas at 10 mL min−1 and
room temperature by maintaining the N2 flow at 20 mL per the successive doses of 10% N2O/He gas were subsequently
min. The samples were stored in vacuum desiccator in airtight titered into the He stream by means of a calibrated injection
bottles. The samples were designated as CZA, CZAC, CZAH, valve (15 μL N2O pulse) at 60 °C temperature. The product
and CZAF for samples without organic compound, and with gases N2 and retained N2O were analyzed by TCD. The
citric acid, hydrazine hydrate, and formic acid, respectively. The amount of N2 and N2O was calibrated with the ZA sample,
CZA sample was reduced using H2 at 300 °C for 4 h and N2O consumption and N2 liberated were calculated, both
designated as CZAD. giving good agreement values within 1% error.
2.3. Catalyst Characterization. The prepared CZA Bulk Cu elemental analysis was carried out using a Horiba
samples using organic compounds were characterized using XGT X-ray fluorescence instrument. A Shimadzu IRAffinity-1
various techniques, such as XRD, BET, field emission scanning FT-IR instrument was used to record the spectra of the samples
electron microscopy (FE-SEM), TPR, XRF, Fourier transform by the KBr pellet method with 4.0 cm−1 resolution in the
infrared (FT-IR), HRTEM, STEM and XPS. wavenumber range 4000−500 cm−1 to study the presence of
Powder XRD measurements were performed for fresh and carbonaceous compounds in the catalysts after calcination
spent catalysts, the 2θ values were between 10−80° obtained by under N2 atmosphere. A complex formation was confirmed.
scanning at 2θ = 0.020° per step using a PAN alytical X’pert 2.4. Catalytic Activity Test. The vapor phase hydro-
PRO X-ray diffractometer with graphite monochromatized Cu genation of furfural was carried out in a fixed-bed down flow
Kα radiation source (λ = 1.5406 Å). The Debye−Scherrer quartz reactor (400 mm long and 20 mm inner diameter)
equation was used to estimate the mean Cu crystallite size under atmospheric pressure. The catalyst of 1.5 g was packed at
based on the diffraction peaks of the Cu (111) plane. the center of the reactor between the plugs of glass wool and
Surface and pore characteristics of the catalysts were the reactor was set to the desired reaction temperature (160−
measured using a Quanta Chrome Nova-1000 surface analyzer 280 °C). Furfural (LHSV 3.6 s−1) 1.8 mL/h was fed
under liquid nitrogen temperature. Prior to the measurements continuously from the syringe pump (model NE-1000) to a
sample were degassed at 300 °C for 4 h to remove preadsorbed vaporizer, the vapors formed were mixed with H2 gas with a
gas and moisture. The BET method was employed to measure ratio 15:1 of H2/furfural. This mixed gaseous stream was passed
the quantity of nitrogen absorbed, and the cumulative volumes through the reactor. The products were condensed in an ice-
of pores were obtained by the BJH method from the desorption cooled trap and analyzed by a chemito GC-1000 gas
isotherms. chromatograph with TR-WAX capillary column (30 m length,
Morphology of the samples was recorded using a JEOL JSM- 0.32 mm thickness, and 0.5 μm internal diameter) attached
6490 FESEM under high vacuum mode. with flame ionization detector. The calibration and quantifica-
The reducibility of the sample was determined using a tion of reactant (FAL) conversion and products (FOL, MF)
BELCAT-II Microtrac (made in Japan) temperature-pro- yield was made using standard samples purchased from Sigma
grammed reduction analyzer. The measurements were Aldrich, India.
performed by placing 50 mg of catalyst sample to the quartz
tube and ramping the temperature with a rate 10 °C/min, 3. RESULTS AND DDISCUSSION
under 20 mL/min of Ar up to 500 °C, and holding for 1 h. 3.1. Catalyst Characteristics. 3.1.1. Surface Character-
Then the sample was cooled to room temperature and a stream istics. Surface area and pore characteristics of CZA samples
of 10% H2/Ar was flushed with a flow rate of 20 mL/min. After were provided in Table1. The CZA sample prepared without
1507 DOI: 10.1021/acs.iecr.7b04701
Ind. Eng. Chem. Res. 2018, 57, 1506−1515
Industrial & Engineering Chemistry Research Article

Table 1. Surface Characteristics of the Prepared Catalysts 3.1.2. XRD Analysis. Powder X-ray diffraction patterns for all
the prepared catalysts are shown in Figure 2. The CZA catalyst
surface pore pore average TPR-H2
area volume diameter particle size consumption
catalyst (m2/g) (cm3/g) (Å) (nm) (mmol/g)
CZA 28.3 0.049 66 28 1.86
CZAC 48.3 0.083 54 14 0.13
CZAF 36.0 0.051 63 20 0.32
CZAH 38.0 0.061 62 20 0.45

