You are on page 1of 10

View Article Online / Journal Homepage / Table of Contents for this issue

Catalysis Dynamic Article Links

Science & Technology


Cite this: Catal. Sci. Technol., 2012, 2, 1967–1976

www.rsc.org/catalysis PAPER
Catalytic hydrogenolysis of biodiesel derived glycerol to 1,2-propanediol
over Cu–MgO catalysts
M. Balaraju, K. Jagadeeswaraiah, P. S. Sai Prasad and N. Lingaiah*
Received 2nd February 2012, Accepted 3rd May 2012
DOI: 10.1039/c2cy20059g
Published on 03 May 2012. Downloaded on 11/4/2021 7:18:13 AM.

Selective hydrogenolysis of glycerol to 1,2-propanediol over Cu–MgO catalysts is reported.


A series of Cu–MgO catalysts with varying Cu content were prepared by a co-precipitation
method. The physico-chemical properties of the catalysts were derived from BET surface area,
X-ray diffraction, temperature programmed reduction of hydrogen, temperature programmed
desorption of carbon dioxide, X-ray photoelectron spectroscopy, transmission electron
microscopy and dissociative N2O adsorption techniques. The activity results showed that
Cu content on MgO has a significant role in glycerol conversion as well as formation of
1,2-propanediol. Well dispersed Cu species and accessible basic sites of MgO are the essential
requirements for high glycerol hydrogenolysis activity. 20 wt% of Cu content on MgO is
identified as an optimum Cu content. The catalyst is equally active even with crude glycerol and
glycerol containing alkali salts as impurity. Different reaction parameters were evaluated and
optimized reaction conditions were established.

1. Introduction renewable resources such as glycerol by conversion into useful


chemicals is essential for development of the society.15 Many
1,2-Propanediol (1,2-PDO), a three carbon diol with a sterio- researchers have focused on the catalytic conversion of bio-
genic center at the central carbon atom, is an important glycerol into value-added chemicals by various reactions.14,16,17
medium-value commodity chemical with annual production One such method is the conversion of glycerol into 1,2-PDO by
of 1 billion pounds in the United States.1,2 It is used in the hydrogenolysis. This approach overcomes the dependency on
preparation of polyester resins, liquid detergents, pharma- fossil fuel and substantially alters the price of 1,2-PDO.18
ceuticals, cosmetics, tobacco humectants, flavors and fragrances, Hydrogenolysis of glycerol leads to the formation of 1,2-
personal care products, paints, animal feed, antifreeze etc.1,3–5 propanediol, 1,3 propanediol (1,3-PDO) as main products,
It can be used as an alternative to toxic ethylene glycol based ethylene glycol (EG) as a degradative product and 1-propanol
deicing agents.1 The present industrial method for the synth- (1-PO), 2-propanol (2-PO) as over hydrogenolysis products.
esis of 1,2-PDO is by hydrolysis of propylene oxide with water Glycerol hydrogenolysis is generally carried out in liquid phase
at temperatures between 125 1C and 200 1C with a pressure of under elevated H2 pressure and temperature.11 Noble metal
20 bar.6,7 The propylene glycol market is under severe pressure catalysts such as supported Rh, Ru, Pt and Pd were studied
due to rapid increase in oil and natural gas costs. Propylene, extensively for glycerol hydrogenolysis.19,20 Ru based catalysts
which is a precursor to propylene oxide used to make propylene were identified as efficient catalysts for glycerol hydrogenolysis
glycol, has seen significant rise in price.8 An alternative route and the main drawback with Ru based catalysts is that they also
to synthesize 1,2-PDO is from renewable feedstocks through cleave C–C bonds which leads to the formation of EG.7,19,20
hydrogenolysis of sugars, higher polyols or glycerol in the Copper and nickel based catalysts are used as alternatives to
presence of a metal catalyst.9–12 noble metal catalysts as they are cheap and active under mild
Glycerol is a by-product of the biodiesel industry, formed reaction conditions. Moreover, the Cu and Ni based catalysts
during the transesterification of edible, non-edible oils or are known for their poor hydrogenolysis activity towards C–C
animal fats.13 Now, additional glycerol from biodiesel production bonds and high selectivity for C–O bond hydro-dehydrogenation.
is flooding the market and substantially influencing the glycerol Y. Nakagawa and Tomishige described in detail about the
price in the global market.5,14 The effective utilization of latest trends in catalyst development for hydrogenolysis of
glycerol to propylene glycol in a recent review.21
Copper based catalysts, particularly Cu–ZnO catalysts,
Catalysis Laboratory, I&PC Division, Indian Institute of Chemical have been studied for glycerol hydrogenolysis.22,23 The cata-
Technology, Hyderabad 500607, India.
E-mail: nakkalingaiah@iict.res.in; Fax: +91 40 2716 0921; lysts are active under mild reaction conditions and do not
Tel: +91 40 2719 3163 require a separate solid acid catalyst. The ZnO present in the

