You are on page 1of 7

Catalysis Today 212 (2013) 120–126

Contents lists available at SciVerse ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Hydrogenolysis of glycerol to 1,3-propanediol over bifunctional catalysts


containing Pt and heteropolyacids
Shanhui Zhu a,b , Yanan Qiu c , Yulei Zhu a,d,∗ , Shunli Hao d , Hongyan Zheng d , Yongwang Li a,d
a
State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, PR China
b
Graduate University of Chinese Academy of Sciences, Beijing 100039, PR China
c
Institute of Science & Technology Information of Shanxi, Taiyuan 030001, PR China
d
Synfuels China Co. Ltd., Taiyuan 030032, PR China

a r t i c l e i n f o a b s t r a c t

Article history: Hydrogenolysis of glycerol to 1,3-propanediol was conducted over zirconia supported bifunctional cata-
Received 11 March 2012 lysts containing Pt and heteropolyacids using H4 SiW12 O40 (HSiW), H3 PW12 O40 (HPW) and H3 PMo12 O40
Received in revised form 27 June 2012 (HPMo) as active compounds. Pt/ZrO2 was also examined for comparison. The as-prepared catalysts were
Accepted 2 September 2012
characterized by BET, CO chemisorption, XRD, Raman spectra, NH3 -TPD and FTIR of adsorbed pyridine.
Available online 18 October 2012
Compared with Pt/ZrO2 , heteropolyacids modified Pt/ZrO2 catalysts showed higher acidity and better
catalytic performance of glycerol hydrogenolysis to 1,3-propanediol. Among them, Pt-HSiW/ZrO2 exhib-
Keywords:
ited superior performance due to the high Brønsted acid sites and good thermal stability. Independent
Glycerol
Hydrogenolysis of the heteropolyacid type, the concentration of Brønsted acid sites appeared as a key to the selective
1,3-Propanediol formation of 1,3-propanediol from glycerol hydrogenolysis, whereas the concentration of Lewis acid sites
Heteropolyacids was related to the formation of 1,2-propanediol. We also investigated the reaction network and proposed
a possible reaction pathway.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction synthesis via hydration of acrolein or hydroformylation of ethyl-


ene oxide [12]. The development of an alternative environmentally
The catalytic transformation of biomass-derived feedstocks for benign process for the production of 1,3-PDO based on the biomass-
the synthesis of value-added fuels, chemicals and materials has derived glycerol as a sustainable feedstock is a real challenge.
been attracting much attention, due to the diminishing reserves Supported Ir, Pt, Ru, Rh, Pd and Cu catalysts have been used
and rising price of fossil resources [1,2]. Recent development of for glycerol hydrogenolysis to 1,3-PDO [9,10,13–21]. Among these
biodiesel production by transesterification of plant oils and animal catalysts, supported Pt catalysts are more selective for this reac-
fat has made a large amount of glycerol available as a renewable tion than others. To obtain 1,3-PDO effectively, it is necessary
feedstock for the production of various fuels and valuable chemicals to add acid component to the catalysts due to the bi-functional
[2–5]. Significant research efforts have been focused on the conver- mechanism. Previous research [9,22] shows that hydrogenoly-
sion of glycerol by various catalytic processes such as reforming, sis of glycerol to 1,3-PDO proceeds via dehydration of glycerol
oxidation, dehydration, hydrogenolysis, etherification, esterifica- to 3-hydroxypropaldehyde (3-HPA) followed by hydrogenation.
tion and so on [3–10]. According to the above mechanism, acid catalysts play an important
Hydrogenolysis of glycerol to 1,2-propanediol (1,2-PDO) and role in glycerol hydrogenolysis. Various acid catalysts including
1,3-propanediol (1,3-PDO) is a noticeable pathway due to its H2 WO4 , H2 SO4 , sulfated zirconia, WO3 /ZrO2 , and WO3 /TiO2 /SiO2
production of renewable value-added chemicals. 1,3-PDO owes have been used [9,10,18,21,23,24]. Studies by Chaminand et al. [23]
much higher economical value than 1,2-PDO, in particular, as an showed that hydrogenolysis of glycerol to 1,3-PDO on Rh/SiO2 cat-
important monomer in the synthesis of polyester fibers [4,5,11]. alyst achieved 4% yield in the presence of H2 WO4 at 200 ◦ C and
Industrial production of 1,3-PDO is currently based on chemical 8.0 MPa. Tomishige and co-workers [18] reported that the yield of
1,3-PDO reached 38% over Ir-ReOx /SiO2 catalyst with H2 SO4 as an
additive in a batch reactor. However, the liquid acids would cause
several problems in terms of catalyst separation, reactor corro-
∗ Corresponding author at: State Key Laboratory of Coal Conversion, Institute of
sion and environmental protection. Recently, Oh et al. [24] have
Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, PR China.
Tel.: +86 351 7117097; fax: +86 351 7560668. communicated the use of Pt-sulfated zirconia with 1,3-dimethyl-2-
E-mail address: zhuyulei@sxicc.ac.cn (Y. Zhu). imidazolidinone in glycerol hydrogenolysis but such strong acidic