using organic compounds as reducing agent showed the least


surface area (28 m2/g), pore volume (0.049 cm3/g), and more
average pore diameter. As usual, copper oxide particles are not
uniformly distributed on mixed oxides and resulted in low
surface area. This confirms that the particle size is bigger than
others. The citric acid CZAC sample showed the highest
surface area and a pore volume with the least average pore
diameter and particle size, due to the smaller particle size. As
expected, the Cu species is much smaller compared to its oxide
form. The higher surface area of the sample is caused by the
complexation of citric acid with metal ions and upon calcination Figure 2. XRD patterns of (a) CZA, (b) CZAC, (c) CZAF, and (d)
under N2 atmosphere decomplexation leading to the generation CZAH.
of metallic species, whereas the formic acid and hydrazine
samples showed lower surface area and pore volume with showed that diffraction peaks at 2θ value 32.48 (110), 35.38
higher average pore diameter and agglomeration of particles (002), 35.53 (1̅11), 38.64 (111), 38.97 (200), and 46.85 (2̅02)
leading to larger size. are attributed to the CuO form with monoclinic symmetry
The complex formation between the Cu2+ ions and citric was (JCPDS 80-0076) with ZnO, Al2O3 and no peaks observed for
confirmed by the FT-IR spectrum. Figure 1 showed that spectra metallic Cu. The organic compound samples showed major
diffraction peaks at 43.32 (111), 50.45 (200), and 74.13 (220)
that corresponded to metallic copper and possess face-centered
lattice with cubic symmetry (JCPDS 04-0836). Also very
minute peaks at 36.50 (111), 42.42 (200) and 61.55 (220) were
present for Cu2O having a primitive lattice with cubic symmetry
(JCPDS 77-0199).20,23 This information evidently proves that
in the nitrogen atmosphere the Cu2+ species converts to the Cu
species in the presence of hydrogen produced by the
decomposition of organic compounds (citric acid, formic
acid, and hydrazine) during calcination.
3.1.3. HRTEM Analysis. Figure 3 shows transmission electron
micrographs of the prepared catalysts. Alumina was hardly
detected in all the catalytic samples and mainly found in the
amorphous phase. Various morphologies were detected for the
CZA catalysts, such as round and irregular shaped Cu particles.
The catalyst prepared without reducing agent (Figure 3A)
resulted in bigger particle size and irregular shape of copper and
zinc-alumina oxides clusters, whereas the CZAC catalyst
Figure 1. FT-IR spectra of complex, citric acid, and CZAC samples. exhibited a marked interface between Cu and zinc−alumina
oxide particles. The Cu particles were in spherical shape, and
surfaces of the metal particles typically covered the smaller
of citric acid, which has three main bands, one main intense ZnO. Additionally, homogeneous, smaller Cu particles of 9 nm
band between 1799 to 1649 cm−1 for the CO stretching, a were observed. The latter samples (CZAF and CZAH)
broader band at 3649−3084 cm−1 for the acid O−H (both acid exhibited inhomogeneous particle shapes which tended to
and alcohol) stretching with hydrogen bonding, the band agglomerate and form indistinct structures, which may be
between 1479 and 1385 cm−1 for the C−O stretching. The related to the structure of the organic compound used.
spectra of the Cu−citric acid complex showed a shift of the 3.1.4. TPR Analysis. The reduction profiles of all prepared
CO and C−O stretching band at 1608 and 1423 cm−1 catalysts are shown in Figure 4 and consumption hydrogen is
respectively, with a decrease in intensity at 3387 cm−1 being provided in Table 2. The TPR profile of CZA consists of one
due to a decrease in the bond strength of O−H (acid) broad and intense peak between the temperature range of 160
stretching by complex formation with the Cu metal ion. The to 300 °C; the deconvolution can bifurcate into two peaks, one
spectra also show one small sharp band at 3547 cm−1 from the centered at 230 °C and another at 265 °C. As per the literature
OH (alcohol) of citric acid with the absence of hydrogen survey,24−28 the CuO particles reduction temperature will vary
bonding. The spectra of CZAC after the calcination at 300 °C and depend on the size, dispersion, and crystallinity. The CZA
under N2 showed the absence of carbonaceous moieties in the sample has bigger crystalline particles of CuO as evident by the
sample, which is in good agreement with the CHNS analysis. TEM and SEM images, and it consumed 3.25 mmol g−1 of H2.
1508 DOI: 10.1021/acs.iecr.7b04701
Ind. Eng. Chem. Res. 2018, 57, 1506−1515
Industrial & Engineering Chemistry Research Article

Figure 3. HRTEM images of CZA samples: (A) CZA, (B) CZAC, (C) CZAF, and (D) CZAH.