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 1967–1976 1967
View Article Online

catalyst drives the dehydration step to yield acetol from was separated by filtration and washed thoroughly with
glycerol, which is further hydrogenated on Cu sites to yield de-ionized water to remove traces of potassium ions. The thus
propylene glycol. A. Bienholz et al. reported 46% glycerol obtained precipitate was dried overnight at 120 1C. The
conversion with 90% selectivity towards propylene glycol over prepared catalysts are designated as 10CuMgO, 20CuMgO,
a Cu–ZnO catalyst.24 Wang and Liu obtained 23% glycerol 40CuMgO, 60CuMgO, 80CuMgO. The numbers indicate the
conversion with 83% selectivity to 1,2 propanediol at 200 1C weight percentage of Cu on the MgO support.
with 42 bar of H2 pressure for 12 h over a Cu–ZnO catalyst Conventional Cu supported on SiO2, Al2O3, ZrO2 and ZnO
prepared by a co-precipitation method.23 In our previous catalysts were prepared by a wet impregnation method using
study, Cu–ZnO catalysts with more than 90% 1,2-PDO aqueous solution of Cu(NO3)23H2O. In this method, a calcu-
selectivity were reported.21 Even though all these catalysts lated amount of aqueous solution of copper precursor was
are selective for 1,2-PDO, the overall glycerol conversion is added dropwise to supports. After allowing overnight adsorp-
limited and requires long reaction times. tion of the metal on the support, the excess solution was first
High glycerol conversion and selectivity to 1,2-PDO was evaporated to near dryness on a water bath and then the
achieved over catalytic systems containing precious metals partially dried material was dried in an air oven at 120 1C for
along with base, such as Pt/C + CaO, PtRu/C + NaOH 12 h. All catalysts were calcined finally at 400 1C for 3 h.
Published on 03 May 2012. Downloaded on 11/4/2021 7:18:13 AM.

and Ru/TiO2 + LiOH.25–27 When base is used instead of acid


as a co-catalyst the overall activity is high. They found that
2.2 Catalysts characterization
glycerol hydrogenolysis activity depends on the alkaline
strength of the base material. The glycerol hydrogenolysis The BET surface areas of the catalyst samples were calculated
proceeds in a three step mechanism in alkali solution. Initially, from N2 adsorption–desorption data acquired on an Autosorb-1
glycerol dehydrogenates to glyceraldehyde, followed by the instrument (Quantachrome, USA) at liquid N2 temperature.
dehydration of glyceraldehyde to 2-hydroxyacrolein which X-ray powder diffraction (XRD) patterns of the catalysts
subsequently hydrogenates to 1,2-PDO.28 In the case of were recorded on a Rigaku Miniflex (M/s. Rigaku Corporation,
Cu–ZnO catalysts, it is proved that the reaction proceeds Japan) X-ray diffractometer using Ni filtered Cu Ka radiation
through an acetol mechanism.22,23 If the acidic support for (l = 1.5406 Å) with a scan speed of 21 min1 and a scan range
Cu is replaced by a basic support, the reaction mechanism is of 2–801 at 30 kV and 15 mA. Particle sizes of Cu and Mg
expected to be different leading to high glycerol conversion oxides were calculated by using the Debye–Scherrer equation.
and selectivity to 1,2-PDO. There are few reports where Cu is Temperature programmed reduction (TPR) of the catalysts
supported on basic supports like MgO and hydrotalcites. was carried out in a flow of 10% H2–Ar mixture gas at a flow
X. Zheng et al. prepared solid base supported Cu catalysts rate of 30 ml min1 with a temperature ramp of 10 1C min1.
and examined for glycerol hydrogenolysis activity. They Before the TPR run the catalysts were pretreated with Ar gas
achieved high glycerol conversion with retaining of high 1,2- at 300 1C for 2 h. The hydrogen consumption was monitored
PDO (495%) selectivity over Cu–MgO and Cu/Mg/Al hydro- using a thermal conductivity detector.
talcite catalysts.29,30 Most of the researchers studied glycerol XPS measurements were conducted on a KRATOS AXIS
hydrogenolysis activity on pure and synthetic glycerol. There 165 with a DUAL anode (Mg and Al) apparatus using the
are no detailed studies about the characterization of Cu–MgO Mg Ka anode. The non-monochromatized Al-Ka X-ray source
catalysts to understand the observed catalytic activities. (hn = 1486.6 eV) was operated at 12.5 kV and 16 mA. Before
The present work is related to development of a highly acquisition of the data the sample was out-gassed for about 3 h
active non-noble Cu–MgO catalyst for the hydrogenolysis of at 100 1C under a vacuum of 1.0  107 torr to minimize
glycerol to 1,2-PDO under mild reaction conditions. The study surface contamination. The XPS instrument was calibrated
aims to understand the role of support, basicity in Cu–MgO using Au as standard. For energy calibration, the carbon 1S
catalysts. The catalytic activity of the Cu–MgO catalyst is photoelectron line was used. The carbon 1S binding energy
studied using crude glycerol and alkali salts containing glycerol was taken as 285 eV. Charge neutralization of 2 eV was used to
to know the practical applicability. The catalysts are charac- balance the charge up of the sample. The spectra were
terized by various techniques and the relation between the deconvoluted using a Sun Solaris based Vision-2 curve resolver.
glycerol hydrogenolysis activity and surface and structural The location and the full width at half maximum (FWHM)
characteristics of Cu–MgO catalysts is discussed. value for the species were first determined using the spectrum of
the pure sample. Symmetric Gaussian shapes were used in all
cases. Binding energies for identical samples were, in general,
2. Experimental reproducible within 0.1 eV.
The total basicity of the catalysts was measured by
2.1 Catalyst preparation
temperature programmed desorption of CO2 (TPD-CO2). In
A series of Cu–MgO catalysts were prepared by a co-precipitation a typical experiment, 0.1 g of catalyst was loaded and
method. Calculated amounts of aqueous solutions of pretreated in He gas at 300 1C for 2 h. After pretreatment
Cu(NO3)23H2O and Mg(NO3)26H2O were taken and preci- the temperature was brought to 100 1C and the adsorption of
pitated with dropwise addition of 0.1 M aqueous potassium CO2 was carried out by passing a mixture of 10% CO2
carbonate solution. During the precipitation the solutions balanced He gas over the catalyst for 1 h. The catalyst surface
were maintained at 70 1C. The precipitate was aged further was flushed with He gas at 100 1C for 2 h to flush off the
at the same temperature for 3 h. After cooling, the precipitate physisorbed CO2. The TPD of the catalysts was carried out in