0920-5861/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cattod.2012.09.011
S. Zhu et al. / Catalysis Today 212 (2013) 120–126 121

catalysts deactivate quickly as carbonaceous deposits block the cat- 200 ◦ C, then purged with He at the same temperature for 1 h and
alyst surface. Additionally, the usage of organic solvent will greatly finally cooled down to 50 ◦ C. The CO chemisorption was operated
reduce the environmental and economical viability. by pulse injection of pure CO at 50 ◦ C. The Pt particle size was cal-
Heteropolyacids (HPAs) have demonstrated to display out- culated by assuming an adsorption of one CO molecule per surface
standing catalytic performance for both homogeneous and platinum atom.
heterogeneous acid-catalyzed reactions because of its well-defined Powder X-ray diffraction (XRD) patterns were conducted at
structure and widely tunable acidity [8,25–27]. HPAs possess room temperature on a D2/max-RA X-ray diffractometer (Bruker,
unique physicochemical properties, with their structural mobil- Germany) using Cu K␣ radiation at 30 kV and 10 mA. The X-ray
ity and multifunctionality, which is useful in industrial processes patterns were recorded in 2 values ranging from 10◦ to 90◦ at the
such as the liquid–phase dehydration of alcohols and hydration of scanning rate of 5◦ /min.
esters or olefins [26–28]. Additionally, heterogeneous acid catal- Raman spectra were recorded with a LabRAM HR800 System
ysis by HPAs has attracted much interest due to its potential of equipped with a CCD detector at room temperature. The 325 nm of
great economic rewards and green benefits. Compared to con- the He-Cd laser was used as the exciting source with a power of
ventional solid acid catalysts such as oxides and zeolites, HPAs 30 MW.
possess strong acidity, uniform acid sites and easily tunable acid- NH3 -TPD was carried out in the same apparatus as CO
ity. Generally, tungsten HPAs owns stronger acidity, higher thermal chemisorption. Prior to each run, the catalyst sample (0.3 g) was
stability and lower oxidation potential than molybdenum HPAs, pretreated in He at 350 ◦ C for 1 h, then cooled to 100 ◦ C and was
which makes tungsten HPAs used more widely [29]. However, HPAs saturated with pure NH3 for 30 min. After being purged with He
lack thermal stability and leach easily in polar solvents. To solve for 30 min, the sample was heated to 700 ◦ C at a heating rate
this problem, HPAs are often immobilized on supports such as of 10 ◦ C/min and the NH3 desorption was monitored with a TCD
SiO2 or ZrO2 . Kozhevnikov and co-workers [27] have studied the detector.
interaction between HPW and support with ammonia adsorption To evaluate and analyze the nature and acid amount of cata-
calorimetry, MAS NMR and FTIR and suggested increasing interac- lysts, FTIR spectra of adsorbed pyridine (Py-IR) of solid samples
tion in the following order: SiO2 < TiO2 < Nb2 O5 < ZrO2 . Moreover, were carried out on a VERTEX70 (Bruker) FT-IR spectrophotome-
our previous study [13] showed that the Pt-HSiW/ZrO2 catalyst ter, equipped with a deuterium triglycine sulphate (DTGS) detector.
was rather durable during one-step hydrogenolysis of glycerol to Prior to pyridine adsorption, the samples were heated to 300 ◦ C
biopropanols. Taken together, it is worthwhile to study supported under vacuum, and then cooled to 30 ◦ C. The Py-IR spectra were
HPAs as acid catalysts for glycerol hydrogenolysis. then recorded at 200 ◦ C after applying vacuum for 30 min. The
In the present investigation, several zirconia supported Pt- quantitative calculation of Brønsted and Lewis acid sites was deter-
HPAs (HSiW, HPW and HPMo) bi-functional catalysts were mined by the integrated area of the adsorption bands at about 1540
prepared, characterized, and tested in aqueous solution for glyc- and 1450 cm−1 , respectively. The acidity of Brønsted and Lewis
erol hydrogenolysis. Special emphasis was also laid on formation acid sites was calculated by using the adsorption coefficient values
route of glycerol hydrogenolysis and degradation based on the reported in the previous paper [30].
catalytic performance in the reaction of products and detected
intermediates. 2.3. Catalytic tests

Hydrogenolysis of glycerol was conducted in a vertical fixed-


2. Experimental
bed reactor (i.d. 12 mm, length 600 mm) with an ice-water trap. In
a typical run, 2.0 g catalyst (20–40 mesh) was charged in the con-
2.1. Catalyst preparation
stant temperature section of the reactor, with quartz sand packed
in both ends. Prior to the test, the catalyst sample was in situ
ZrO2 (Jiangsu Qianye Co., Ltd., China) was used as support of
reduced in a stream of pure H2 (100 ml/min) at 200 ◦ C for 2 h. After
the catalysts. H2 PtCl6 ·6H2 O (AR), HSiW (AR), HPW (AR) and HPMo
reduction, the system was cooled to the desired temperature. A
(AR) were purchased from Sinopharm Chemical Reagent Co., Ltd.
10 wt.% glycerol aqueous solution was fed continuously into the
All the catalysts were prepared by incipient wetness impregnation
reactor with an HPLC pump along with a co-feed H2 of gas flow-
method. Typical procedures for the preparation of Pt-HSiW/ZrO2
ing at 100 cm3 min−1 . The liquid and gas products were cooled and
catalyst are as follows. The Pt/ZrO2 catalyst was prepared by
collected in a gas–liquid separator immersed in an ice-water trap.
impregnation of zirconia support with an aqueous solution of
The products were obtained when the reaction reached the steady
H2 PtCl6 ·6H2 O. The impregnated sample was dried overnight at
state. Additionally, the reaction performance of 1,3-PDO, 1,2-PDO
110 ◦ C and then calcined at 400 ◦ C in static air for 4 h. After impreg-
and ethylene glycol (EG) was also evaluated in order to reveal the
nated HSiW onto Pt/ZrO2 , Pt-HSiW/ZrO2 was dried overnight at
reaction pathway of glycerol hydrogenolysis.
110 ◦ C and then calcined at 350 ◦ C in static air for 4 h. Pt-HPW/ZrO2
The liquid products were analyzed by gas chromatography using
and Pt-HPMo/ZrO2 were prepared in the similar way to that of
a capillary column (DB-WAX, 30 m × 0.32 mm) equipped with a FID
Pt-HSiW/ZrO2 . The Pt and HPAs loadings were fixed at 2 wt.% and
detector. The tail gas was off-line analyzed by a gas chromatogra-
15 wt.%, respectively.
phy with a capillary column (OV-101, 60 m × 0.25 mm) and a TCD
detector. The products obtained were also identified by GC–MS. The
2.2. Catalyst characterization conversion of glycerol and selectivity of products were calculated
as follows:
N2 adsorption–desorption isotherms were recorded at
Conversion (%)
−196 ◦ C using a Micromeritics ASAP 2420 instrument after
degassing at 300 ◦ C for 8 h in vacuum. BET surface area and moles of glycerol (in) − moles of glycerol (out)
= × 100
BJH pore size distribution were calculated by desorption moles of glycerol (in)
isotherms.
CO chemisorption was measured at 50 ◦ C in Auto Chem. 2920 moles of one product
equipment (Mircromeritics, USA). Prior to the measurement, about Selectivity (%) = × 100.
moles of all products
0.2 g catalyst sample was first reduced in situ for 2 h in pure H2 at
122 S. Zhu et al. / Catalysis Today 212 (2013) 120–126

