You are on page 1of 8

Applied Catalysis A: General 378 (2010) 11–18

Contents lists available at ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Gas-phase dehydration of glycerol to acrolein catalysed by caesium


heteropoly salt
Abdullah Alhanash, Elena F. Kozhevnikova, Ivan V. Kozhevnikov *
Department of Chemistry, University of Liverpool, Crown Street, Liverpool L69 7ZD, UK

A R T I C L E I N F O A B S T R A C T

Article history: Caesium 12-tungstophosphate, Cs2.5H0.5PW12O40 (CsPW), possessing strong Brønsted acid sites is an
Received 9 December 2009 active catalyst for the dehydration of glycerol to acrolein in the gas-phase process at 275 8C and 1 bar
Received in revised form 7 January 2010 pressure. The initial glycerol conversion amounts to 100% at 98% acrolein selectivity, however, decreases
Accepted 28 January 2010
significantly with the time on stream (40% after 6 h) due to catalyst coking, without impairing acrolein
Available online 6 February 2010
selectivity. Doping CsPW with platinum group metals (PGM) (0.3–0.5%) together with co-feeding
hydrogen improve catalyst stability to deactivation, while maintaining high selectivity to acrolein. The
Keywords:
enhancing effect of PGM was found to increase in the order: Ru  Pt < Pd. The catalyst 0.5%Pd/CsPW
Glycerol
Acrolein
gives 96% acrolein selectivity at 79% glycerol conversion, with a specific rate of acrolein production of
Dehydration 23 mmol h1 gcat1 at 275 8C and 5 h time on stream, exceeding that reported previously for supported
Acid catalysis heteropoly acids (5–11 mmol h1 gcat1 per total catalyst mass). Evidence is presented regarding the
Heteropoly acid nature of acid sites required for the dehydration of glycerol to acrolein, supporting the importance of
strong Brønsted sites for this reaction.
ß 2010 Elsevier B.V. All rights reserved.

1. Introduction heterogeneous catalysis appears to be much more attractive. A


number of solid acid catalysts for the dehydration of glycerol have
Production of chemicals by catalytic transformation of bio- been reported including metal sulphates and phosphates, zeolites,
sustainable resources is a global challenge [1]. Recent development supported mineral acids, niobia, sulphated zirconia and heteropoly
of biodiesel production by transesterification of plant oils and acids (for a recent review see [8] and references therein). The
animal fat has made a large amount of glycerol available as a synthesis of acrolein is an endothermic reaction with a standard
renewable feedstock for the synthesis of a wide range of value- enthalpy of 19.8 kJ mol1 for a gas-phase process. Typically, the
added chemicals [1–5]. Significant research effort has been focused reaction is carried out in the temperature range of 260–350 8C
on the conversion of glycerol by catalytic processes such as using an aqueous glycerol as a feedstock. Since much steam is
reforming, oxidation, hydrogenolysis, esterification, etherification present in the gas flow, the catalyst should have good water
[1–6]. However, one of the most important directions for the tolerance. The catalyst efficiency in acrolein synthesis is enhanced
utilisation of glycerol is its dehydration to acrolein. Acrolein is an with increasing catalyst acidity [3,4,8]. Supported Keggin hetero-
important intermediate for the chemical and agricultural indus- poly acids, possessing very strong Brønsted acidity, have been
tries, which is currently produced by oxidation of petroleum- found amongst the most efficient catalysts for glycerol to acrolein
derived propene [7]. It can also be obtained from glycerol via acid- conversion in the gas phase [9–12]. In particular, with 12-
catalysed dehydration in the gas or liquid phase [8]. The tungstosilicic acid, H4SiW12O40, possessing a higher water toler-
development of an alternative environmentally benign process ance compared to other Keggin heteropoly acids, acrolein yields of
for the production of acrolein based on the biomass-derived 80% have been achieved at an optimum temperature of 275 8C
glycerol as a sustainable feedstock is a real challenge [2–5,8]. [9].
The dehydration of glycerol to acrolein has been known since As regards to the reaction mechanism, the dehydration of
the nineteenth century [3,4,8]. Homogeneous and heterogeneous glycerol on acid catalysts is suggested to proceed via the formation
acid catalysts as well as biocatalysts have been used for this of 3-hydroxypropanal, with 1-hydroxyacetone (acetol) formed as a
reaction [3,4,8]. Due to technical and environmental advantages, relatively stable by-product (Scheme 1) [8,9,12–14]. Acetol is the
main product (>90% selectivity) in the gas-phase dehydration of
glycerol on mixed oxide catalysts, such as copper chromite [15],
* Corresponding author. Fax: +44 151 794 3588. and supported metal catalysts, e.g., Cu/Al2O3 [16]. This may
E-mail address: I.V.Kozhevnikov@liverpool.ac.uk (I.V. Kozhevnikov). indicate that Lewis acid sites and metal sites are important for the

0926-860X/$ – see front matter ß 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2010.01.043
12 A. Alhanash et al. / Applied Catalysis A: General 378 (2010) 11–18

supported on Aerosil 300 silica, 20%HPW/SiO2, and the corre-


sponding Pd-doped catalyst 2%Pd/20%HPW/SiO2 were prepared as
described elsewhere [28]. Zn(II)–Cr(III) (1:1) mixed oxide catalysts
of 45–180 mm particle size were prepared as described earlier [29].
All the catalysts were stored in a desiccator over P2O5. Table 1
shows the texture of catalysts under study (surface area, pore
volume and average pore diameter as well as metal dispersion).

2.3. Catalyst characterisation


Scheme 1. Dehydration of glycerol over acid catalysts.