1968 Catal. Sci. Technol., 2012, 2, 1967–1976 This journal is c The Royal Society of Chemistry 2012
View Article Online

a He gas flow at a flow rate of 30 ml min1 with a temperature the reaction temperature while maintaining the required H2
ramp of 10 1C min1. The CO2 desorption was monitored pressure. During the reaction a decrease in hydrogen pressure
using a thermal conductivity detector (TCD) of a gas is observed. After the reaction, the gas phase products were
chromatograph. collected in a gas bag and the liquid phase products were
In order to determine the exposed copper surface area and separated from the catalyst by filtration. Products in the liquid
the dispersion of the catalysts, dissociative N2O adsorption–H2 phase and the gas phase were analyzed by a gas chromato-
TPR reverse titration experiments were carried out as graph (Shimadzu 2010) equipped with an FID and a TCD.
described in the literature.31 In a typical experiment, temperature The liquid products were separated on a capillary INNOWax
programmed reduction (TPRI) was carried out to reduce the (30 m  0.25 mm  0.25 mm) column. Products were
CuO phase to Cu(0) with a 5% H2–Ar mixture with a heating also identified by GC–MS (Shimadzu, GCMS-QP2010).
rate of 10 1C min1. Then the sample which is in the form of The gas products were analyzed by using a Porapak Q column
Cu(0) phase is exposed to N2O to oxidize Cu to Cu2O by (4 m  3 mm) equipped with a thermal conductivity detector.
adsorptive decomposition of N2O at 80 1C by a continuous The products identified during glycerol hydrogenolysis are
N2O flow for 0.5 h. Then, again TPRII was carried out for the 1,2-propanediol (1,2-PDO), 1-propanol (1-PO), and 2-propanol
second time on the re-oxidized Cu2O surface in order to reduce (2-PO) (hydrogenolysis products) and ethylene glycol (EG),
Published on 03 May 2012. Downloaded on 11/4/2021 7:18:13 AM.

Cu2O to Cu. The thermal conductivity detector (TCD) was ethanol, methanol, ethane and methane.
used to measure the amount of H2-uptake in TPR experi- Conversion of the glycerol was calculated on the basis of the
ments. The Cu dispersions (D) and specific Cu surface area (S) following equation:
of the catalysts were calculated by the following equations. Moles of glycerol consumed
Conversion ð%Þ ¼  100
2  H2 consumption in TPRII Moles of glycerol initially charged
DispersionðD%Þ ¼  100
H2 consumption in TPRI
The selectivity of the products was calculated on carbon
Specific Cu surface area ðSÞ basis.
2  H2 consumption in TPRII  N Selectivity ð%Þ
¼ m2 g1
H2 consumption in TPRI  MCu  1:4  109
Moles of carbon in a specific product
where N = Avogadro’s constant, MCu = atomic mass ¼  100
Moles of carbon in all detected products
(63.456 g mol1), 1.47  1019 is the number of surface Cu
atoms per unit surface area (0.0711 nm2).
The morphology features of the catalysts were characterized 3. Results and discussion
by transmission electron microscopy (TEM). TEM investiga-
tions were carried out using a Philips CM20 (100 kV) 3.1 Catalysts characterization
transmission electron microscope equipped with a NARON Table 1 summarizes the physico-chemical properties of Cu–MgO
energy-dispersive spectrometer with a germanium detector. catalysts. BET surface area of the catalysts decreased sharply
The specimens were prepared by dispersing the samples in from 43 to 12 m2 g1, when copper loading increased from 10
methanol using an ultrasonic bath and evaporating a drop of to 60 wt%. This can be attributed to the formation of CuO
resultant suspension onto the lacey carbon support grid. The clusters on the MgO support which might be blocking the
sizes of the catalyst particles were measured by digital micro- pores of moderate surface area of MgO.
graph software (version 3.6.5, Gatan Inc.). X-ray powder diffraction patterns of calcined Cu–MgO
catalysts are shown in Fig. 1. The diffraction peaks corres-
2.3 Activity measurements
ponding to CuO phases were not observed for low copper
Hydrogenolysis of glycerol was carried out in a 100 ml haste containing catalysts (10CuMgO and 20CuMgO) suggesting
alloy PARR 4843 autoclave. Required quantity of glycerol the high dispersion of copper. The most intense reflections
diluted in de-ionized water and catalysts were charged into the corresponding to CuO at 2y of 35.51 and 38.81 were clearly
autoclave. The reactor was purged with H2 three times to flush seen for the catalysts with high CuO content (60CuMgO and
off the air. After the purge, the reactor was heated to 80CuMgO). The presence of peaks at 2y values of 42.81 and

Table 1 Physico-chemical properties of Cu–Mg catalysts

Crystal sized (nm)


Catalyst Cu/Mg ratioa BET surface area (m2 g1) Metal surface areab (m2 g1) Basicityc ( 106 mol g1) Cu MgO
10CuMgO 0.005 42.7 115.6 321 12.1 33.8
20CuMgO 0.12 39.5 120.4 290 13.7 31.7
40CuMgO 0.21 20.2 102.2 250 14.3 30.5
60CuMgO 0.58 11.8 70.9 180 25.9 28.9
80CuMgO 0.71 9.6 30.9 90 29.9 27.6
a
Atomic ratio determined by XPS analysis. b Cu metal surface area determined from N2O chemisorption. c Basicity of catalyst measured by CO2
desorption studies. d Crystal size of catalyst measured by an XRD technique.