particle size being not well dispersed, when compared to the


amount needed when the catalyst is prepared without any
reducing agent. This clearly demonstrates the role of reducing
agent during the catalyst preparation.
3.1.5. STEM and SAED Analysis. The particle size
distribution and microstructures of Cu nanoparticles of CZAs
samples were further confirmed by STEM images (Figure 5).
All the samples at higher resolution showed fringe images, and
from that we calculated d(hkl)-spacing values for Cu species
which are in the fcc lattice form. These d-values were calculated
Figure 4. TPR profiles of (a) CZA, (b) CZAC, (c) CZAF, and (d)
by using digital Micrograph 3 software. The calculated values
CZAH.
were between two lines from three different organic
compounds. The CZAs sample showed similar patterns with
Table 2. Metal Dispersion by Pulsed Titration of N2Oa
the values of d(111) = 0.205 nm and d(200) = 0.179 nm. The CZA
amount of surface Cu in sample prepared without organic compound has the majority of
N2 O Cu atoms area of Cu samples CuO species, and it showed the values of d(111) = 0.245 nm
consumed on surface atoms (XRF) % Cu
catalysts (μmol/g) (mmol/g) (m2/g) (mmol/g) dispersion majorly with d(200) = 0.232 nm. Very small amounts of Cu2O
CZA 30.54 0.284 11.71 1.58 18 species are also observed with d(111) = 0.245 nm and d(200) =
CZAC 94.22 0.876 36.14 1.51 58 0.213 nm being in good agreement with the diffraction peaks of
CZAF 64.05 0.597 24.57 1.53 39 all the samples.29 SAED patterns of CZAC, CZAH, and CZAF
CZAH 69.92 0.651 26.82 1.48 44 samples showed the presence of Cu metallic phases and the
a formation of different planes (002), (11̅ 1), (200), and (111).
Pulsed chemisorption of N2O by TPX instrument, bulk Cu analysis
by XRF. Figure 5 and the CZA sample showed the CuO phase with
different planes (002), (111), (200), (220), and (311) in the
SAED patterns; these are also in good agreement with the XRD
The catalysts prepared using organic compounds as reducing and STEM analyses.
agents showed no significant TPR reduction peaks, which 3.1.6. SEM and EDXA. The morphology of the CZA samples
clearly indicated that copper species are reduced to metallic and distribution of Cu species on the surface was studied by
form during the catalyst preparation step. All the reducing elemental mapping. The scanning electron micrograph of the
agents used to convert Cu2+ to Cu are effective for the samples show that the metal oxides are mixed well throughout
reduction. Citric acid, formic acid, and hydrazine first resulted the entire catalyst without forming any separated clusters. The
in metal complex formation and during calcinations at 300 °C EDXA profile shows that the catalyst surface is composed of
under inert atmosphere the metal complex decomposes. Cu, Zn, and Al species. The SEM images of CZA (Figure 6)
Decomposition results in hydrogen generation, which converts showed an agglomeration of Cu−Zn clusters over Al2O3. The
Cu2+ species into metallic Cu species by in situ reduction. TPR distribution of metal clusters is consistent throughout the
profiles of these samples showed that the peak at 175 °C surface of Al2O3. The CZAC catalyst almost resulted in a
indicates these samples require a lower temperature and a uniform plate-like structure. CZAH and CZAF showed smaller
negligible amount of hydrogen for the reduction of Cu2O Cu−Zn clusters and better dispersion of Cu−Zn clusters over
particles. The particles are present in negligible amounts and Al2O3.
also are smaller and are highly dispersed, while in the case of 3.1.7. XPS Analysis. The XPS technique gives insights into
the formic acid sample, a peak is seen at a slightly higher the oxidation state of copper species on the surface of the
intensity, confirming that formic acid has less of a tendency to catalysts prepared with and without reducing agents. The Cu 2p
reduce the CuO species. Some of the Cu2O species remain in photoelectron spectra of prepared catalysts are shown in Figure
the sample, and these species are less dispersed and are bigger 7. The figure gives the information about the oxidation state of
with irregular shape, needing a greater amount of hydrogen surface Cu species in CZA, CZAC, CZAF, and CZAH samples.
(0.34 mmol g−1) for a complete reduction to get the Cu The catalyst prepared without organic compound (CZA)
species. The hydrazine, being a strong reducing agent results in showed a broad peak centered at 933.99 eV for the Cu(2+)O
metallic copper even at room temperature, but the particle size and peaks in the region between 940.5 and 944.5 and 962 eV
of the ground mixture was more comparable to that of the correspond to satellite shakeup peaks that show the presence of
CZAC. In all the cases, the amount of hydrogen consumed for CuO species in the samples. Whereas the catalysts prepared
the reduction is greater (0.47 mmol g−1), due to the higher with organic reducing agents shows a sharp peak at 932.4
1509 DOI: 10.1021/acs.iecr.7b04701
Ind. Eng. Chem. Res. 2018, 57, 1506−1515
Industrial & Engineering Chemistry Research Article

Figure 5. STEM images of CZA and CZAC and SAED patterns of CZAC.

Figure 6. SEM images of (A) CZA, (B) CZAC, (C) CZAF, (D) CZAH, and (E) EDXA pattern of CZAC.

representing together with the absence of typical satellite peaks of small amount of unreduced species (Cu2O). The formic acid
at 940−945 eV for the Cu2+ species related to the unoccupied used may not be sufficient enough to reduce all the copper
3d states, the reduced copper species in the form of Cu and/or present in the catalyst species (it will produce the least amount
Cu+. CZAC samples give only a single sharp peak at 932.34 eV of hydrogen during calcination among the reducing agents
and the absence of a satellite peak confirms the presence of used). Hydrazine being a strong reducing agent, still resulted in
only Cu metallic species. However, CZAF samples showed an three kinds of copper being present after the calcination under
additional peak along with Cu species. There is a small peak inert atmosphere. The sample CZAH showed a broad peak
which is due to (deconvoluted peak at 933.34 eV) the presence which is in the range that falls to all three kinds of Cu species,
1510 DOI: 10.1021/acs.iecr.7b04701
Ind. Eng. Chem. Res. 2018, 57, 1506−1515
Industrial & Engineering Chemistry Research Article