Elemental analysis of catalysts was performed by inductively


formation of acetol, whereas the formation of 3-hydroxypropanal coupled plasma (ICP) atomic emission spectroscopy on a Spectro
followed by its dehydration to acrolein is favoured in the presence Ciros CCD spectrometer. Surface area and porosity of catalysts
of strong Brønsted acid sites. were determined by the BET method from nitrogen physisorption
The major drawback to the gas-phase dehydration of glycerol measured on a Micromeritics ASAP 2000 instrument at 196 8C.
on acid catalysts is catalyst deactivation due to extensive coke Prior to the measurement, the samples were evacuated at 250 8C
deposition on the catalyst surface as well as the difficulty of direct for 3 h. Powder X-ray diffraction (XRD) pattern of catalysts was
usage of the crude glycerol obtained from biodiesel production [8]. recorded on a Philips X’Pert PRO diffractometer with Cu Ka
Catalyst regeneration by combustion of coke has been attempted (l = 1.54 Å) radiation. Thermogravimetric analysis (TGA) of fresh
[8], either in situ continuously [17,18] or ex situ periodically [19], and spent catalysts was performed using a Perkin Elmer TGA 7
although with limited success. Glycerol dehydration using instrument at a heating rate of 20 8C min1 under N2 or air flow.
fluidised catalyst bed with catalyst circulation between the reactor Diffuse reflectance infrared Fourier transform (DRIFT) spectra of
and regenerator similar to the fluidised catalytic cracking process fresh and spent catalysts were taken on a Nicolet NEXUS FTIR
has been demonstrated [14]. It should be noted that regeneration spectrometer using powdered catalyst mixtures with KBr. The
of heteropoly acid catalysts by combustion of coke is difficult DRIFT spectra of adsorbed pyridine were obtained using the same
because of relatively low thermal stability of heteropoly acids [20]. instrument equipped with a controlled environment chamber
Here we report the acidic heteropoly salt, caesium 12- (Spectra-Tech Inc., model 0030-101) as described elsewhere [29].
tungstophosphate Cs2.5H0.5PW12O40 (CsPW), as an efficient cata- Metal dispersion on the catalyst surface was measured by
lyst for the dehydraion of glycerol to acrolein in the gas phase. hydrogen chemisorption on a Micromeritics TPD/TPR 2900
CsPW is well-known as a water-insoluble strong Brønsted acid and instrument as described previously [29]. A catalyst sample
a versatile solid acid catalyst possessing high thermal stability (100 mg) pre-exposed to air at room temperature was placed in
(500 8C) and water tolerance [21–23]. We also attempt to a glass sample tube connected to the instrument and stabilised at
enhance catalyst lifetime and reduce coke deposition by doping 100 8C under nitrogen flow. 50 mL pulses of pure H2 were injected
CsPW with platinum group metals (PGM), such as Ru, Pd and Pt, in the flow in 3 min intervals until the catalyst was saturated with
and co-feeding hydrogen to the reaction system. This approach, hydrogen. Hydrogen did not adsorb on CsPW under such
involving bifunctional acid–metal catalysis, has been effectively conditions, and no reduction of CsPW in the PGM-doped catalysts
used in petrochemical industry, for example, alkane hydroisome- was observed. From temperature-programmed reduction studies
risation [24,25]. Finally, new mechanistic evidence is presented [30], solid H3PW12O40 doped with 0.5% Pd is reduced by H2 only
regarding the nature of acid sites required for the dehydration of above 300 8C. The metal dispersion, D, defined as the fraction of
glycerol to acrolein. metal (M) at the surface, D = Ms/Mtotal, was calculated assuming
the stoichiometry of H2 adsorption: MsO + 1.5H2 ! MsH + H2O
2. Experimental [31,32]. The average size of metal particles, d, was obtained from
the empirical equation d (nm) = 0.9/D [32].
2.1. Chemicals
2.4. Catalyst testing
RuCl3xH2O (99.98%), H2PtCl6xH2O (99%), Pd(OAc)2 (99%),
H3PW12O40xH2O, Cs2CO3 (99%) and acetol (90%) were purchased Dehydration of glycerol was carried out in a vertical fixed-bed
from Aldrich. Glycerol (>99%) was from Alfa Aesar. N2 and H2 gases Pyrex tubular reactor of 28 cm length and 0.8 cm internal diameter
(>99%) were supplied by the British Oxygen Company. with a catalyst bed set at the middle of reactor length. A glass wool
plug was placed in the top of reactor to enhance evaporation of
2.2. Catalyst preparation liquid glycerol feed before reaching the catalyst bed. The reactor
was placed in an electrically heated furnace. The temperature was
Cs2.5H0.5PW12O40 (CsPW) was prepared from H3PW12O40 and controlled by a thermocouple which was located near catalyst bed.
Cs2CO3 as a white crystalline powder as described elsewhere [26]. Nitrogen or hydrogen was used as a carrier gas at a flow rate of
Cs2HPW12O40 and Cs3.5H0.5SiW12O40 were prepared similarly. 15 mL min1, which was controlled by Brooks mass flow
PGM-doped CsPW catalysts were prepared by impregnating CsPW controllers. The reactor was packed with 0.30 g of catalyst powder
with a solution of appropriate metal precursor: 0.1 M aqueous (45–180 mm particle size). A glycerol–water (10:90, w/w) solution
solution of RuCl3 and H2PtCl6 in the case of Ru/CsPW and Pt/CsPW was fed into the top of reactor by a Kontron Instruments 422 HPLC
catalysts respectively [27] and 0.02 M benzene solution of pump at a flow rate of 0.14 mL min1, which corresponds to a
Pd(OAc)2 in the case of Pd/CsPW [28]. The slurry was stirred at glycerol WHSV of 2.8 h1. Reaction products together with
room temperature, dried in a rotary evaporator at 60 8C, followed unconverted glycerol were collected hourly in an ice trap (0 8C)
by reduction by H2 at 250 8C for 2 h. The catalysts were sieved into and subjected to offline GC analysis (Varian CP-3800 gas
a powder of 45–180 mm particle size. Some PGM-doped catalysts chromatograph, flame ionisation detector, 30 m  0.32 mm CP-
containing 0.5 wt% of metal dopant were prepared by thoroughly WAX 52 CB capillary column) using ethanol as a standard. Glycerol
mixing CsPW with carbon-supported PGM catalysts, 10%Pd/C, conversion, product selectivities and product yields were calcu-
10%Pt/C and 5%Ru/C, obtained from Aldrich. H3PW12O40 (HPW) lated using Eqs. (1)–(3). Carbon balance was estimated as the
A. Alhanash et al. / Applied Catalysis A: General 378 (2010) 11–18 13