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 1967–1976 1969
View Article Online

Table 2 Binding energies of Cu–MgO catalysts

Binding energy (eV)


Catalyst Cu 2p3/2 Cu 2p1/2 Mg 2p O 1s
10CuMgO 934.8, 943.1 954.6, 963.4 51.6 531.6, 533.6
20CuMgO 934.4, 943.3 954.1, 963.5 51.4 531.6, 533.4
40CuMgO 934.5, 943.4 954.4, 963.4 51.3 531.1, 533.8
80CuMgO 934.3, 943.6 954.5, 963.5 51.5 532.1, 534.0

The surface of the Cu–MgO catalysts was characterized by


an XPS technique. The binding energy values of Cu, Mg and
O are listed in Table 2. Fig. 3 shows the Cu 2p spectra of
calcined Cu–MgO catalysts. The catalysts show binding
energy values of 934 and 954 eV related to Cu 2p3/2 and
Cu 2p1/2 respectively. The binding energy corresponding to
Published on 03 May 2012. Downloaded on 11/4/2021 7:18:13 AM.

Cu 2p1/2 was due to a spin–orbit coupling, with an energy gap


of more than 20 eV of Cu 2p3/2. The main peaks were followed
Fig. 1 X-ray diffraction patterns of calcined Cu–Mg catalysts. (*) by strong satellite peaks around 944 and 964 eV confirming the
CuO and (#) MgO phases. presence of Cu2+ species in the calcined catalysts.33,34 In fact,
transition metal ions with unfilled 3d orbitals are reported to
show prominent satellite peaks in the core level XPS spectra
due to electron shake-up transitions because of the ligand to
metal 3d charge transfer. The structure of the satellite peak,
i.e., the peak number, intensity and splitting reflect the nature
of chemical bonding of the transition metal ions.35,36 The peak
intensity of Cu 2p electrons in XP spectra significantly
increased with increase in Cu loading from 10 to 80%. The
shift in binding energies was not observed even for a high Cu
content catalyst. These results suggested a high homogeneity
of CuO sites on the support. The O 1s spectra showed two
bands in the binding range of 530–534 eV for all the catalysts.
The Mg 2p binding energy value of 51 eV is almost the same
for all the samples. The XPS results suggest the presence of
Cu2+ species on MgO in the calcined catalysts. The surface
Cu/Mg atomic ratio of Cu–MgO catalysts was determined by
using an XPS technique and results are listed in Table 1. The
surface Cu/Mg ratio increased with Cu content of catalysts. Very
low surface Cu to Mg ratio was observed for low Cu content
Fig. 2 Temperature programmed reduction patterns of Cu–Mg catalysts. catalysts due to high Cu dispersion on the MgO support.
The copper surface area of Cu–MgO catalysts was deter-
62.21 suggests the crystalline phases of MgO in the all samples. mined by a dissociative N2O adsorption–H2 TPR reverse
The crystallite sizes of Cu and MgO particles of the catalysts titration method and the results are summarized in Table 1.
were calculated using the X-ray line broadening technique and The Cu metal area of the catalysts was found to be 115.6, 120.4
are listed in Table 1. and 102.2 m2 g1 for 10CuMgO, 20CuMgO, 40CuMgO catalysts
The reducibility of the catalysts was studied using tempera- respectively. Surface area of Cu was drastically decreased for the
ture programmed reduction of hydrogen (TPR). The TPR catalyst with high Cu loading (60CuMgO and 80CuMgO) due to
profiles of the catalysts are shown in Fig. 2. As expected, the formation of cluster particles of CuO. XRD and TPR results are
hydrogen consumption increased with increase in the copper in support of the formation of CuO clusters.
content on the support. A sharp reduction peak was observed The TEM images of the catalysts are shown in Fig. 4. The
at 240 1C for the 20CuMgO catalyst. This peak is due to the particle sizes of CuO in these catalysts were estimated by
reduction of CuO to Cu. The reduction peak was shifted TEM analysis. The calculated copper mean particle sizes of
remarkably to high temperature with increased broadness, as 10CuMgO, 20CuMgO, 40CuMgO, 60CuMgO and 80CuMgO
Cu content increased on the support. The high temperature were 10.0, 12.0, 15.7, 24.5 and 30.5 nm, respectively. The
peak can be attributed to the reduction of the cluster or bulk particle sizes of the catalysts obtained from TEM analysis were
CuO particles.32 A broad reduction peak for the catalyst with comparable with those obtained from XRD results. It was
high Cu content might be due to merging of reduction peaks identified that the copper metal particles in Cu–MgO catalysts
related to dispersed and cluster CuO species. This indicates were smaller than those of Cu–ZnO catalysts.22 The difference
that copper oxide species with different redox behavior were in metal particle size on different supports can be related to the
present on the MgO support. intrinsic property of the support. Many previous studies have

1970 Catal. Sci. Technol., 2012, 2, 1967–1976 This journal is c The Royal Society of Chemistry 2012
Published on 03 May 2012. Downloaded on 11/4/2021 7:18:13 AM. View Article Online

Fig. 3 XPS of calcined Cu–Mg catalysts. (a) 10CuMgO, (b) 20CuMgO, (c) 60CuMgO, (d) 80CuMgO.