Table 1
Physicochemical properties and acidities of Pt/ZrO2 and Pt-HPAs/ZrO2 catalysts.

Catalyst Surface area Pore size Pore volume Pt size (nm) Pt dispersion Total acidity Brønsted acidity Lewis acidity Total acidity
(m2 /g−1 ) (nm) (cm3 g−1 ) (%) (mmol NH3 /gcat )a (␮mol/gcat. )b (␮mol/gcat )b (␮mol/gcat )b

Pt/ZrO2 61.0 9.7 0.202 3.01 37.6 0.49 0 164.82 164.82


Pt-HSiW/ZrO2 54.8 9.8 0.164 3.16 35.8 0.53 86.56 85.59 172.15
Pt-HPW/ZrO2 57.0 9.4 0.171 3.06 37.0 0.63 62.89 117.18 180.07
Pt-HPMo/ZrO2 55.3 8.9 0.170 3.69 30.4 0.66 20.09 179.03 199.12
a
The amount of acid sites was determined by quantifying the desorbed NH3 from NH3 -TPD.
b
The amount of acid sites was determined by quantifying the desorbed pyridine from Py-IR.

3. Results and discussion

Pt-PMo/ZrO2
3.1. Catalyst characterizations

3.1.1. Physicochemical properties of catalysts


Table 1 lists the textural properties derived from nitrogen Pt-HPW/ZrO2
physisorption isotherms and dispersion of Pt determined by CO

Intensity
chemisorption of as-prepared catalysts. As can be seen, both surface
area and pore volume of Pt-HPAs/ZrO2 samples decreased as com- Pt-HSiW/ZrO2
pared to Pt/ZrO2 . This decrease may be attributed to pore blockage
caused by heteropolyacids (HPAs) as their Keggin units (diameter Pt/ZrO2
≈1.2 nm) are small enough to enter mircropores of ZrO2 [31]. As
shown in Table 1, all the catalysts have small Pt particle size and
good dispersion of Pt. Compared with Pt/ZrO2 sample, no remark- ZrO2
able changes of Pt particle size are observed for any supported HPAs
catalysts, implying that the HPAs have little effect on the disper-
sion of Pt. The slight decrease in Pt dispersion of Pt-HPMo/ZrO2 10 20 30 40 50 60
can be probably related to the partial coverage by HPMo or decom- 2θ (º)
position component. Cui et al. [32] have studied the effect of Mo
promoter in iron-based catalysts by high-resolution transmission Fig. 1. XRD patterns of bulk ZrO2 , Pt/ZrO2 and Pt-HPAs/ZrO2 catalysts.

electron microscopy and suggested that the presence of high Mo-


content would weaken the dispersion of iron active sites due to the below 700 cm−1 interfere with the characteristic Raman bands
coverage effect. of HPAs species, while the range above 700 cm−1 is free from
Assuming that a cross section of the heteropolyacid (HPA) characteristic bands of ZrO2 and hence relevant to the structure
molecule is 144 Å2 , the 15% HPA loading will correspond to determination of HPAs species. The typical Raman spectra bands of
45 m2 /g monolayer coverage [27]. Due to the 61.0 m2 /g surface Pt-HSiW/ZrO2 appeared at 998 and 975 cm−1 , which were assigned
area of Pt/ZrO2 , the supported HPA catalysts can be adequately to W O symmetric and asymmetric stretching vibrations of HSiW
viewed as having sub-monolayer HPA coverage. The XRD patterns Keggin structure, respectively [13]. The Pt-HPW/ZrO2 sample also
(discussed below) confirmed the high dispersion of HPA on the sup- displayed two overlapping bands at around 1008 and 990 cm−1 ,
port surface. The sub-monolayer HPA coverage can ensure strong which can be ascribed to the symmetric and asymmetric stretch-
HPA/support interaction which can strengthen the thermal stabil- ing vibrations of the terminal W O bond of HPW Keggin anions,
ity and avoid the leaching of HPA during the reaction. Additionally, respectively [33]. For Pt-HPMo/ZrO2 , Raman band for the Keggin
the 15 wt.% loading is sufficient to provide efficient acid catalysis for structure was observed at 996 cm−1 but a weak band characteris-
glycerol hydrogenolysis. Chai et al. [33] have investigated glycerol tic of MoO3 was also identified at 818 cm−1 [31]. Furthermore, a
dehydration over a series of HPW loading on zirconia and found
that the 15 wt.% HPW/ZrO2 showed superior catalytic performance
and most of the Keggin HPW remained intact after calcination up
to 650 ◦ C. HPMo
1/4
The XRD patterns of bulk ZrO2 and as-prepared catalysts are
illustrated in Fig. 1. ZrO2 presented in all the samples exhibited
Pt-HPMo/ZrO2
characteristic peaks corresponding to monoclinic phase of ZrO2
(e.g., 24.1◦ , 28.3◦ , 31.5◦ , 34.3◦ , 49.4◦ ) [25]. It is of interest to note
that neither diffraction peaks of Keggin structure nor decomposi- HPW 1/4
Intensity