Table 1
Catalysts texture.

Catalysta SBETb (m2 g1) Pore volumec (cm3 g1) Pore diameterd (Å) Metal dispersione

CsPW 130 0.06 32


Cs2HPW12O40 70 0.034 19
Cs3.5H0.5 SiW12O40 149 0.11 30
0.3%Pt/CsPW 84 0.051 24 0.52
0.5%Ru/CsPW 117 0.094 32 0.54
0.5%Pd/CsPW 84 0.051 24 0.76
20%HPW/SiO2 205 0.74 144
2%Pd/20%HPW/SiO2 200 0.86 175 0.66
Zn–Cr (1:1) 132 0.11 33
0.3%Pd/Zn–Cr (1:1) 112 0.09 35 0.37
a
CsPW stands for Cs2.5H0.5PW12O40 and HPW for H3PW12O40.
b
BET surface area.
c
Single point total pore volume.
d
Average pore diameter by the BET method.
e
From H2 adsorption.

percentage of the reacted carbon atoms found in organic products. the glycerol WHSV 2.8 h1, corresponding to the glycerol molar
Typically, it was close to 100% within 5% experimental error. Non- flow rate of 30 mmol h1 gcat1. Nitrogen was used as a carrier gas.
analysed gaseous by-products and coke deposited on the catalyst had To withstand severe steaming the catalyst should have strong
small effect on the carbon balance and could be neglected unless steam resistance. The results presented here show that the bulk
stated otherwise. CsPW salt was fit for the purpose.
The CsPW catalyst exhibited high initial activity, with the
Moles of glycerol reacted
Glycerol conversion ð%Þ ¼  100 (1) glycerol conversion amounting to 100% at 98% acrolein selectivity
Moles of glycerol fed
after 1 h on stream. The conversion, however, decreased sharply
Moles of product with the time on stream, down to 40% after 6 h, without
Product yield ð%Þ ¼  100 (2) impairing acrolein selectivity (Fig. 1). Significant amount of coke
Moles of glycerol fed
(8.9 wt%) was deposited on the catalyst surface during reaction
Product selectivity (Table 2), which was the likely cause of catalyst deactivation [8,9–
Product selectivity ð%Þ ¼  100 (3) 12]. Similar deactivation has been reported previously for
Glycerol conversion
supported heteropoly acids [9–12], as well as for other solid acid
catalysts [8]. With CsPW, despite catalyst deactivation, the
3. Results and discussion selectivity to acrolein remained high for a long time, 92% after
12 h on stream. Acetol was the main by-product, but with only 2–
3.1. Glycerol dehydration over CsPW 3% selectivity. Amongst other by-products were found acetone,
acetaldehyde, propanal and propanoic acid together with traces of
The first entry in Table 2 shows the performance of CsPW bulk unidentified compounds.
heteropoly salt in glycerol dehydration in the gas phase at an Cs2HPW12O40, possessing smaller surface area than CsPW
optimum temperature of 275 8C and 1 bar pressure. Previously, the (Table 1), showed the same high acrolein selectivity, but at
same optimum temperature has been established for supported considerably lower glycerol conversion, decreasing from 47% after
heteropoly acids [9,10]. It should be noted that the reaction was 1 h to barely 4% after 3 h on stream (not shown in Table 2). Acidic
carried out in the presence of a large excess of steam as a result of caesium 12-tungstosilicate, Cs3.5H0.5SiW12O40, performed quite
using 10 wt% glycerol aqueous solution as a feed. The molar ratio of similar to CsPW, however with a lower glycerol conversion and
[glycerol]:[N2]:[H2O] in the incoming gas flow was 1:4.4:46 and faster catalyst deactivation (Table 2).

Table 2
Catalyst performance in dehydration of glycerol.a.