followed by the addition of a second hydrogen on the oxygen of


alkoxide to form an alcohol.35 Furfuryl alcohol upon
dehydration leads to the formation of a cation, in which a
further reduction results in methyl furan. Furfural hydro-
genation is more active in the presence of Cu metallic species
especially on the (111) plane.36,37 The exposed species in the
catalyst has weak interaction with elongation of the CO bond
rather than the furanic ring of FAL. Furfural tends to adsorb on
the catalyst surface through carbonyl oxygen in a η1 (O)-
aldehyde configuration as previously reported in the
literature.38−41 DFT calculations, reported by earlier workers,
predict a strong repulsion of the Cu (111) surface to the furan
ring. This is presumably due to the 3d orbital of the surface Cu
atoms and the antibonding orbital of the aromatic furan
ring.12,36,42−45
Catalytic activity was studied by varying the reaction
Figure 7. XPS spectra of catalyst samples. parameters such as effect of the reaction temperature, time
on stream, and furfural flow for vapor phase hydrogenation of
but by the consideration of area under the curve upon furfural to furfuryl alcohol over CZAC catalyst. These have
deconvolution, it is confirmed that there three peaks which been discussed in the preceding sections.
have the ratio of area of the peaks at binding energies. The 3.2.1. On Stream Stability Behavior. Time on stream
peaks are observed at 932.34, 933.34, and 933.99 eV with the behavior was studied for 30 h. Initially (first 6 h) 78% FAL
ratio of 72:19:9 corresponding to the Cu, Cu2O, and CuO, conversion with 92% and 8% selectivity toward FOL and MF
respectively.30−32 XPS results confirm that all the reducing respectively, was observed. After 6 h, the conversion of FAL
agents used resulted in metallic copper either in the preparation slightly decreased to 74% with a slight increase in selectivity for
step or during calcination; however the degree of reducing FOL (97%) with decreased MF selectivity (3%). This confirms
capacity depends on the strength and number of hydrogen deactivation free stable catalytic activity up to 30 h. A further
produced during calcination. Citric acid resulted exclusively in increase in stream time for more than 30 h showed a slight
metallic copper, whereas other reducing agents result in a decreased of FAL (8%) conversion, and the selectivity of FOL
mixture of species with the majority being Cu, merging well reached almost to 100%. A typical time on stream activity of
with the reduction profiles obtained by TPR and XRD. From CZAC catalyst was shown in Figure 8. It exhibited a steady and
these observations Cu species only existed for CZAC samples, stable activity and selectivity (MF was negligible and not
and CZAF and CZAH samples exhibited all three kinds of Cu considered for studies).
species, but the ratio of Cu2O and CuO species was very less
compared to the Cu, which indicated that these results are in
good agreement with STEM and XRD results.
3.1.8. Cu Dispersion by N2O Chemisorption. Oxygen
chemisorption via N2O decomposition on CZAs catalyst
samples were considered to determine the surface Cu atoms
on the catalysts. Initially, N2O decomposition on CZAs were
confirmed. Table 2 shows the results of N2O consumption and
dispersion and surface area of metallic Cu. These measurement
for CZAs samples were carried at 60 °C by sequential injection
of N2O and calculated by following the previous report, by
considering the Cu density 14.9 × 1019 atoms per unit surface
area.33,34 The pulsed titration results of CZAs samples showed
N2O consumption in the following order CZA < CZAH <
CZAF < CZAC. These results directly evident that the
presence of metallic Cu on the zinc−alumina mixed oxide
cluster. Among the CZAs samples CZAC showed more amount
of N2O consumption, it showed that, there is more number of Figure 8. Time on stream behavior on CZAC for FAL conversion.
Cu (58%) atoms on the surface, whereas CZA has only 11%, Reaction conditions: temperature, 200 °C; LHSV = 3.6 h−1; H2/FAL
CZAF 39%, and CZAH 44% of Cu atoms on the surface. This = 3; atmospheric pressure.
is evident that citric showed more dispersion than formic acid
and hydrazine. 3.2.2. Effect of Reaction Temperature. The temperature has
3.2. Catalytic Activity. FAL hydrogenation was performed significant effect on vapor phase hydrogenation of FAL over the
under vapor phase conditions for all the prepared catalysts, and CZAC sample. Figure 9 shows the conversion of FAL and FOL
it was found that CZAC showed a maximum conversion of 76% selectivity as a function of temperature range between 160−280
with a FOL selectivity of 96% and 4% MF. The reaction °C. Increased conversion with the increase in temperature was
pathway of furfural hydrogenation follows the addition of observed. Maximum conversion (84%) and selectivity of FAL
activated molecular hydrogen to the carbonyl (aldehyde) group (98%) was observed at 200 °C. The selectivity toward the FOL
of the FAL. Initially the addition of hydrogen to the carbonyl reduces with increase in reaction temperature on account of
carbon leads to the formation of an alkoxide intermediate formation of the double hydrogenated product methyl furan
1511 DOI: 10.1021/acs.iecr.7b04701
Ind. Eng. Chem. Res. 2018, 57, 1506−1515
Industrial & Engineering Chemistry Research Article