shown that the support material can influence the metal particle glyceraldehyde and it subsequently dehydrates to hydroxy-
size and consequently influence the catalytic performance.37,38 acrolein on basic sites of catalyst. Further hydrogenation of
these reaction intermediates to 1,2-PDO took place over
metallic sites. A considerable amount (B5%) of acetol (inter-
3.2 Activity measurements
mediate) was observed in our previous studies whereas acetol
3.2.1 Influence of Cu loading on MgO on glycerol hydro- is not identified in the present study.22 Therefore, it is con-
genolysis. Table 3 reports glycerol hydrogenolysis activity of cluded that the reaction proceeds over the catalysts with acidic
Cu–MgO catalysts. About 30% glycerol conversion with 92% supports as well as over those with basic supports through
selectivity to 1,2-PDO was achieved over the 10CuMgO different reaction mechanisms as shown in Scheme 1.
catalyst. As copper content on MgO increased from 10 to From the above reaction mechanism, it is known that basicity
20 wt%, the glycerol conversion also raised substantially from and copper metal area are essential for glycerol hydrogenolysis.
30 to 50% with retaining of the similar selectivity to 1,2-PDO. During glycerol hydrogenolysis, Cu and MgO sites play an
Conversion of glycerol is gradually decreased with further important role in glycerol conversion and 1,2-PDO selectivity.
increase in Cu content. The high copper containing catalysts The observed glycerol hydrogenolysis activity is related to the
particularly 60CuMgO and 80CuMgO showed very low activ- Cu metal area and basicity of the catalyst. The relation of Cu
ity. From above results it is noticed that low Cu containing metal surface area and basicity of the catalysts with glycerol
catalysts showed high glycerol hydrogenolysis activity. It is conversion is shown in Fig. 5. A linear correlation is observed
also reported that in the case of Cu/MgO high glycerol between Cu metal surface area and glycerol conversion as
conversion is achieved with low Cu content catalysts.29 The shown in Fig. 5(A). The low Cu containing catalyst showed
present results are comparable with the results reported by maximum Cu metal area and exhibited high activity. Metal
Yuan et al.29 However, they reported about 72% glycerol surface area and dispersion rapidly decreased for the catalysts
conversion when the reaction was carried out for 20 h. with high Cu content such as 60CuMgO and 80CuMgO due to
Consistent selectivity to 1,2-PDO was observed for all Cu–MgO formation of multi-layer CuO clusters on the support. The
catalysts irrespective of glycerol conversion. TPR, XRD and TEM analyses are in support of these observa-
Previously, many authors reported Ru or Cu supported tions. The relation between glycerol conversion and catalyst
on solid acid catalysts for glycerol hydrogenolysis.5,22,23,38,39 basicity shown in Fig. 5(B) suggests that the catalyst with high
Glycerol hydrogenolysis was also carried out by adding basicity exhibits high glycerol conversion. The basicity of low
base as a co-catalyst over carbon supported noble metal Cu content catalysts is also high. Basicity of the 10CuMgO
catalysts.25–27 The hydrogenolysis activity of the present catalyst was found to be 321 mmol g1. As CuO content
Cu–MgO catalyst is higher than that of reported Cu–ZnO increased from 10 to 80%, basicity of the Cu–MgO catalysts
catalysts and comparable with that of hybrid Ru/C + gradually decreased from 321 to 90 mmol g1 due to blockage of
Amberlyst and Pt/C or Ru/C + NaOH. Montassier et al. surface MgO sites by CuO clusters. As these catalysts contain
proposed a reaction mechanism for the conversion of glycerol low amounts of MgO, their overall basicity is also minimum.30
to 1,2-PDO under basic reaction conditions.28 According to These results indicate that the glycerol conversion depends on
this reaction mechanism, dehydrogenation of glycerol leads to the basicity and Cu metal area. As a whole a judicious amount

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 1967–1976 1971
Published on 03 May 2012. Downloaded on 11/4/2021 7:18:13 AM. View Article Online

Fig. 4 TEM images of Cu–Mg catalysts. (a) 10CuMgO, (b) 20CuMgO, (c) 40CuMgO, (d) 60CuMgO, (e) 80CuMgO.

Table 3 Influence of Cu content in Cu–Mg catalysts on glycerol temperature for glycerol hydrogenolysis, the reaction was
hydrogenolysis carried out from 160 to 220 1C. A significant influence of the
Selectivity (%) reaction temperature on glycerol conversion and selectivity to
1,2-PDO was observed. With increase in reaction temperature,
Catalyst Conv. (%) 1,2-PDO EG Others
a uniform increase in glycerol conversion was observed.
10CuMgO 30.2 92.4 5.4 2.8 Substantial increase in glycerol conversion was observed
20CuMgO 49.3 92.3 5.9 1.8 between reaction temperatures of 180 and 200 1C. The selectivity
40CuMgO 36.6 90.3 6.7 3.0
60CuMgO 18.0 90.0 6.9 3.1 to 1,2-PDO increased up to 200 1C and decreased with further
80CuMgO 6.1 87.6 5.6 6.8 increase in reaction temperature. This indicates that when
Reaction conditions: 20 wt% glycerol aqueous solution: 50 ml, H2 reaction temperature is more than 200 1C, excessive hydro-
pressure: 40 bar, reaction time: 8 h, catalyst weight: 0.6 g (6%), genolysis of glycerol and 1,2-PDO is taking place and that
reaction temperature: 200 1C. leads to formation of lower alcohols like methanol, ethanol
and other gaseous products. A reaction temperature of 200 1C
was identified as the optimum for selective hydrogenolysis of
of basicity and Cu metal surface area is required for the catalyst glycerol to propylene glycol over Cu–MgO catalysts.
to show better glycerol conversion and selectivity.
3.2.3 Effect of glycerol concentration on glycerol hydro-
3.2.2 Effect of reaction temperature on hydrogenolysis of genolysis. The effect of glycerol concentration on the hydro-
glycerol. Fig. 6 shows the influence of reaction temperature on genolysis of glycerol was studied and the results are shown in
hydrogenolysis of glycerol. In order to optimize the reaction Fig. 7. The conversion of glycerol decreased with increase in

1972 Catal. Sci. Technol., 2012, 2, 1967–1976 This journal is c The Royal Society of Chemistry 2012
View Article Online

Scheme 1 Glycerol hydrogenolysis mechanism over acid or base catalysts.