Catalyst Gas flow Conversionb (%) Selectivityb (%) Cc (wt%)


d
Acrolein Acetol Others

CsPW N2 41 (100) 94 (98) 2.4 3.6 8.9


Cs3.5H0.5 SiW12O40 N2 21 (91) 94 (97) 3.3 2.7
0.5%Ru/CsPW H2 51 (100) 94 (97) 2.8 3.2 6.8
0.3%Pt/CsPW H2 47 (100) 95 (87) 2.5 2.5 8.0
0.5%Pd/CsPW H2 79 (97) 96 (83) 1.4 2.6 4.5
5%Ru/C + CsPW (1:10)e H2 50 (95) 92 (94) 3.5 4.5
10%Pt/C + CsPW (1:20)e H2 74 (100) 5.8 (0) 3.3 15
10%Pd/C + CsPW (1:20)e H2 83 (100) 31 (4) 2.9 52
20%HPW/SiO2 N2 60 (100) 95 (97) 3.6 1.4
2%Pd/20%HPW/SiO2 H2 72 (100) 94 (85) 3.5 1.5
a
Reaction conditions: 275 8C, 0.30 g catalyst powder of 45–180 mm particle size, 15 mL min1 gas flow rate (N2 or H2), 10 wt% glycerol aqueous solution fed at
0.14 mL min1 flow rate (2.8 h1 glycerol WHSV), 5 h time on stream.
b
In brackets, glycerol conversion and acrolein selectivity after 1 h on stream.
c
Carbon content in used catalysts after reaction (5–6 h on stream) from combustion analysis.
d
A mixture of organic by-products including acetaldehyde, acetone, propanal, propanoic acid and other unidentified components.
e
Mechanical mixtures of CsPW with PGM/C in weight ratios shown in brackets with a total PGM content of 0.5 wt%. 75–90% and 14–50% loss of carbon balance was
observed for the Pt and Pd catalyst respectively, indicating the formation of gaseous products (CO and CO2). No attempt was made to investigate it in a more detail.
14 A. Alhanash et al. / Applied Catalysis A: General 378 (2010) 11–18

catalysts, suffers from rapid deactivation due to coke formation.


Regeneration of heteropoly acid catalysts has been reviewed
recently [20]. Traditional regeneration by coke combustion, which
is typically carried out at 500 8C, is difficult because of insufficient
thermal stability of heteropoly acids. Doping heteropoly acid
catalysts with platinum group metals (PGM) such as Pd and Pt
improves their regeneration by reducing the combustion temper-
ature to 350 8C [20]. However, the disadvantage of such
regeneration is that the catalytic process must be discontinued.
Here we attempted to enhance catalyst lifetime and reduce
coke deposition by doping CsPW with PGM such as Ru, Pd and Pt
and co-feeding hydrogen into the reaction system. This approach
has been effectively used in the industrial hydroisomerisation of
alkanes, which is carried out on PGM-doped alumina or zeolite in
the presence of hydrogen [24,25]. The metal additives, on the one
hand, promote bifunctional metal–acid mechanism of alkane
Fig. 1. Time course for glycerol dehydration over CsPW at 275 8C in N2 flow.
isomerisation and, on the other, reduce catalyst coking by
hydrogenating coke precursors. Solid Keggin heteropoly acids
Within the Misono’s classification of mechanisms of heteropoly such as H3PW12O40 doped with Pd and Pt have been reported as
acid catalysis (i.e., bulk-type and surface-type mechanisms) [21], the bifunctional catalysts for alkane hydroisomerisation [34–37].
dehydration of glycerol on the water-insoluble and rather hydro- Two types of PGM-doped CsPW catalysts were tested: (i)
phobic CsPW can be viewed as a surface-type acid-catalysed catalysts obtained by impregnation of CsPW with a PGM precursor
reaction occurring on the accessible surface proton sites of CsPW. followed by reduction with H2 and (ii) mechanical mixtures of
From the results in Table 2, we can estimate the turnover frequency CsPW with carbon-supported PGM (5%Ru/C, 10%Pt/C and 10%Pd/
(TOF) for glycerol conversion on this catalyst using SBET = 130 m2 g1 C). The first type had 0.3–0.5 wt% metal loading and the second
for the fresh catalyst (Table 1) and assuming a Keggin unit cross 0.5 wt%. Representative results on their performance in glycerol
section of 144 Å2 [21,22]. This gives an initial TOF of 405 h1 per one dehydration are shown in Table 2.
surface proton site at 1 h time on stream. After 5 h on stream, this First, the supported PGM catalysts (0.5%Ru/CsPW, 0.3%Pt/CsPW
value decreases to 166 h1 due to catalyst deactivation. These are and 0.5%Pd/CsPW) were tested using the same procedure as for the
approximate figures because the catalyst surface area was shrinking undoped CsPW above, i.e., at 275 8C in N2 flow. Under such
during reaction down to 97 m2 g1 after 5 h on stream, which would conditions, their performance was similar to that of the undoped
lead to higher TOF values. The corresponding values of the specific CsPW. Then they were tested in H2 flow, whereupon they showed
rate of acrolein production are 30 and 12 mmol h1 gcat1 for 1 and considerable improvement in their stability to deactivation
5 h time on stream respectively. These compare well with the compared to CsPW (Table 2, Fig. 2). This improvement, however,
best values reported for supported heteropoly acid catalysts: strongly depended on the metal dopant. Thus the Ru and Pt catalysts
5.1 mmol h1 gcat1 at 275 8C and 5 h time on stream for exhibited a relatively small enhancement at longer times on stream
30%H4SiW12O40/SiO2 and 10.5 mmol h1gcat1 at 315 8C and (cf. Figs. 1 and 2(A and B)). In contrast, the Pd catalyst, 0.5%Pd/CsPW,
9–10 h time on stream for 10%H3PW12O40/ZrO2 [11,12], which are provided significant enhancement of catalyst performance
calculated per total catalyst mass. (Fig. 2(C)). It is important that PGM doping had little (Pd and Pt)
In contrast to CsPW, heteropoly acids, for example the parent or no effect (Ru) on acrolein selectivity (Table 2, Fig. 2). It should be
acid H3PW12O40, are readily water-soluble and highly hydrophilic noted, however, that in the initial stage of reaction (1 h time on
compounds [21–23]. Therefore, the above heteropoly acid stream), Pd and Pt reduced the selectivity to acrolein to 83% and 87%
catalysts are most likely to dehydrate glycerol via the bulk-type respectively, which then increased to 95–96% (Fig. 2(B and C)). It was
mechanism by the Misono’s classification [21], which would mean found that the drop in selectivity was due to an increased formation
that all heteropoly acid protons are available for reaction. With of propanoic acid in the initial stage of reaction.
that in mind, the following TOF values for glycerol conversion were The PGM doping combined with H2 supply was found to reduce
calculated for these catalysts: 14.3 h1 for 30%H4SiW12O40/SiO2 the amount of coke deposited on the catalyst surface (Table 2),
(275 8C, 5 h) and 148 h1 for 10%H3PW12O40/ZrO2 (315 8C, 9–10 h). which is in line with the enhancement of catalyst stability to
These values are lower than our TOF value 166 h1 (275 8C, 5 h) for deactivation. Thus Ru and Pt had only small effect on the amount of
CsPW. This result can be explained by suggesting that under our coke formed, whereas Pd had pronounced effect, reducing the
reaction conditions in the presence of large excess of steam the amount of coke in half, down to 4.5%, which constitutes only 0.8%
hydrophobic CsPW has less hydrated hence stronger proton sites of carbon supplied into the reactor with the feed. This shows that
as compared to the hydrophilic heteropoly acids. It should be the enhancing effect of PGM in glycerol dehydration is related to
pointed out that in the absence of water CsPW is a weaker acid than their coke inhibition, with Pd being the most effective dopant.
the bulk parent H3PW12O40 as measured by ammonia adsorption Fig. 3 shows the TGA/TPO (temperature-programmed oxida-
calorimetry and ammonia TPD [23,33], with the values of initial tion) of the spent 0.5%Pd/CsPW catalyst after 6 h on stream in H2 at
NH3 adsorption enthalpy 165 and 195 kJ mol1 and NH3 275 8C and, for comparison, the TGA for the fresh catalyst. The fresh
desorption temperature 557 and 577 8C for CsPW and catalyst looses hydration water below 200 8C and exhibits no
H3PW12O40 respectively [23]. It should be noted, however, that further weight loss up to 620 8C. The TPO for the spent catalyst
supported heteropoly acids have weaker acid sites compared to provides an insight into the nature of coke formed. The TPO profile
bulk ones [22]. shows a broad combustion peak in the temperature range of 350–
550 8C, indicating that this coke is a mixture of aliphatic (easier to
3.2. Glycerol dehydration over PGM-doped CsPW burn) and polyaromatic (more difficult to burn) carbonaceous
deposits [38].
As demonstrated above, the bulk CsPW catalyst shows high Therefore, doping CsPW with PGM together with co-feeding H2
initial activity in glycerol dehydration, but, like other solid acid improves catalyst stability to deactivation without impairing the
A. Alhanash et al. / Applied Catalysis A: General 378 (2010) 11–18 15