3.6 h−1 was considered as the optimal flow of FAL to get


maximum conversion, and more important industrial needed
product FOL, and minimize the side product MF.
3.3. Comparison of CZA’s Catalyst Samples for
Hydrogenation Reaction. Under optimized reaction con-
ditions, at 200 °C, H2/FAL ratio = 3 and LHSV is 3.6 h−1. The
comparison between the different catalysts was studied. The
copper species in the CZA catalyst were reduced using
molecular hydrogen (CZAD), and the activity results were
given in Table 2. Among the catalysts studied, the CZA sample
showed the least activity of 34% FAL conversion with 23 and
11% yields of FOL and MF, respectively. The CZAD sample
showed 80% conversion with yields of 65% FOL, 10% MF, and
5% furan, whereas, CZAC showed maximum conversion (88%)
with a high yield of FOL (81%) and a very less amount of MF
7%, other catalysts CZAF and CZAH were shown to increase
conversion and yield but still less than the CZAC (Table 3). At
a reaction temperature of 240 °C, all the catalysts showed
Figure 9. Effect of temperature on hydrogenation of FAL over CZAC. slightly increased conversion with varied product yield
Reaction conditions: atmospheric pressure; H2/FAL ratio = 3; LHSV distribution. At higher temperature, the yield of MF was
= 3.6 h−1; time = 4 h. increased and also a minor amount of furan was also observed.
This provides information about the cause of increased
(MF) as a side-product. The formation of MF is due to temperature for the C−C bond breaking in the MF leading
hydrogenalysis of the alcoholic group at higher reaction to a more stabilized furan formation. The conversion and yield
temperature. Thus, 240 °C seems to be an optimum of the more important product FOL was selectively formed
temperature for FAL conversion in yielding higher amounts more in CZAC, because it showed a maximum of Cu species
of FOL with a negligible amount of MF.
with active (111) plane exposed on the surface with lower
3.2.3. Effect of Furfural Flow. FAL conversion and FOL
particle size, whereas CZAD, CZAF, and CZAH showed
selectivity were measured as a function of FAL flow rate
increased particle size, and in addition to Cu (111) species,
(LHSV) from 2.36 to 9.86 h−1 under the optimized conditions
there is some amounts of Cu+ species that can also observed,
of 200 °C with a H2/FAL ratio of 3. The results are shown in
Figure 10. The reaction was performed using CZAC catalyst leading to decreased conversion and yield of desired product.
The CZA catalysts showed the least conversion and product
because it contains only Cu2O in the presence of molecular
hydrogen during the reaction, showing very less activity. Also
particles are agglomerated leading to higher particle size.
3.4. Correlation between the Surface Cu Species and
FOL Yield. The Cu species has a main role in conversion,
because metallic Cu species are responsible for the reduction
reaction. The more dispersed Cu species becomes available for
the reactant molecule adsorption leading to reduction of the
carbonyl group to the alcohol group. The catalysts synthesis
using organic compounds showed improved hydrogenation,
because of the complex formation with the organic compound,
and are dispersed well on the surface because this Cu species
surface area will be higher. CZA prepared without using the
organic compound has very less dispersion as evident by the
N2O titration consequently surface area of the Cu species is
lesser and also availability of the Cu species on the surface is
too less so that. The new factor called efficiency of surface H2
occupancy was found to be useful for the prediction of product
yield and conversion of reactant. So to check the effect of the
Figure 10. Conversion of FAL and selectivity of FOL at various FAL Cu particle size, surface area, and dispersion responsibility on
flow rates over CZAC for FAL hydrogenation. Reaction conditions: the yield, correlations were made on the FOL yield with
temperature, 200 °C; H2/FAL ratio = 3; atmospheric pressure, time = individual factors. Table 4 shows the good correlation for all the
4 h. factors: particle size (0.96), Cu surface area (0.94), and
dispersion (0.96). An increase or decrease in any one the
sample. A very low LHSV (2.4 h−1) favors the formation of MF factors will directly effect the product and conversion. The
and causes a decrease in the selectivity toward FOL. At an factor called efficiency surface hydrogen occupancy (ESHO),
appropriate LHSV, 3.6 h−1 and more than that, the conversion the product of surface Cu area, particle size, and dispersion, was
of FAL decreased with an increase in selectivity toward FOL correlated with FOL yield to determine the combined influence
with decreased selectivity of MF. This is attributed to reduced of all the three factors. It showed a better correlation coefficient
contact time between the catalyst and reactant. Hence LHSV (0.99) with lower particle size and better dispersion of Cu
1512 DOI: 10.1021/acs.iecr.7b04701
Ind. Eng. Chem. Res. 2018, 57, 1506−1515
Industrial & Engineering Chemistry Research Article

Table 3. Conversion and Yield of Products under Different Catalysts for Hydrogenation of FALa
200 °C 240 °C
CZA CZAD CZAC CZAF CZAH CZA CZAD CZAC CZAF CZAH
FAL conversion (%) 34 80 88 78 75 38 86 91 84 83
yield (%) FOL 23 65 83 70 69 20 62 77 70 68
MF 11 10 05 08 06 08 16 11 12 10
furan 05 08 03 03 05
a
Reaction conditions: H2/FAL ratio = 3; atmospheric pressure, LHSV, 3.6 h−1.

Table 4. Correlation between the Different Factors with FAL Conversion and FOL Yield
catalysts surface area of Cu atoms (m2/g) average particle size (nm) % Cu dispersion ESHOa (103) FAL conversion FOL yield TOF (h−1)
CZA 11.71 28 18 5.90 34 23 14.6
CZAC 36.14 14 58 29.34 88 83 32.4
CZAF 24.57 20 39 21.50 78 70 26.3
CZAH 26.82 20 44 24.78 78 69 25.8
correlation r 0.98 0.94 0.96 0.999
a
ESHO: Efficiency surface H2 occupancy