tion products WO3 or MoO3 were detected, most probably because


Pt-HPW/ZrO2
of the fine dispersion of HPAs or their decomposition species on
the support surface. As discussed above, the sub-monolayer HPA
coverage can avoid HPA crystallites on the support surface. On the HSiW 1/4
other hand, there were also no diffraction peaks of Pt, indicating Pt-HSiW/ZrO2
that Pt was highly dispersed on the support surface, which was
well in accordance with the CO chemisorption results. ZrO2
Raman spectroscopy is known to be well suited for observa-
tion the state of HPAs in the catalysts. As shown in Fig. 2, it has 200 400 600 800 1000 1200
been identified and attributed Raman bands at 222, 310, 337, -1
Raman shift (cm )
382, 474, 499, 540, 559, 620 and 636 cm−1 to monoclinic zirco-
nia [25], as evidenced by XRD results. The strong Raman bands Fig. 2. Raman spectra of bulk heteropolyacids and different catalysts.
S. Zhu et al. / Catalysis Today 212 (2013) 120–126 123

broad band in the range of 980–750 cm−1 was detected, which was
assigned to the stretching vibrations of the bridges between tri-
L B
molybdic groups of Keggin structure caused by dehydration process
[31]. The broadening of this region and presence of MoO3 indi-
cated a partial decomposition of HPMo Keggin structure during the
high temperature calcination. These observations are consistent
with earlier literature that HPMo would undergo decomposition

Absorbance
Pt-HPMo/ZrO2
at around 350 ◦ C to form molybdenum and phosphorous oxides
[29,31].
Pt-HPW/ZrO2
In comparison with the bulk HPAs, a decrease in the intensity
and a broadening of the Keggin bands over supported HPAs is also Pt-HSiW/ZrO2
observed. This is probably due to the strong HPAs interaction with
Zr OH groups which is an acid-base reaction where the Zr OH
Pt/ZrO2
groups act as a base and HPAs as a Brønsted acid. Some NMR could
also support these findings. Oliveira et al. [25] reported in their
studies on ZrO2 supported HPW catalyst two lines present at −15.0
and −13.0 ppm. The first chemical shift is related to the Keggin unit 1400 1450 1500 1550 1600 1650 1700
-1
and the second is assigned to zirconia surface bridges. A similar Wavenumber (cm )
shift was also observed by Alsalme et al. [27] that indicated strong
interaction between finely dispersed amorphous HPW species and Fig. 4. Py-IR spectra of Pt/ZrO2 and Pt-HPAs/ZrO2 catalysts.
ZrO2 .
In summary, it can be concluded that the thermal stability of
Similar patterns were also obtained for Pt-HSiW/ZrO2 . The max-
HSiW and HPW was stronger than HPMo; the Keggin structure of
imum TPD peak position of supported HPAs catalysts decreased
HSiW and HPW was well preserved on the support surface, but that
in the following order: Pt-HPW/ZrO2 (479 ◦ C) > Pt-HSiW/ZrO2
of HPMo was partially destroyed.
(400 ◦ C) > Pt-HPMo/ZrO2 (337 ◦ C). The results indicates that the acid
strength of HPAs is in the order of HPW > HSiW > HPMo, consistent
3.1.2. Acidic properties of catalysts well with the previous report [36]. The amount of acid sites can be
Appropriate acidity is essential for the selective hydrogenol- calculated from TPD desorption peak area and the results are listed
ysis of glycerol to 1,3-PDO due to the bi-functional mechanism in Table 1. The total acidities of the as-prepared catalysts were in an
[9,14,22,34]. The NH3 -TPD technique has been extensively order of Pt-HPMo/ZrO2 > Pt-HPW/ZrO2 > Pt-HSiW/ZrO2 > Pt/ZrO2 .
employed to probe the acidity of solid acid catalysts. Meanwhile, Fig. 4 illustrates the FTIR spectra after pyridine adsorption on
the FT-IR after pyridine adsorption is a useful tool to determine the Pt/ZrO2 and Pt-HPAs/ZrO2 . Pyridine adsorption at Brønsted acid
nature and amount of acid sites. sites and Lewis acid sites exhibited typical bands centering at
The NH3 -TPD profiles of the as-prepared samples are pre- 1540–1548 cm−1 and 1445–1460 cm−1 , respectively [13]. The con-
sented in Fig. 3. Generally the strength of solid acid sites centration of Brønsted acid and Lewis acid are obtained from the
with TPD profiles can be classified into three types by the integral intensities of the typical bands and the results are pre-
temperature of desorption NH3 weak (150–300 ◦ C), moderate sented in Table 1. The results show that Pt/ZrO2 possessed only
(300–450 ◦ C) and strong (450–650 ◦ C) [35]. The NH3 -TPD result Lewis acid sites. The loading of HPAs resulted in a noticeable
of Pt/ZrO2 showed poorly resolved desorption peaks in two dis- increase in the Brønsted acid on all the samples except that on
tinct temperature regions, centering about 255 ◦ C and 437 ◦ C, HPMo, in which HPMo partially decomposed to MoO3 . On the
respectively. On the other hand, Pt-HPMo/ZrO2 exhibited an over- other hand, the amount of Lewis acid sites diminished after HPAs
lapping peak including one main peak and two shoulders. In impregnation except HPMo. The decrease can be related to block-
the case of Pt-HPW/ZrO2 , two main peaks and a weak shoulder age of Lewis acid sites due to deposition of HPAs. The increase in
were observed, around 222 ◦ C, 479 ◦ C and 157 ◦ C, respectively. Lewis acid sites of Pt-HPMo/ZrO2 sample may be derived from the
decomposition product MoO3 . It can be seen that the total acidity
determined by Py-IR correlated well with the NH3 -TPD results and
HPAs had significant influence on the relative amount of Brønsted
acid and Lewis acid.