Fig. 4. DRIFT spectra for (a) fresh and (b) spent 0.5%Pd/CsPW catalyst.

high selectivity to acrolein. The enhancing effect of PGM increases


in the order: Ru  Pt < Pd. The catalyst 0.5%Pd/CsPW gives 96%
acrolein selectivity at 79% glycerol conversion (76% yield) at 275 8C
and 5 h time on stream. It performs with an average TOF of 320 h1
(estimated as for the undoped CsPW above using SBET = 84 m2 g1
for the fresh catalyst). Again, this is an approximation because the
catalyst surface was shrinking during reaction down to 68 m2 g1
after 5 h on stream, which would lead to a higher TOF value. For
this catalyst, the specific rate of acrolein production amounts to
23 mmol h1 gcat1 at 5 h on stream, exceeding that reported for
supported heteropoly acids.
Fresh and spent 0.5%Pd/CsPW catalysts were characterised by
FTIR and XRD to reveal that the Keggin (primary) structure of the
PW12O403 anion as well as the crystal (secondary) structure of
CsPW remained unchanged after reaction. Both fresh and spent
catalysts exhibited the same well-known infrared spectrum
characteristic of the Keggin structure (Fig. 4) [21]. Likewise, the
fresh CsPW and 0.5%Pd/CsPW as well as the spent 0.5%Pd/CsPW
exhibited the same cubic XRD pattern characteristic of CsPW
(Fig. 5). The XRD of CsPW has been studied in detail in comparison
with that of the parent acid H3PW12O40 by Okuhara et al. [39]. It
should be noted that the XRD of Pd/CsPW, both fresh and used, did
not show any pattern of Pd metal. Thus the most intense 1 1 1
diffraction peak of Pd phase at 2Q = 40.1 is not seen in Fig. 5. This
Fig. 2. Time course for glycerol dehydration over (A) 0.5%Ru/CsPW, (B) 0.3%Pt/CsPW
indicates a fine dispersion (D) of Pd metal in the catalyst in
and (C) 0.5%Pd/CsPW at 275 8C in H2 flow. agreement with our measurement by hydrogen chemisorption
(D = 0.76, Table 1). This value corresponds to an average Pd particle
size d = 0.9/D = 1.2 nm.
The performance of catalysts prepared by mechanical mixing of
CsPW with carbon-supported PGM is also presented in Table 2. The
Ru catalyst, 5%Ru/C + CsPW (1:10), showed very similar results to