particle influence on a higher surface area, and showed better 16.6 nm, which confirms deactivation free activity for more
catalytic activity for the FAL hydrogenation to yield FOL. than 30 h of reactants stream.
3.5. Characteristics of Used Catalysts. The used catalyst
sample was characterized to know the deactivation and stability 4. CONCLUSION
using XRD, N2-physisorption, and HRTEM. The results
showed stable and deactivation free catalytic activity. The The solid state dry synthesis of metallic copper on mixed oxides
XRD pattern (Figure 11) showed major peaks at 43.317 (111), using organic compound as reducing agents was reported. The
metal complexes decomposed during the calcination, produces
hydrogen which results in situ reduction of copper on the oxide
cluster surface. The structure and composition of organic the
compound used play an important role in morphology and
reduction behavior of the final catalyst. Citric acid-assisted
synthesis (CZAC) resulted in better dispersion of metallic
copper nanoparticles on mixed oxides with higher surface area.
The CZAC catalyst showed the best performance more than 30
h on-stream for the catalytic hydrogenation of furfural to
furfuryl alcohol with higher selectivity. The term efficiency
surface hydrogen occupancy (ESHO) was coined to correlate
the effect of metallic dispersion, particle size, and their surface
Figure 11. XRD pattern of (CZACR) used catalysts. area for the hydrogenation reaction; it showed a better
correlation coefficient. This can be extended to industrial
50.449 (200), and 74.126 (220) responsible for the Cu metallic catalysis. This method of synthesis showed to be less energetic
species and minor peaks for the CuO and Cu2O, which and more economic for industrial catalyst synthesis.


indicates that the catalyst retained almost all its characteristics
after use. The surface and pore characteristics measured by the ASSOCIATED CONTENT
BET method showed a slightly decreased surface area of 43 m2/
g with a pore volume of 0.076 cm3/g and average pore diameter * Supporting Information
S

of 53.12 Å. The HRTEM images of CZAC (Figure 12) showed The Supporting Information is available free of charge on the
some agglomeration of particles with increased particle size of ACS Publications website at DOI: 10.1021/acs.iecr.7b04701.

Figure 12. HRTEM images of (A) CZAC and (B) CZACR catalysts.

1513 DOI: 10.1021/acs.iecr.7b04701


Ind. Eng. Chem. Res. 2018, 57, 1506−1515
Industrial & Engineering Chemistry Research Article

Comparison of different catalysts with our catalyst, UV− (13) Sitthisa, S.; Sooknoi, T.; Ma, Y.; Balbuena, P. B.; Resasco, D. E.
visible spectra of CZAs catalysts for the confirmation of Kinetics and mechanism of hydrogenation of furfural on Cu/SiO2
catalysts. J. Catal. 2011, 277, 1−13.
complex formation and reaction mechanism (PDF) (14) Sitthisa, S.; Resasco, D. E. Hydrodeoxygenation of Furfural Over

■ AUTHOR INFORMATION
Corresponding Author
Supported Metal Catalysts: A Comparative Study of Cu, Pd and Ni.
Catal. Lett. 2011, 141, 784−791.
(15) Wu, J.; Shen, Y.; Liu, C.; Wang, H.; Geng, C.; Zhang, Z. Vapor
phase hydrogenation of furfural to furfuryl alcohol over environ-
*E-mail: venkatesha.312@gmail.com. Tel.: 9901003120. mentally friendly Cu-Ca/SiO2 catalyst. Catal. Commun. 2005, 6, 633−
ORCID 637.
(16) Chen, X.; Li, H.; Luo, H.; Qiao, M. Liquid phase hydrogenation
N. J. Venkatesha: 0000-0002-9091-9744 of furfural to furfuryl alcohol over Mo-doped Co-B amorphous alloy
Notes catalysts. Appl. Catal., A 2002, 233, 13−20.
The authors declare no competing financial interest. (17) Antunes, M. M.; Lima, S.; Neves, P.; Magalhaes, A. L.; Fazio, E.;


Fernandes, A.; Neri, F.; Silva, C. M.; Rocha, S. M.; Ribeiro, M. F.;
ACKNOWLEDGMENTS Pillinger, M.; Urakawa, A.; Valente, A. A. One-pot conversion of
furfural to useful bio-products in the presence of a Sn,Al-containing
The authors would like to thank the Rajya Vokkaligara Sangha zeolite beta catalyst prepared via post-synthesis routes. J. Catal. 2015,
and Principal of Bangalore Institute of Technology for the 329, 522−537.
facilities provided. The authors also extended their thanks to (18) Halilu, A.; Ali, T. H.; Atta, A. Y.; Sudarsanam, P.; Bhargava, S.
Prof. S. M. Shivaprasad, ICMS JNCASR, Mr. G. Vijaykumar K.; Abd Hamid, S. B. Highly Selective Hydrogenation of Biomass-
and Mr. Eranjaneya, Central College BUB, for XRD, Mr. Derived Furfural into Furfuryl Alcohol Using a Novel Magnetic
Srikanth, NCL Pune, for TPR measurements and Mr. M. Nanoparticles Catalyst. Energy Fuels 2016, 30, 2216−2226.
Shivakumar, BMSCE, for SEM. (19) Lee, S. P.; Chen, Y. W. Selective Hydrogenation of Furfural on