3.2. Effect of supported heteropolyacids on glycerol


hydrogenolysis
Pt-HPMo/ZrO2
The hydrogenolysis of glycerol in aqueous solution was car-
Intensity

Pt-HPW/ZrO2 ried out over Pt/ZrO2 and Pt-HPAs/ZrO2 . Fig. 5 illustrates glycerol
conversion and 1,3-PDO selectivity as a function of reaction
temperature. Generally glycerol conversion increased remarkably
Pt-HSiW/ZrO2 whereas 1,3-PDO selectivity decreased with the rising temperature
from 160 to 240 ◦ C over all catalysts. It can be seen that temperature
Pt/ZrO2 has a positive effect on glycerol conversion, as claimed by previous
literature [9,10,13]. However, higher temperature disfavored the
formation of 1,3-PDO and increased by-products 1,2-PDO and 1-
100 200 300 400 500 600 propanol (1-PO). Gong et al. [9] reported similar phenomenon over
o
Temperature ( C) bi-functional Pt/WO3 /TiO2 /SiO2 catalyst during hydrogenolysis of
glycerol in a batch reactor and supposed that elevated temperature
Fig. 3. NH3 -TPD profiles of Pt/ZrO2 and Pt-HPAs/ZrO2 catalysts. can be favorable to activate terminal hydroxyl groups of glycerol.
124 S. Zhu et al. / Catalysis Today 212 (2013) 120–126

2
100 Lewis acid sites/[µmol/m ]
Pt/ZrO2 90 120 150 180
Pt-HSiW/ZrO2
80 Pt-HPW/ZrO2 20 20

Pt-HPMo/ZrO2
Conversion (%)

60 15 15

1,3-PDO yield/%
1,2-PDO yield/%
40 10 10

20 5 5

0 0 0
160 180 200 220 240
o
Temperature ( C ) 0 20 40 60 80
2
Brønsted acid sites/[µmol/m ]

Fig. 6. Correlation between the amount of acids sites and product yield obtained at
180 ◦ C of the reaction.
50 Pt/ZrO2
Pt-HSiW/ZrO2 In comparison to Pt/ZrO2 , supported HPAs catalysts enhanced the
Pt-HPW/ZrO2 catalytic activity considerably and the most prominent benefit
40
Selectivity of 1,3-PDO (%)

rose from the significant impact on product distribution. For sup-


Pt-HPMo/ZrO2 ported HPAs catalysts, large amounts of sequential hydrogenolysis
30
products (1-PO/2-PO) were formed and the selectivity to degrada-
tion products (EG and ethanol) diminished moderately. Among the
four catalysts, Pt-HSiW/ZrO2 was the most selective catalyst as it
20 showed 48.1% selectivity to 1,3-PDO at 180 ◦ C.
It is well known that the dispersion of active metal and acidity of
catalyst play a key role in the bi-functional hydrogenolysis mech-
10 anism of glycerol [13,14]. It was found that the addition of HPAs
to Pt/ZrO2 slightly decreased the surface area and the dispersion of
Pt. These results inferred that the increased hydrogenolysis ability
0 of Pt-HPAs/ZrO2 was related to the acidity of catalyst. As shown
in Tables 1 and 2, it can be observed that the conversion increased
160 180 200 220 240
o
gradually with the total amount of acid sites over these catalysts. In
Temperature ( C ) lower reaction temperature, Pt-HPMo/ZrO2 exhibited the highest
activity related to high amount of acid sites. However, the catalytic
Fig. 5. Effect of reaction temperature on glycerol conversion and 1,3-PDO selec-
tivity over Pt/ZrO2 and Pt-HPAs/ZrO2 catalysts. Reaction conditions: 5.0 MPa, 2.0 g activity became less than other catalysts at 220 and 240 ◦ C, prob-
catalysts; 10 wt.% glycerol aqueous solution, H2 /glycerol = 137:1 (molar ratio), ably due to the bad hydrothermal stability of HPMo. Varisli et al.
WHSV = 0.09 h−1 . [37] reported that HPMo was less stable than HSiW and HPW at
temperatures higher than 200 ◦ C in dehydration of ethanol.
On the other hand, the increase of reaction temperature promoted It is clearly stated in literature that Brønsted acid sites favors
the excessive hydrogenolysis of propanediols (PDOs) to propanols the removal of the secondary hydroxyl group from glycerol to gen-
(1-PO + 2-PO). erate 1,3-PDO [9,13,24]. As presented in Tables 1 and 2, the yield
The activity and main products selectivity of Pt/ZrO2 and Pt- to 1,3-PDO is proportional to the concentration of Brønsted acid
HPAs/ZrO2 in glycerol hydrogenolysis at 180 ◦ C are summarized sites. A close correlation between the 1,3-PDO yield at 180 ◦ C and
in Table 2. The Pt/ZrO2 catalyst produced lower glycerol conver- the concentration of Brønsted acid sites can be derived, as dis-
sion, a high 1,2-PDO selectivity and very poor 1,3-PDO selectivity, played in Fig. 6. Although Brønsted acid sites were not available in
suggesting that Pt/ZrO2 was not effective for 1,3-PDO formation. the fresh Pt/ZrO2 , they could be generated by interaction of Lewis

Table 2
Catalytic performance of glycerol hydrogenolysis over Pt/ZrO2 and Pt-HPAs/ZrO2 catalysts at 180 ◦ C.