Fig. 3. TGA for the 0.5%Pd/CsPW catalyst in air flow: (a) spent catalyst (TG mode), (b) Fig. 5. Powder XRD (Cu Ka radiation) for (a) fresh CsPW, (b) fresh 0.5%Pd/CsPW and
spent catalyst (DTG mode, arbitrary units) and (c) fresh catalyst (TG mode). (c) spent 0.5%Pd/CsPW.
16 A. Alhanash et al. / Applied Catalysis A: General 378 (2010) 11–18

those for the supported Ru catalyst 0.5%Ru/CsPW, albeit with a catalyst 2%Pd/20%HPW/SiO2, similarly to 0.5%Pd/CsPW, showed
slightly lower glycerol conversion and acrolein selectivity. The Pt considerable improvement regarding its stability to deactivation
and Pd catalysts behaved quite differently, however. These two as compared to 20%HPW/SiO2 (Fig. 6). At 5 h on stream, it gave 94%
catalysts gave high glycerol conversions, but very low acrolein acrolein selectivity at 72% conversion (68% yield), with an acrolein
selectivities, especially in the initial stage of reaction. Instead, production rate of 21 mmol h1 gcat1 (105 mmol h1 gHPW1) and
these catalysts produced large amounts of propanoic acid, which a TOF value of 106 h1. Although these results are not as good as
was identified by GC–MS. Moreover, significant loss of carbon those for 0.5%Pd/CsPW (see above), they are still better than those
balance was observed: 75–90% and 14–50% for the Pt and Pd reported for supported heteropoly acids previously [11,12].
catalysts respectively, depending on the time on stream. Larger It should be noted that Pt is usually more active than Pd in
carbon losses were observed in the initial stage of reaction, alkane hydroisomerisation over bifunctional catalysts. Therefore,
decreasing with the time on stream. Since the catalyst weight did the question arises why is Pd more active than Pt in our case? The
not change significantly after reaction, these carbon losses cannot reason for this could be that the nature of coke is different in these
be explained by coking. It can be suggested that these catalysts reactions. In alkane hydroisomerisation, coke is mainly formed by
produced gaseous products such as H2, CO and CO2 by steam olefin polymerisation, hence it has a hydrocarbon composition. In
reforming of glycerol. This would be similar to aqueous phase glycerol to acrolein dehydration, coke is likely to form by acrolein
reforming (APR) of polyols documented by Dumesic and co- polymerisation and, therefore, comprises oxygenated hydrocar-
workers [6]. Pt and Pd are active catalysts in the APR, with Pt being bons. It is conceivable, therefore, that Pd is more effective than Pt in
the most active one [6]. However, as regards to the dehydration of inhibiting the formation of oxygenated coke.
glycerol to acrolein, these catalysts have no advantage over the
supported PGM catalysts discussed above, therefore no attempt 3.3. Mechanistic considerations
was made to study them in a more detail.
For comparison, we also tested silica-supported H3PW12O40 It has been suggested that the dehydration of glycerol to
(HPW) catalysts 20%HPW/SiO2 and 2%Pd/20%HPW/SiO2 under the acrolein proceeds through the formation of 3-hydroxypropanal
same conditions (Table 2, Fig. 6). The performance of the undoped and acetol as the primary products, with the former further
catalyst in N2 at 275 8C was similar to that of CsPW regarding dehydrated to acrolein, whereas the latter a relatively stable by-
acrolein selectivity and even better regarding glycerol conversion: product (Scheme 1) [8–14]. We tested the reactivity of acetol in our
60% conversion at 5 h compared to 41% for CsPW. As a result, conditions at 275 8C passing 5 wt% aqueous solution of acetol over
20%HPW/SiO2 gave a higher specific rate of acrolein production CsPW in N2 flow and over 0.5%Pd/CsPW in H2 flow (acetol
of 17 mmol h1 gcat1 (5 h) per total catalyst mass and WHSV = 1.4 h1). Neither of these tests showed any conversion of
85 mmol h1 gHPW1 per HPW mass. Assuming that all protons acetol, thus confirming that acetol indeed is a relatively stable by-
of HPW were available for the reaction, the TOF values were product in our system. Notably, we did not observe any
calculated to be 147 h1 at 1 h and 88 h1 at 5 h. The Pd-doped hydrogenation of acetol to 1,2-propanediol with Pd/CsPW,
although this reaction has been reported to occur in other systems
[15]. This result can be explained by the large excess of steam in
our system. Another conclusion that can be drawn from the lack of
acetol conversion is that in our system coke is mainly originated
from acrolein and maybe to some extent from 3-hydroxypropanal
(Scheme 2).
Previous work on glycerol dehydration suggests that acrolein
forms on strong Brønsted acid sites [8–14]. Consequently, tungsten
heteropoly acids such as H3PW12O40 and H4SiW12O40 and the
CsPW heteropoly salt presented here, which are all strong, purely
Brønsted acids, are amongst the most efficient catalysts for the
synthesis of acrolein. In this context, it was interesting to test a
purely Lewis solid acid catalyst for the synthesis of acrolein. As
such, we tested Zn(II)–Cr(III) (1:1) mixed oxide and its Pd-doped
derivative 0.3%Pd/Zn–Cr. Recently, these catalysts have been found
active for dehydroisomerisation of a-pinene to p-cymene [40] and
hydrogenation of acetone to methyl isobutyl ketone [29], and the
acid properties of Zn–Cr oxide have been characterised in detail
[29].
Fig. 7 shows the DRIFT spectra of pyridine adsorbed on CsPW
and Zn–Cr (1:1) oxide, which give clear evidence of the nature of
acid sites in these catalysts. As expected, the CsPW possesses
Brønsted acid sites as indicated by the strong band at 1540 cm1. In
contrast, the Zn–Cr oxide possesses only Lewis acid sites as

Fig. 6. Time course for glycerol dehydration at 275 8C over (A) 20%HPW/SiO2 in N2
flow and (B) 2%Pd/20%HPW/SiO2 in H2 flow. Scheme 2. Reaction pathway for glycerol dehydration to acrolein.
A. Alhanash et al. / Applied Catalysis A: General 378 (2010) 11–18 17

Fig. 7. DRIFT spectra of pyridine adsorbed on (a) Zn–Cr (1:1) mixed oxide and (b)
CsPW.