Ni−P, Ni−B, and Ni−P−B Ultrafine Materials. Ind. Eng. Chem. Res.
1999, 38, 2548−2556.
REFERENCES (20) Ramesh, S.; Venkatesha, N. J. Template Free Synthesis of Ni-
(1) Adam, J.; Blazso, M.; Meszaros, E.; Stocker, M.; Nilsen, M. H.; Perovskite: An Efficient Catalyst for Hydrogen Production by Steam
Bouzga, A.; Hustad, J. E.; Gronli, M.; Oye, G. Pyrolysis of biomass in Reforming of Bioglycerol. ACS Sustainable Chem. Eng. 2017, 5, 1339−
the presence of Al-MCM-41 type catalysts. Fuel 2005, 84, 1494−1502. 1346.
(2) He, R.; Ye, X. P.; English, B. C.; Satrio, J. A. Influence of pyrolysis (21) Yan, K.; Liao, J.; Wu, X.; Xie, X. A noble-metal free Cu-catalyst
condition on switchgrass bio-oil yield and physicochemical properties. derived from hydrotalcite for highly efficient hydrogenation of
Bioresour. Technol. 2009, 100, 5305−5311. biomass-derived furfural and levulinic acid. RSC Adv. 2013, 3,
(3) Luo, H. S.; Li, H. I.; Zhuang, L. Furfural Hydrogenation to 3853−3856.
Furfuryl Alcohol over a Novel Ni−Co−B Amorphous Alloy Catalyst. (22) Yfanti, V. L.; Vasiliadou, E. S.; Lemonidou, A. A. Glycerol
Chem. Lett. 2001, 30, 404−405. hydro-deoxygenation aided by in situ H2 generataion via methanol
(4) Audemar, M.; Ciotonea, C.; Vigier, K. O.; Royer, S.; Ungureanu, aqueous reforming over a Cu-ZnO-Al2O3 catalysts. Catal. Sci. Technol.
A.; Dragoi, B.; Dumitriu, E.; Jerome, F. Selective Hydrogenation of 2016, 6, 5415−5426.
Furfural to Furfuryl Alcohol in the Presence of a Recyclable Cobalt/ (23) Graciani, J.; Mudiyanselage, K.; Xu, F.; Baber, A. E.; Evans, J.;
SBA-15 Catalyst. ChemSusChem 2015, 8, 1885−1891. Senanayake, S. D.; Stacchiola, D. J.; Liu, P.; Hrbek, J.; Sanz, J. F.;
(5) Kijenski, J.; Winiarek, P.; Paryjczak, T.; Lewicki, A.; Mikolajska, Rodriguez, J. A. Highly active copper-ceria and copper-ceria-titania
A. Platinum deposited on monolayer supports in selective hydro- catalysts for methanol synthesis from CO2. Science 2014, 345, 546−
genation of furfural to furfuryl alcohol. Appl. Catal., A 2002, 233, 171− 551.
182. (24) Petitjean, L.; Gagne, R.; Beach, E. S.; Xiao, D.; Anastas, P. T.
(6) Rao, R. S.; Baker, R. T. K.; Vannice, M. Furfural hydrogenation
Highly selective hydrogenation and hydrogenolysis using a copper-
over carbon-supported copper. Catal. Lett. 1999, 60, 51−57.
doped porous metal oxide catalyst. Green Chem. 2016, 18, 150−156.
(7) Thomas, D. Catalyst comprising Raney nickel with adsorbed
(25) Hu, Q.; Yang, L.; Fan, G.; Li, F. Hydrogenation of biomass-
molybdenum compound. U.S. Patent No. 4153578, 1979.
derived compounds containing a carbonyl group over a copper-based
(8) Leo, F. J.; Herman, F. J. Copper chromite catalyst for preparation
of furfuryl alcohol from furfural.U.S. Patent No. 4251396, 1981. nanocatalyst: Insight into the origin and influence of surface oxygen
(9) Elizaveta, P. A.; Juldash, M.; Ivan, P. P.; Lidia, G. G.; vacancies. J. Catal. 2016, 340, 184−195.
Jormukhamat, A. G.; Ildgam, B. A.; Aexei, D. M.; Larisa, C. J.; (26) Chanquia, C. M.; Sapag, K.; Rodriguez-Castellon, E.; Herrero, E.
Lazar, P. D.; Bassheva, B. B. Method for preparing furfuryl alcohol. R.; Eimer, G. A. Nature and Location of Copper Nanospecies in
U.S. Patent No. 4261905, 1981. Mesoporous Molecular Sieves. J. Phys. Chem. C 2010, 114, 1481−
(10) Rao, R.; Dandekar, A.; Baker, R. T. K.; Vannice, M. A. 1490.
Properties of Copper Chromite catalysts in hydrogenation reaction. J. (27) Lou, L. L.; Liu, S. X. CuO-containing MCM-48 as catalysts for
Catal. 1997, 171, 406−419. phenol hydroxylation. Catal. Commun. 2005, 6, 762−765.
(11) (a) Liu, H.; Hu, Q.; Fan, G.; Yang, L.; Li, F. Surface synergistic (28) Hao, X. Y. A novel approach to prepare MCM-41 supported
effect in well-dispersed Cu/MgO catalysts for highly efficient vapor- CuO catalyst with high metal loading and dispersion. Microporous
phase hydrogenation of carbonyl compounds. Catal. Sci. Technol. Mesoporous Mater. 2006, 88, 38−47.
2015, 5, 3960−3969. (b) Lesiak, H.; Binczarski, M.; Karski, S.; (29) Niu, X.; Zhao, T.; Yuan, F.; Zhu, Y. Preparation of hollow
Maniukiewicz, W.; Rogowski, J.; Szubiakiewicz, E.; Berlowska, J.; CuO@SiO2 spheres and its catalytic performances for the NO + CO
Dziugan, P.; Witonska, I. Hydrogenation of furfural over Pd−Cu/ and CO oxidation. Sci. Rep. 2015, 5, 9153.
Al2O3 catalysts. The role of interaction between palladium and copper (30) Valencia, D.; Klimova, T. Citric acid loading for MoS2-based
on determining catalytic properties. J. Mol. Catal. A: Chem. 2014, 395, catalysts supported on SBA-15. New catalytic materials with high
337−348. hydrogenolysis ability in hydrodesulfurization. Appl. Catal., B 2013,
(12) Vaidya, P. D.; Mahajani, V. V. Kinetics of Liquid-Phase 129, 137−145.
Hydrogenation of Furfuraldehyde to Furfuryl Alcohol over a Pt/C (31) Jimenez-Gomez, C. P.; Cecilia, J. A.; Martin, D. D.; Tost, R. M.;
Catalyst. Ind. Eng. Chem. Res. 2003, 42, 3881−3885. Gonzalez, J. S.; Robles, J. M.; Mariscal, R.; Torres, P. M. Gas-phase