Catalyst Conversion (%) Selectivity (%)a

1,3-PDO 1,2-PDO 1-PO 2-PO EG Ethanol Acetone Propanal Othersb

Pt/ZrO2 18.7 3.9 76.1 3.2 0.6 6.6 3.9 1.2 0.5 4.0
Pt-HSiW/ZrO2 24.1 48.1 16.5 21.8 4.5 0.7 0.9 0.2 1.1 6.2
Pt-HPW/ZrO2 25.5 32.9 10.9 37.9 5.2 0.6 4.2 0.7 5.3 2.3
Pt-HPMo/ZrO2 27.1 7.8 39.2 30.4 3.2 1.0 1.9 1.1 1.4 14.0

Reaction conditions: 180 ◦ C, 5.0 MPa, 2.0 g catalysts; 10 wt.% glycerol aqueous solution, H2 /glycerol = 137:1 (molar ratio), WHSV = 0.09 h−1 .
a
1,3-PDO, 1,3-propanediol; 1,2-PDO, 1,2-propanediol; 1-PO, 1-propanol; 2-PO, 2-propanol; EG, ethylene glycol.
b
Others: acetol, methanol, acetic acid, propionic acid, propane, methane, and so on.
S. Zhu et al. / Catalysis Today 212 (2013) 120–126 125

Table 3
Reaction results of various reactants over Pt-HSiW/ZrO2 .

Reactant Conversion (%) Selectivity (%)a

1,3-PDO 1,2-PDO 1-PO 2-PO EG Ethanol Acetone Propanal Othersb

Glycerol 24.1 48.1 16.5 21.8 4.5 0.7 0.9 0.2 1.1 6.2
1,2-PDO 47.1 0.0 – 85.2 8.9 0.0 1.8 2.0 0.7 1.4
1,3-PDO 9.1 – 0.0 96.5 1.0 0.0 1.5 0.0 0.6 0.4
EG 53.7 0.0 0.0 0.0 0.0 – 97.7 0.0 0.0 2.3

Reaction conditions: 180 ◦ C, 5.0 MPa, 2.0 g catalysts; 10 wt.% aqueous solution of reactant, WHSV = 0.09 h−1 .
a
1,3-PDO, 1,3-propanediol; 1,2-PDO, 1,2-propanediol; 1-PO, 1-propanol; 2-PO, 2-propanol; EG, ethylene glycol.
b
Others: acetol, methanol, acetic acid, propionic acid, propane, methane, and so on.

acid sites with steam to give rise to 1,3-PDO. Among the tested routes. The reaction results are summarized in Table 3. Regarding
supported HPAs catalysts, HSiW led to the highest 1,3-PDO yield the reaction activity, 1,3-PDO was much lower than 1,2-PDO and
whereas HPMo displayed the lowest 1,3-PDO yield. The excellent glycerol. Additionally, the selectivity to 1-PO was much higher than
performance of HSiW was attributed to high Brønsted acid sites and that to 2-PO in the hydrogenolysis of 1,3-PDO or 1,2-PDO, which
hydrolytic stability even in hydrothermal environment. Atia et al. was related to the high selectivity of 1-PO formation in the glycerol
[31] reported that supported HSiW was more selective than HPW hydrogenolysis. The 1-PO/2-PO ratio in the glycerol hydrogenolysis
and HPMo in glycerol dehydration to acrolein which has the same was similar to that in the 1,2-PDO hydrogenolysis. The combined
intermediate 3-HPA as 1,3-PDO. Tsukuda et al. [36] also acclaimed results indicate that 1-PO and 2-PO are mainly formed via 1,2-PDO
that HSiW showed high performance in glycerol dehydration due in the glycerol hydrogenolysis, consistent well with the previous
to high Brønsted acid strength and resistance to water. On the other report [10].
hand, HPMo always showed the lowest 1,3-PDO selectivity whereas A plausible reaction pathway for glycerol hydrogenolysis and
high 1,2-PDO selectivity due to partial decomposition of HPMo, degradation is presented in Fig. 7. Reaction routes of hydrogenoly-
high Lewis acid sites and bad thermal stability. sis products will be discussed firstly. Previous research [22] shows
According to the literature [8,38], Lewis acid sites are respon- that PDOs formation proceeds via dehydration of glycerol to ace-
sible for the glycerol dehydration to acetol which can further tol and 3-HPA on acid catalyst and consecutive hydrogenation to
hydrogenation to form 1,2-PDO over metal catalysts. Therefore, we 1,2-PDO and 1,3-PDO over metal catalysts, respectively. Alhanash
attempted to correlate the concentration of Lewis acid sites with et al. [8] have suggested that acetol was formed on Lewis acid
1,2-PDO yield at 180 ◦ C. Unfortunately, the correlation is not good, sites whereas acrolein on Brønsted acid sites in glycerol dehydra-
which could be mainly attributed to sequential hydrogenolysis of tion, which is well in agreement with our results. In the case of
1,2-PDO to propanols. Qin et al. [10] proposed that propanols are Pt/ZrO2 (Table 2), the main product was 1,2-PDO due to its large
mainly formed via 1,2-PDO in glycerol hydrogenolysis, and it was Lewis acid sites. However, addition of HSiW to Pt/ZrO2 catalysts
also verified by our separate experiment, as discussed later. Thus, improved the catalytic activity and 1,3-PDO selectivity remarkably
the calculated 1,2-PDO yield should include the observed 1,2-PDO because of the enhanced Brønsted acid. To corroborate this point
and the part converted to propanols. It is interesting to find that further, we also investigated the Pt-free HSiW/ZrO2 catalyst for
a close correlation between the calculated 1,2-PDO yield at 180 ◦ C glycerol hydrogenolysis under identical conditions as Pt/ZrO2 at
and the concentration of Lewis acid sites was obtained, as illus- 180 ◦ C. The result showed that glycerol conversion was only 2.3%
trated in Fig. 6. and the primary product was acrolein with a selectivity of 66.5%,
whereas the selectivity of acetol was only 13.3%, consistent well
3.3. Reaction mechanism of glycerol hydrogenolysis over with the conclusion reported by Alhanash et al. [8]. The 3-HPA
Pt-HSiW/ZrO2 was not detected in any experiment whereas acetol was observed,
implying that acetol is much less reactive than 3-HPA. 3-HPA can
Activity tests of reaction products 1,2-PDO, 1,3-PDO and dehydrate into acrolein or hydrogenate to 1,3-PDO. Additionally,
EG over Pt-HSiW/ZrO2 were also examined under similar 3-HPA can further decompose to form acetaldehyde and formalde-
conditions as used for glycerol to elucidate the reaction hyde through a retro-aldol condensation reaction. Then acetic