Table 3
Glycerol dehydration over Zn–Cr (1:1) mixed oxide catalyst.a.

Catalyst Gas Temperature Conversion Selectivity (%)


flow (8C) (%)
Acrolein Acetol Others

Zn–Cr N2 275 1
Zn–Cr N2 300 9 (13) 34 (38) 42 (40) 24
Zn–Cr N2 350 18 (25) 30 (30) 40 (32) 30
Fig. 8. Time course for glycerol dehydration at 350 8C over (A) Zn–Cr (1:1) in N2 flow
0.3%Pd/Zn–Cr H2 350 49 (44) 32 (23) 31 (22) 37
and (B) 0.3%Pd/Zn–Cr (1:1) in H2 flow.
Reaction conditions as in Table 2. Glycerol conversion and product selectivity are
after 5 h time on stream, in brackets after 1 h on stream.

evidenced by the strong band at 1450 cm1. These Lewis sites have
been found to be of moderate acid strength with an enthalpy of
NH3 adsorption of 155 kJ mol1 [29].
Table 3 and Fig. 8 show the performance of Zn–Cr and 0.3%Pd/
Zn–Cr catalysts in acrolein dehydration. In contrast to CsPW, these
catalysts were not active at 275 8C due to their weaker acidity, but
showed a moderate activity at 300–350 8C, with good performance
stability for 5–6 h on stream (Fig. 8). The important result is that
these catalysts gave acetol as the main product with the selectivity
(31–42%) greater than, or equal to, that of acrolein (30–34%). It is
logical, therefore, to suggest that acetol was formed on Lewis acid Scheme 3. Mechanism of glycerol dehydration on Brønsted acid sites.
sites whereas acrolein on Brønsted acid sites. Although Brønsted
sites were not available in the fresh Zn–Cr oxide, they could be
generated by interaction of the Lewis sites with steam present in proton site should mainly lead to the protonation of the internal
the system to give rise to the acrolein formation. oxygen in the glycerol molecule possessing a higher negative
We propose the following mechanistic schemes to explain charge compared to the terminal oxygens (Scheme 3). Subsequent
these results. Scheme 3 represents glycerol dehydration on strong steps involve elimination of H3O+ to give 1,3-dihydroxypropene
proton sites and Scheme 4 on Lewis sites. Since proton transfer is and its tautomerisation to 3-hydroxypropanal. The latter under-
not limited by steric constrains, interaction of glycerol with a goes further acid-catalysed dehydration to yield acrolein. Finally,

Scheme 4. Mechanism of glycerol dehydration on Lewis acid sites.


18 A. Alhanash et al. / Applied Catalysis A: General 378 (2010) 11–18

the proton site is regenerated by interaction of its conjugate base Acknowledgement