Published on 03 May 2012. Downloaded on 11/4/2021 7:18:13 AM.

Fig. 5 Correlation of (A) Cu metal area and (B) basicity of Cu–MgO catalysts with glycerol conversion.

Fig. 6 Effect of reaction temperature on hydrogenolysis activity over Fig. 7 Effect of glycerol concentration on hydrogenolysis activity
20CuMgO catalyst. Reaction conditions: 20 wt% glycerol aqueous over 20CuMgO catalyst. Reaction conditions: glycerol aqueous
solution: 50 ml, H2 pressure: 40 bar, reaction time: 8 h, catalyst weight: solution: 50 ml, H2 pressure: 40 bar, reaction time: 8 h, catalyst
0.6 g (6%). weight: 0.6 g (6%), reaction temperature: 200 1C.

the concentration of glycerol. The decrease in conversion at 3.2.4 Effect of reaction time on glycerol hydrogenolysis.
higher glycerol concentration may be due to the availability of Reaction time is an important parameter and long reaction
less number of Cu active sites for the large amount of glycerol times (424 h) were adopted usually for glycerol hydrogeno-
molecules present in the reaction medium. However, the lysis to obtain reasonable conversion.4,10 The conversion of
selectivity to 1,2-PDO was unaltered with glycerol concen- glycerol and selectivity to 1,2-PDO at different reaction times
tration, implying the high selectivity of the Cu–MgO catalysts. were studied over the 20CuMgO catalyst. As shown in Fig. 8,
The catalyst showed considerable glycerol conversion even at about 32% glycerol conversion and 86% selectivity towards
high glycerol concentrations. 1,2-PDO were achieved within 4 h of reaction time. A considerable

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 1967–1976 1973
Published on 03 May 2012. Downloaded on 11/4/2021 7:18:13 AM. View Article Online

Fig. 8 Effect of reaction time on hydrogenolysis activity over Fig. 10 Effect of catalyst weight on hydrogenolysis activity over
20CuMgO catalyst. Reaction conditions: 20 wt% glycerol aqueous 20CuMgO catalyst. Reaction conditions: 20 wt% glycerol aqueous
solution: 50 ml, H2 pressure: 40 bar, catalyst weight: 0.6 g (6%), solution: 50 ml, H2 pressure: 40 bar, reaction time: 8 h, reaction
reaction temperature: 200 1C. temperature: 200 1C.

change in glycerol conversion was observed up to a reaction As expected the conversion of the glycerol marginally increased
time of 6 h, thereafter the increase in conversion was marginal. as the hydrogen pressure increased from 20 to 60 bar. The
The Cu–MgO catalysts were very active compared to Cu–ZnO hydrogen atoms at higher pressure compete with the
catalysts which took longer reaction times to obtain reasonable re-adsorption of an intermediate on active catalyst sites
glycerol conversion.21 The catalyst was also selective in forming leading to the decrease in glycerol conversion. A similar
1,2-PDO in more than 90% yield, irrespective of reaction time. observation was reported by Rode et al. for a copper chromite
The higher selectivity to 1,2-PDO even at longer reaction times catalyst.40 The Cu based catalysts are highly active at moderate
reflects that the present catalyst was efficient to selectively hydrogen pressure unlike noble metal catalysts which are active at
cleave only C–O bonds instead of C–C bonds. elevated H2 pressure. Hydrogen pressure in the range of 40 to
80 bar resulted in a marginal decrease in the selectivity to 1,2-PDO
3.2.5 Effect of hydrogen pressure on glycerol hydrogenolysis. from 92 to 89% due to enhanced hydrogenation of 1,2-PDO to
Reactions were carried out at different hydrogen pressures 2-propanol. It was interesting to note that an increase in H2
from 20 to 80 bar at a constant reaction temperature of 200 1C pressure did not affect the selectivity to the degradation product
to determine the effect of hydrogen pressure on the overall EG because copper is selective for cleavage of C–O bonds.
activity and the results are shown in Fig. 9. Generally, glycerol
hydrogenolysis at low hydrogen pressure is important to maximize 3.2.6 Effect of catalyst weight on glycerol hydrogenolysis.
the utility of the existing equipment for performing hydrogenolysis. Fig. 10 shows the variation in the glycerol conversion with
increase in the catalyst weight during glycerol hydrogenolysis.
The glycerol conversion varied with change in catalyst loading.
As the catalyst weight increased from 4 to 10%, the glycerol
conversion increased from 32 to 60%. The increase in catalyst
weight results in more number of available active sites for
hydrogenolysis reaction to take place. A similar increase in
hydrogenolysis activity with increase in catalyst concentration
is also reported.2

3.2.7 Glycerol hydrogenolysis on crude glycerol. Hydro-


genolysis of glycerol was studied using 5% sodium sulphate
containing synthetic glycerol and crude glycerol obtained from
a biodiesel plant to know the practical applicability of the
present catalyst. These results are listed in Table 4. The
bioglycerol obtained from the biodiesel industry generally
contains alkali salts, as alkali is used as catalyst for biodiesel
synthesis. During the neutralization of alkali it leads to
Fig. 9 Effect of reaction pressure on hydrogenolysis activity over formation of alkali salts. It is important to study glycerol hydro-
20CuMgO catalyst. Reaction conditions: 20 wt% glycerol aqueous genolysis over Na2SO4 containing glycerol. The present catalyst
solution: 50 ml, reaction time: 8 h, catalyst weight: 0.6 g (6%), reaction showed about 41% glycerol conversion with 92% 1,2-PDO
temperature: 200 1C. selectivity even with sodium sulphate containing glycerol.