OH + CH4

O O OH
+H2 -H2O +H2
OH OH
OH
-H2O -H2O
+H2 OH
O

-H2O
-H2O [O]
HO OH +H2
HO O HO OH OH
OH
O O
-H2 [O]
-H2 O + O
OH

HO OH HO -CO -H2O
O HO +H2
OH OH
O OH O

Fig. 7. Reaction scheme of glycerol hydrogenolysis and degradation reactions.


126 S. Zhu et al. / Catalysis Today 212 (2013) 120–126

acid can be formed by the oxidation of acetaldehyde in aqueous Acknowledgements


solution.
Propanal and acetone are formed through PDOs dehydra- This work was financially supported by the Natural Science
tion and then hydrogenate to 1-PO and 2-PO, respectively. It is Foundation of China (No. 20976185), and the Major State Basic
reported that propanal can be further oxidized to produce pro- Research Development Program of China (973 Program) (No.
pionic acid in aqueous solution [39]. Ethanol and methane are 2012CB215305).
produced by C C cleavage of 1,2-PDO. Additionally, trace propane
is formed by sequential hydrogenolysis of propanols. Methane References
can be formed from methanol hydrogenation or from PDOs
degradation. [1] A. Corma, S. Iborra, A. Velty, Chemical Reviews 107 (2007) 2411–2502.
[2] G.W. Huber, S. Iborra, A. Corma, Chemical Reviews 106 (2006) 4044–4098.
Degradation product EG showed higher hydrogenolysis activity [3] M. Pagliaro, R. Ciriminna, H. Kimura, M. Rossi, C. Della Pina, Angewandte Chemie
than glycerol and PDOs over Pt-HSiW/ZrO2 . The fact that EG was International Edition 46 (2007) 4434–4440.
not formed in the reaction of PDOs hydrogenolysis indicated [4] C.H.C. Zhou, J.N. Beltramini, Y.X. Fan, G.Q.M. Lu, Chemical Society Reviews 37
(2008) 527–549.
that it was formed directly from glycerol. The EG formation [5] A. Behr, J. Eilting, K. Irawadi, J. Leschinski, F. Lindner, Green Chemistry 10 (2008)
proceeds via dehydrogenation of glycerol to glyceraldehyde and 13–30.
subsequent C C bond cleavage. The mechanism by which the [6] D. Liang, J. Gao, H. Sun, P. Chen, Z. Hou, X. Zheng, Applied Catalysis B: Environ-
mental 106 (2011) 423–432.
C C bond cleavage occurs on Pt sites is subject to some debate
[7] D.L. King, L. Zhang, G. Xia, A.M. Karim, D.J. Heldebrant, X. Wang,
that it proceeds via decarbonylation or retro-aldol condensation. T. Peterson, Y. Wang, Applied Catalysis B: Environmental 99 (2010)
In the presence of base, Maris and Davis [40] have asserted that 206–213.
EG formation is attributed to retro-aldol reaction rather than [8] A. Alhanash, E.F. Kozhevnikova, I.V. Kozhevnikov, Applied Catalysis A: General
378 (2010) 11–18.
through decarbonylation. King et al. [7] proposed that C C bond [9] L. Gong, Y. Lu, Y. Ding, R. Lin, J. Li, W. Dong, T. Wang, W. Chen, Applied Catalysis
cleavage was formed through glyceraldehyde decarbonylation A: General 390 (2010) 119–126.
under an acidic condition. It seems that both offer plausible [10] L.-Z. Qin, M.-J. Song, C.-L. Chen, Green Chemistry 12 (2010) 1466–1472.
[11] A. Behr, J. Eilting, K. Irawadi, J. Leschinski, F. Lindner, Chimica Oggi 26 (2008)
pathways which are highly dependent on the operation conditions 32–36.
and catalytic system used. Very recently, Vlachos and co-works [12] G.A. Kraus, Clean-Soil Air Water 36 (2008) 648–651.
[41] have identified the most favorable decomposition pathway [13] S. Zhu, Y. Zhu, S. Hao, H. Zheng, T. Mo, Y. Li, Green Chemistry 14 (2012)
2607–2616.
of glycerol on Pt with quantum mechanics calculations and the [14] M. Balaraju, V. Rekha, P.S.S. Prasad, B.L.A.P. Devi, R.B.N. Prasad, N. Lingaiah,
pathway is shown as follows: C3 H8 O3 → CHOHCHOHCH2 OH → Applied Catalysis A: General 354 (2009) 82–87.
CHOHCHOHCHOH → CHOHCOHCHOH → COHCOHCHOH → [15] L. Ma, D.H. He, Catalysis Today 149 (2010) 148–156.
[16] S. Panyad, S. Jongpatiwut, T. Sreethawong, T. Rirksomboon, S. Osuwan, Catalysis
COCOHCHOH → CO + COHCHOH. These above results indicate Today 174 (2011) 59–64.
that EG formation in our reaction chemistry proceeded most [17] M.G. Musolino, L.A. Scarpino, F. Mauriello, R. Pietropaolo, ChemSusChem 4
probably via decarbonylation of glyceraldehyde. However, not CO (2011) 1143–1150.
[18] Y. Amada, Y. Shinmi, S. Koso, T. Kubota, Y. Nakagawa, K. Tomishige, Applied
but CO2 was detected in glycerol hydrogenolysis. Wawrzetz et al.
Catalysis B: Environmental 105 (2011) 117–127.