1514 DOI: 10.1021/acs.iecr.7b04701


Ind. Eng. Chem. Res. 2018, 57, 1506−1515
Industrial & Engineering Chemistry Research Article

hydrogenation of furfural to furfuryl alcohol over Cu/ZnO catalysts. J.


Catal. 2016, 336, 107−115.
(32) Guerreiro, E. D.; Gorriz, O. F.; Rivarola, J. B.; Arrua, L. A.
Characterization of Cu/SiO2 catalysts prepared by ion exchange for
methanol dehydrogenation. Appl. Catal., A 1997, 165, 259−271.
(33) Zhu, H.; John, G.; Wei, B. Synthesis of Assembled copper
nanoparticles from copper-chelating glycolipid nanotubes. Chem. Phys.
Lett. 2005, 405, 49−52.
(34) Eren, B.; Heine, C.; Bluhm, H.; Somorjai, G. A.; Salmeron, M.
Catalyst Chemical State during CO Oxidation Reaction on Cu(111)
Studied with Ambient-Pressure X-ray Photoelectron Spectroscopy and
Near Edge X-ray Adsorption Fine Structure Spectroscopy. J. Am.
Chem. Soc. 2015, 137, 11186−11190.
(35) Yang, F.; Choi, Y. M.; Liu, P.; Hrbek, J.; Rodriguez, J. A.
Autocatalytic Reduction of a Cu2O/Cu(111) Surface by CO: STM,
XPS, and DFT Studies. J. Phys. Chem. C 2010, 114, 17042−17050.
(36) Liu, D.; Fernandez, Y.; Ola, O.; Mackintosh, S.; Maroto-Valer,
M.; Parlett, C. M. A.; Lee, A. F.; Wu, J. C. S. On the impact of Cu
dispersion on CO2 photoreduction over Cu/TiO2. Catal. Commun.
2012, 25, 78−82.
(37) Jensen, J. R.; Johannessen, T.; Livbjerg, H. An improved N2O-
method for measuring Cu-dispersion. Appl. Catal., A 2004, 266, 117−
122.
(38) Shi, Y.; Zhu, Y.; Yang, Y.; Li, Y.-W.; Jiao, H. Exploring Furfural
Catalytic Conversion on Cu(111) from Computation. ACS Catal.
2015, 5, 4020−4032.
(39) Lahti, M.; Chaudhuri, A.; Pussi, K.; Hesp, D.; McLeod, I. M.;
Dhanak, V. R.; King, M. O.; Kadodwala, M.; MacLaren, D. A. The
structural analysis of Cu(111)-Te (√3 × √3) R30° and (2√3 ×
2√3) R30° surface phases by quantitative LEED and DFT. Surf. Sci.
2014, 622, 35−43.
(40) Van Druten, G. M. R.; Ponec, V. Hydrogenation of carbonylic
compounds: Part II: The influence of alcohols on the hydrogenation of
carbonyl compounds. Appl. Catal., A 2000, 191, 163−176.
(41) Hernandez, D. V.; Caballero, J. M. R.; Gonzalez, J. S.; Tost, R.
M.; Robles, J. M.; Cruz, M. A. P.; Lopez, A. J.; Huesca, R. H.; Torres,
P. M. Furfuryl alcohol from furfural hydrogenation over copper
supported on SBA-15 silica catalysts. J. Mol. Catal. A: Chem. 2014, 383,
106−113.
(42) Avery, N. R. EELS identification of the adsorbed species from
acetone adsorption on Pt(111). Surf. Sci. 1983, 125, 771−786.
(43) Srivastava, R. D.; Guha, A. K. Kinetics and mechanism of
deactivation of Pd-Al2O3 catalyst in the gaseous phase decarbonylation
of furfural. J. Catal. 1985, 91, 254−262.
(44) Bradley, M. K.; Robinson, J.; Woodruff, D. P. The structure and
bonding of furan on Pd(111). Surf. Sci. 2010, 604, 920−925.
(45) Shi, Y.; Zhu, Y.; Yang, Y.; Li, Y.-W.; Jiao, H. Mechanisms of
Mo2C(101)-Catalyzed Furfural Selective Hydrodeoxygenation to 2-
Methylfuran from Computation. ACS Catal. 2016, 6, 6790−6803.

1515 DOI: 10.1021/acs.iecr.7b04701


Ind. Eng. Chem. Res. 2018, 57, 1506−1515

You might also like