1974 Catal. Sci. Technol., 2012, 2, 1967–1976 This journal is c The Royal Society of Chemistry 2012
View Article Online

Table 4 Hydrogenolysis of crude glycerol over 20CuMgO catalyst The Cu–MgO catalysts are highly active and selective to
1,2-PDO in glycerol hydrogenolysis under mild reaction
Selectivity (%)
conditions. The content of Cu and its dispersion, metal surface
Type of glycerol Conversion (%) 1,2-PDO EG Others area and the basicity of the catalyst are responsible for the
Pure glycerol 49.3 92.3 5.9 1.8 high glycerol hydrogenolysis activity. The catalyst with low Cu
5% Na2SO4 + pure glycerol 41.6 92.3 5.0 2.7 content (20 wt%) on MgO showed the highest activity with
Crude glycerol 44.5 92 4.9 3.1 50% glycerol conversion and 92% selectivity to 1,2-PDO. The
Pure glycerola 45.8 92.3 5.8 1.9
Pure glycerolb 39.7 92.2 4.5 3.3 glycerol conversion also depends on the reaction parameters,
and optimum reaction conditions were established. The
Reaction conditions: 20 wt% glycerol aqueous solution: 50 ml, H2
pressure: 40 bar, reaction time: 8 h, catalyst weight: 0.6 g (6%), reaction
present Cu–MgO catalyst tolerates even alkali salts and other
temperature: 200 1C. a Activity on 200 ml scale. b Activity on 500 ml scale. impurities present in crude glycerol.

Acknowledgements
Table 5 Comparison of conventional Cu catalysts with 20CuMgO
catalyst for the hydrogenolysis of glycerol The authors thank Department of Science & Technology,
Published on 03 May 2012. Downloaded on 11/4/2021 7:18:13 AM.