[42] have observed similar phenomenon during aqueous-phase [19] Y. Shinmi, S. Koso, T. Kubota, Y. Nakagawa, K. Tomishige, Applied Catalysis B:
hydrogenolysis of glycerol over Pt/Al2 O3 and argued that CO2 Environmental 94 (2009) 318–326.
is generated by decarbonylation of glyceraldehyde followed by [20] L. Huang, Y.L. Zhu, H.Y. Zheng, G.Q. Ding, Y.W. Li, Catalysis Letters 131 (2009)
312–320.
water gas shift. Additionally, glyceraldehyde is also not detected [21] T. Kurosaka, H. Maruyama, I. Naribayashi, Y. Sasaki, Catalysis Communications
because it is very reactive. Interestingly, we observed dihydroxy- 9 (2008) 1360–1363.
acetone which can be produced from glycerol dehydrogenation or [22] I. Gandarias, P.L. Arias, J. Requies, M.B. Guemez, J.L.G. Fierro, Applied Catalysis
B: Environmental 97 (2010) 248–256.
glyceraldehyde transformation. Liang et al. [6] have disclosed that [23] J. Chaminand, L. Djakovitch, P. Gallezot, P. Marion, C. Pinel, C. Rosier, Green
glyceraldehyde is more reactive than dihydroxyacetone over Pt Chemistry 6 (2004) 359–361.
based catalyst. Glyceraldehyde was not detected in our products [24] J. Oh, S. Dash, H. Lee, Green Chemistry 13 (2011) 2004–2007.
[25] C.F. Oliveira, L.M. Dezaneti, F.A.C. Garcia, J.L. de Macedo, J.A. Dias, S.C.L. Dias,
most likely because it can undergo decarbonylation or convert to K.S.P. Alvim, Applied Catalysis A: General 372 (2010) 153–161.
dihydroxyacetone rapidly. As shown in Table 3, the major product [26] O. Toshio, Catalysis Today 73 (2002) 167–176.
of EG hydrogenolysis was ethanol and trace of acetaldehyde. [27] A.M. Alsalme, P.V. Wiper, Y.Z. Khimyak, E.F. Kozhevnikova, I.V. Kozhevnikov,
Journal of Catalysis 276 (2010) 181–189.
The EG can undergo dehydration to acetaldehyde intermediate [28] M. Misono, Catalysis Today 144 (2009) 285–291.
followed by subsequent hydrogenation to ethanol. [29] A. Micek-Ilnicka, Journal of Molecular Catalysis A: Chemical 308 (2009) 1–14.
[30] C.A. Emeis, Journal of Catalysis 141 (1993) 347–354.
[31] H. Atia, U. Armbruster, A. Martin, Journal of Catalysis 258 (2008)
71–82.
4. Conclusions [32] X. Cui, J. Xu, C. Zhang, Y. Yang, P. Gao, B. Wu, Y. Li, Journal of Catalysis 282 (2011)
35–46.
The results of this study indicate that heteropolyacids mod- [33] S.H. Chai, H.P. Wang, Y. Liang, B.Q. Xu, Applied Catalysis A: General 353 (2009)
213–222.
ified Pt/ZrO2 catalysts were effective in glycerol hydrogenolysis
[34] M. Chia, Y.J. Pagán-Torres, D. Hibbitts, Q. Tan, H.N. Pham, A.K. Datye, M. Neurock,
to 1,3-PDO. Addition of HPAs to Pt/ZrO2 catalysts increased the R.J. Davis, J.A. Dumesic, Journal of the American Chemical Society 133 (2011)
catalytic activity and 1,3-PDO selectivity considerably due to the 12675–12689.
[35] A.S. de Oliveira, S.J.S. Vasconcelos, J.R. de Sousa, F.F. de Sousa, J.M. Filho, A.C.
enhanced acidity. Compared to Pt-HPW/ZrO2 and Pt-HPMo/ZrO2 ,
Oliveira, Chemical Engineering Journal 168 (2011) 765–774.
Pt-HSiW/ZrO2 obtained the highest selectivity to 1,3-PDO, which [36] E. Tsukuda, S. Sato, R. Takahashi, T. Sodesawa, Catalysis Communications 8
was probably related to the high Brønsted acid sites and good (2007) 1349–1353.
thermal stability. The 1,3-PDO yield was proportional to the con- [37] D. Varisli, T. Dogu, G. Dogu, Chemical Engineering Science 62 (2007)
5349–5352.
centration of the Brønsted acid sites, whereas the 1,2-PDO yield [38] Y.T. Kim, K.-D. Jung, E.D. Park, Applied Catalysis B: Environmental 107 (2011)
was proportional to the concentration of Lewis acid sites. Finally, 177–187.
the reaction pathways have been proposed, and the results indi- [39] F. Wang, J.-L. Dubois, W. Ueda, Journal of Catalysis 268 (2009) 260–267.
[40] E.P. Maris, R.J. Davis, Journal of Catalysis 249 (2007) 328–337.
cate that 1-PO and 2-PO are mainly formed via 1,2-PDO in glycerol [41] Y. Chen, M. Salciccioli, D.G. Vlachos, Journal of Physical Chemistry C 115 (2011)
hydrogenolysis. In addition, the formation of by-product EG is con- 18707–18720.
firmed to originate from decarbonylation of glyceraldehyde under [42] A. Wawrzetz, B. Peng, A. Hrabar, A. Jentys, A.A. Lemonidou, J.A. Lercher, Journal
of Catalysis 269 (2010) 411–420.
given experimental conditions.

You might also like