with H3O+. Similar mechanism for acrolein formation has been
suggested previously [8–14]. Financial support from the EPSRC (research grant EP/E039847)
Interaction of glycerol with Lewis acid sites occurs differently is gratefully acknowledged.
as indicated by our results with Zn–Cr oxide. This interaction, in
contrast to the glycerol protonation above, is bound to be affected References
by steric constrains. As a result, it is the terminal OH group of
glycerol rather than the internal one that is more likely to interact [1] A. Corma, S. Iborra, A. Velty, Chem. Rev. 107 (2007) 2411–2501.
with Lewis acid site as shown in Scheme 4, where two oxo-bridged [2] M. Pargliaro, R. Ciriminna, H. Kimura, M. Rossi, C. DellaPina, Angew. Chem. Int. Ed.
46 (2007) 4434–4440.
metal ions (M) represent the Lewis acid site in the Zn–Cr oxide.
[3] C. Zhou, J.N. Beltramini, Y. Fan, G.Q. Lu, Chem. Soc. Rev. 37 (2008) 527–549.
Concerted transfer of the terminal OH group of glycerol to the M [4] Y. Zheng, X. Chen, Y. Shen, Chem. Rev. 108 (2008) 5253–5277.
and migration of the H+ from the internal carbon atom to the [5] A. Behr, J. Eilting, K. Irawadi, J. Leschinski, F. Lindner, Green Chem. 10 (2008) 13–
bridging O atom of the oxide gives 2,3-dihydroxypropene 30.
[6] G.W. Huber, J.W. Shabaker, J.A. Dumesic, Science 300 (2003) 2075–2077.
together with the hydrated active site in the catalyst. The 2,3- [7] K. Weissermel, H.J. Arpe, Industrial Organic Chemistry, 3rd ed., Wiley, 1997.
dihydroxypropene is then tautomerised to yield acetol, and the [8] B. Katryniok, S. Paul, M. Capron, F. Dumeignil, Chem. Sus. Chem. 2 (2009) 719–730.
Lewis acid site is regenerated by thermal dehydration of its [9] E. Tsukuda, S. Sato, R. Takahashi, T. Sodesawa, Catal. Commun. 8 (2007) 1349–
1353.
hydrated form. [10] H. Atia, U. Armbruster, A. Martin, J. Catal. 258 (2008) 71–82.
The proposed mechanisms (Schemes 3 and 4) adequately [11] S.H. Chai, H.P. Wang, Y. Liang, B.Q. Xu, Green Chem. 10 (2008) 1087–1093.
represent the dehydration of glycerol on Brønsted and Lewis acid [12] S.H. Chai, H.P. Wang, Y. Liang, B.Q. Xu, Appl. Catal. A 353 (2009) 213–222.
[13] W. Suprun, M. Lutecki, T. Haber, H. Papp, J. Mol. Catal. A 309 (2009) 71–78.
sites. New evidence presented here regarding the nature of acid [14] A. Corma, G.W. Huber, L. Sauvanaud, P. O’Connor, J. Catal. 257 (2008) 163–171.
sites required for the dehydration of glycerol to acrolein supports [15] C.-W. Chiu, M.A. Dasari, G.J. Suppes, W.R. Sutterlin, AIChE J. 52 (2006) 3543–3548.
the importance of strong Brønsted sites for this reaction. [16] S. Sato, M. Akiyama, R. Takahashi, T. Hara, K. Inui, M. Yokota, Appl. Catal. A 347
(2008) 186–191.
[17] J.L. Dubois, C. Duquenne, W. Hoelderich, J. Kervennal, WO 2006087084 (2006).
4. Conclusion [18] H. Kasuga, M. Okada, JP 2008137950 (2008).
[19] Y. Arita, H. Kasuga, M. Kirishiki, JP 2008110298 (2008).
[20] I.V. Kozhevnikov, J. Mol. Catal. A 305 (2009) 104–111.
This study has demonstrated that the water-insoluble Cs
[21] T. Okuhara, N. Mizuno, M. Misono, Adv. Catal. 41 (1996) 113–252.
heteropoly salt, Cs2.5H0.5PW12O40 (CsPW), possessing strong [22] I.V. Kozhevnikov, Chem. Rev. 98 (1998) 171–198.
Brønsted acid sites and high water tolerance is an active catalyst [23] T. Okuhara, Chem. Rev. 102 (2002) 3641–3666.
for the dehydration of glycerol to acrolein in the gas-phase process [24] J.A. Moulijn, P.W.N.M. van Leeuwen, R.A. van, Santen (Eds.), Catalysis, Elsevier,
Amsterdam, 1993.
at 275 8C and 1 bar pressure. The catalyst exhibits high initial [25] A. Chica, A. Corma, J. Catal. 187 (1999) 167–176.
activity, with a glycerol conversion of 100% at 98% acrolein [26] Y. Izumi, M. Ono, M. Kitagawa, M. Yoshida, K. Urabe, Micropor. Mater. 5 (1995)
selectivity. The activity, however, decreases significantly with the 255–262.
[27] A. Alhanash, E.F. Kozhevnikova, I.V. Kozhevnikov, Catal. Lett. 120 (2008) 307–311.
time on stream due to catalyst coking, without impairing acrolein [28] E.F. Kozhevnikova, E. Rafiee, I.V. Kozhevnikov, Appl. Catal. A 260 (2004) 25–34.
selectivity. It has been found that doping CsPW with platinum [29] F. Al-Wadaani, E.F. Kozhevnikova, I.V. Kozhevnikov, J. Catal. 257 (2008) 199–205.
group metals (PGM) together with co-feeding hydrogen improves [30] K. Katamura, T. Nakamura, K. Sakata, M. Misono, Y. Yoneda, Chem. Lett. (1981)
89–92.
catalyst stability to deactivation, while maintaining high selectivi- [31] J.E. Benson, M. Boudart, J. Catal. 4 (1965) 704–710.
ty to acrolein. The enhancing effect of PGM increases in the order: [32] J.E. Benson, H.S. Hwang, M. Boudart, J. Catal. 30 (1973) 146–153.
Ru  Pt < Pd. The catalyst 0.5%Pd/CsPW performs with a turnover [33] E.F. Kozhevnikova, I.V. Kozhevnikov, J. Catal. 224 (2004) 164–169.
[34] T. Baba, Y. Hasada, M. Nomura, Y. Ohno, Y. Ono, J. Mol. Catal. A 114 (1996) 247–
frequency of 320 h1 at 275 8C and 5 h time on stream and 255.
gives 96% acrolein selectivity at 79% glycerol conversion (76% [35] T. Okuhara, Appl. Catal. A 256 (2003) 213–224.
yield). It provides a specific rate of acrolein production of [36] N. Essayem, Y. Ben Taârit, P.Y. Gayraud, G. Sapaly, C. Naccache, J. Catal. 204 (2001)
157–162.
23 mmol h1 gcat1, which exceeds the specific rates reported
[37] N. Essayem, Y. Ben Taârit, C. Feche, P.Y. Gayraud, G. Sapaly, C. Naccache, J. Catal.
in the literature for supported Keggin heteropoly acids (5– 219 (2003) 97–106.
11 mmol h1 gcat1 per total catalyst mass). New mechanistic [38] J. Barbier, Stud. Surf. Sci. Catal. 34 (1987) 1–18.
evidence has been obtained regarding the nature of acid sites [39] T. Okuhara, H. Watanabe, T. Nishimura, K. Inumaru, M. Misono, Chem. Mater. 12
(2000) 2230–2238.
required for the dehydration of glycerol to acrolein, which supports [40] F. Al-Wadaani, E.F. Kozhevnikova, I.V. Kozhevnikov, Appl. Catal. A 363 (2009)
the importance of strong Brønsted sites for this reaction. 153–156.

You might also like