New Delhi, India, for financial support. MB and KJ thank


Catalyst Conversion (%) Selectivity to 1,2PDO (%)
CSIR for the award of Senior Research Fellowships.
20% Cu–SiO2 20.0 83.6
20% Cu–Al2O3 25.3 85.6
20% Cu–ZrO2 10.2 62.0 References
20% Cu–ZnO 37.0 92.0
20% Cu–MgO 49.3 92.3 1 Chemical Marketing Reporter, 1092-0110, 260 (11) (September 24,
2001) 30.
Reaction conditions: 20 wt% glycerol aqueous solution: 50 ml, H2 2 M. A. Dasari, P. Kiatsimkul, W. R. Sutterlin and G. J. Suppes,
pressure: 40 bar, reaction time: 8 h, catalyst weight: 0.6 g (6%), Appl. Catal., A, 2005, 281, 225–231.
reaction temperature: 200 1C. 3 A. Behr, J. Eilting, K. Irawadi, J. Leschinski and F. Lindner, Green
Chem., 2008, 10, 13–30.
4 J. Chaminand, L. Djakovitch, P. Gallezot, P. Marion, C. Pinel and
The decrease in glycerol conversion is marginal when com- C. Rosier, Green Chem., 2004, 6, 359–361.
pared to the results of pure glycerol. 5 M. Balaraju, V. Rekha, B. L. A. P. Devi, R. B. N. Prasad, P. S. S.
Prasad and N. Lingaiah, Appl. Catal., A, 2010, 384, 107–114.
Crude glycerol was obtained from a biodiesel plant where 6 R. Perrin and J. P. Scharff, Chimie industrielle, ed. Masson, 1993.
sodium hydroxide is used as catalyst. Demethanolised crude 7 M. Balaraju, V. Rekha, P. S. S. Prasad, B. L. A. P. Devi, R. B. N.
glycerol was used for glycerol hydrogenolysis under optimized Prasad and N. Lingaiah, Appl. Catal., A, 2009, 354, 82–87.
8 A. Brandner, K. Lehnert, A. Bienholz, M. Lucas and P. Claus,
reaction conditions. A maximum glycerol conversion of about Top. Catal., 2009, 52, 278–287.
44% with 92% selectivity toward 1,2-PDO was obtained. The 9 S. P. Chopade, D. J. Miller, J. E. Jackson, T. A. Werpy, J. G. Frye
activity results show that the Cu–MgO catalyst is active even Jr. and A. H. Zacher, US Pat., US 6291725 to Board of Trustees
for crude glycerol to synthesize 1,2-PDO. It is very important operating Michigan State University, Battelle Memorial Institute,
Pacific Northwest Laboratory, 2001.
to mention that the present catalyst is resistant towards the 10 J. C. Chao and D. T. A. Huibers, US Pat., US 4366332 to
salts and other impurities present in crude glycerol. Hydrocarbon Research Inc, 1982.
11 B. Casale and A. M. Gomez, US Pat., US 5276181 to Novamont
3.2.8 Comparison of present Cu/MgO catalyst with conven- S.p.A, 1994.
tional Cu catalysts. The conventional Cu based catalysts are 12 P. Bloom, US Pat., US 7928148 to Archer Daniels Midland
Company, 2008.
prepared by an impregnation method keeping the same weight 13 J. M. Marchetti, V. U. Miguel and A. F. Errazu, Renewable
percentage of Cu as that of Cu/MgO catalysts. These catalysts Sustainable Energy Rev., 2007, 11, 1300–1311.
were studied for the glycerol hydrogenolysis under the optimum 14 D. T. Johnson and K. A. Taconi, Environ. Prog., 2007, 26, 338–348.
15 R. D. Cortright, R. R. Davda and J. A. Dumesic, Nature, 2002,
reaction conditions and the results are shown in Table 5. The 418, 964–967.
Cu supported on supports such as alumina, silica, zirconia, zinc 16 M. Pagliaro, R. Ciriminna, H. Kimura, M. Rossi and C. D. Pina,
oxide exhibited lower activity when compared to Cu/MgO Eur. J. Lipid Sci. Technol., 2009, 111, 788–799.
catalysts. The Cu supported on conventional supports (Al2O3, 17 N. R. Shiju, D. R. Brown, K. Wilson and G. Rothenberg, Top.
Catal., 2010, 53, 1217–1223.
SiO2, ZrO2) showed relatively less activity under the present 18 L. C. Meher, R. Gopinath, S. N. Naik and A. K. Dalai, Ind. Eng.
experimental conditions. The results suggest that the glycerol Chem. Res., 2009, 48, 1840–1846.
hydrogenolysis activity depends on the nature of the catalysts 19 T. Miyazawa, Y. Kusunoki, K. Kunimori and K. Tomishige,
J. Catal., 2006, 240, 213–221.
and their method of preparation. The commercial Cu catalysts
20 Y. Kusunoki, T. Miyazawa, K. Kunimori and K. Tomishige,
copper and copper chromite (Sud-Chemie) catalysts gave about Catal. Commun., 2005, 6, 645–649.
21.1 and 46.6% yield of 1,2PDO with 39.8 and 85% selectivity 21 Y. Nakagawa and K. Tomishige, Catal. Sci. Technol., 2011, 1,
respectively.2 These results suggest that the present catalyst is 179–190.
22 M. Balaraju, V. Rekha, P. S. S. Prasad, R. B. N. Prasad and
highly active and selective for the hydrogenolysis of glycerol. N. Lingaiah, Catal. Lett., 2008, 126, 119–124.
23 S. Wang and H. Liu, Catal. Lett., 2007, 117, 62–67.
24 A. Bienholz, F. Schwab and P. Claus, Green Chem., 2010, 12,
4. Conclusions 290–295.
25 E. P. Maris and R. J. Davis, J. Catal., 2007, 249, 328–337.
The use of a basic support like MgO for Cu based catalysts 26 E. P. Maris, W. C. Ketchie, M. Murayama and R. J. Davis,
alters the reaction mechanism of glycerol hydrogenolysis. J. Catal., 2007, 251, 281–294.

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 1967–1976 1975
View Article Online

27 J. Feng, J. B. Wang, Y. F. Zhou, H. Y. Fu, H. Chen and X. J. Li, 34 C. C. Chusuei, M. A. Brookshier and D. W. Goodman, Langmuir,
Chem. Lett., 2007, 1274–1275. 1999, 15, 2806–2808.
28 C. Montassier, D. Giraud and J. Barbier, Heterogeneous Catalysis 35 D. C. Frost, A. Ishitani and C. A. McDowell, Mol. Phys., 1972,
and Fine Chemicals, 1988, pp. 165–170. 24, 861.
29 Z. Yuan, J. Wang, L. Wang, W. Xie, P. Chen, Z. Hou and 36 T. A. Carlson, Photoelectron and Auger Spectroscopy, Plenum
X. Zheng, Bioresour. Technol., 2010, 101, 7088–7092. Press, New York, 1975.
30 Z. Yuan, L. Wang, J. Wang, S. Xia, P. Chen, Z. Hou and 37 J. Feng, H. Fu, J. Wang, R. Li, H. Chen and X. Li, Catal.
X. Zheng, Appl. Catal., B, 2011, 101, 431–440. Commun., 2008, 9, 1458–1464.
31 C. J. G. Van Der Grift, A. F. H. Wielers, B. P. J. Jogh, J. Van 38 E. S. Vasiliadou, E. Heracleous, I. A. Vasalos and
Beunum, M. De Boer, M. Versluijs-Helder and J. W. Geus, A. A. Lemonidou, Appl. Catal., B, 2009, 92, 90–99.
J. Catal., 1991, 131, 178–189. 39 L. Guo, J. Zhou, J. Mao, X. Guo and S. Zhang, Appl. Catal., A,
32 S. Xu, C. Huang, J. Zhang and B. Chen, Korean J. Chem. Eng., 2009, 367, 93–98.
2009, 26, 1568–1573. 40 C. V. Rode, A. A. Ghalwadkar, R. B. Mane, A. M. Hengne,
33 R. Kam, C. Selomulya, R. Amal and J. Scott, J. Catal., 2010, 273, S. T. Jadkar and N. S. Biradar, Org. Process Res. Dev., 2010, 14,
73–81. 1385–1392.
Published on 03 May 2012. Downloaded on 11/4/2021 7:18:13 AM.

1976 Catal. Sci. Technol., 2012, 2, 1967–1976 This journal is c The Royal Society of Chemistry 2012

You might also like