Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Journal of Molecular Catalysis A: Chemical 414 (2016) 18

Contents lists available at ScienceDirect

Journal of Molecular Catalysis A: Chemical


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

Enhancing catalytic performance of phosphorus-modied ceria


supported VPO catalysts for n-butane oxidation
Hua-Yi Wu a,b,c , Peng Jin a,b , Yi-fei Sun d , Mei-Hua Yang a,b , Chuan-Jing Huang a,b, ,
Wei-Zheng Weng a,b , Hui-Lin Wan a,b,
a
State Key Laboratory of Physical Chemistry of Solid Surfaces, National Engineering Laboratory for Green Chemical Productions of Alcohols, Ethers and
Esters, Xiamen University, Xiamen 361005, PR China
b
Department of Chemistry, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, PR China
c
Department of Pharmacy, Xiamen Medical College, Xiamen 361008 PR China
d
Department of Chemical and Materials Engineering, University of Alberta, Edmonton, Alberta T6G 2V4, Canada

a r t i c l e

i n f o

Article history:
Received 24 September 2015
Received in revised form
14 December 2015
Accepted 27 December 2015
Available online 2 January 2016
Keywords:
n-Butane
Oxidation
Maleic anhydride
Supported VPO catalyst
Ceria

a b s t r a c t
Vanadium phosphorus oxide (VPO) catalysts supported on CeO2 and P-modied CeO2 (P-CeO2 ) have been
rstly synthesized for selective oxidation of n-butane to maleic anhydride (MA). The catalysts supported
on P-CeO2 showed higher catalytic performance, especially higher MA selectivity than catalysts on pure
ceria, and their specic activities (normalized to per unit mass of VPO loading) were also superior to
that of bulk VPO. TG, TEM, XRD, Raman, H2 -TPR and XPS results revealed that the structure and property
of VPO are largely affected by the support. The loaded VPO existed almost wholly as VOPO4 phases on
pure ceria, while mainly in the form of (VO)2 P2 O7 with small amount of VOPO4 phases on phosphorusmodied ceria. Both VPO/P-CeO2 and bulk VPO catalysts have the similar VPO phase structure, but the
former showed the enhanced reducibility of V4+ phase. These differences in structure and properties of
the catalysts were discussed and related to their different catalytic performances.
2016 Published by Elsevier B.V.

1. Introduction
The 14-electron partial oxidation of n-butane to maleic anhydride (MA) is the most successful sample for the commercial
application of light parafn conversion [1]. The most effective catalyst for this reaction is vanadiumphosphorus-oxide with vanadyl
pyrophosphate ((VO)2 P2 O7 ) as the main component [2,3]. So far,
great effort has been devoted to getting insight into the peculiar
properties of the catalyst, while questions still exist as to the form
of the active phase (amorphous versus crystallized) and the nature
of the active components [48].
Denitely, supported VPO catalysts have several superiorities
compared with bulk VPO catalysts, such as larger surface area to
volume ratio of active component, more active sites available for
per unit mass of catalyst and better mechanical strength. In recent

Corresponding authors at: State Key Laboratory of Physical Chemistry of


Solid Surfaces, National Engineering Laboratory for Green Chemical Productions
of Alcohols, Ethers and Esters, Xiamen University, Xiamen 361005, PR China.
Fax: +86 592 2183047.
E-mail addresses: huangcj@xmu.edu.cn (C.-J. Huang), hlwan@xmu.edu.cn
(H.-L. Wan).
http://dx.doi.org/10.1016/j.molcata.2015.12.024
1381-1169/ 2016 Published by Elsevier B.V.

years, several groups have reported their investigations on VPO catalysts supported on SiO2 , TiO2 , SiC, Al-containing MCM-41, SBA-15,
ZrO2 , RGO and ZrP [916]. Notably, the preparation methods have
great inuences on phase compositions and the catalytic performance of the VPO catalysts [9,10,13,17]. The results of plenty of
literature have reached the agreement that VPO catalysts synthesized in an organic medium have better performance than those
synthesized in an aqueous solution [1820]. Generally, in organic
synthesis, a mixture of iso-butyl/benzyl alcohols is adopted for producing VPO catalysts and the use of an organic medium yields an
increase in surface area of the precursor and enhances the activity of
the catalyst [1922]. Nevertheless, as reported previously [2325],
supported VPO catalysts, especially for those prepared in aqueous
medium, usually had a certain amount of VOPO4 such as -VOPO4
and -VOPO4 . The existences of a massive amount of -VOPO4 and
-VOPO4 were harmful to the catalytic performance of supported
VPO catalysts. Besides, it is found that the nature of the support
also affects phase compositions and catalytic performance of the
VPO catalysts [12,26]. Nie et al. [12] modied MCM-41 with certain
amount of aluminum and concluded that the incorporated Al3+ in
the framework of MCM-41 support might have an impact on MA
selectivity. Buena et al. [26] adopted hydrophobic and hydrophilic

H.-Y. Wu et al. / Journal of Molecular Catalysis A: Chemical 414 (2016) 18

silica as the supports and found that (VO)2 P2 O7 could be formed


easily on the former support.
Ceria is usually used as catalyst support or catalyst in various applications such as CO oxidation, three-way catalysts (TWCs),
selective oxidation and the watergas shift [2730]. It has attracted
widely attention for its distinctive properties, including high lattice oxygen mobility, high oxygen-storage capacity and ease of
exchange of surface lattice oxygen with gas-phase oxygen [31,32].
Due to these properties, supporting VPO catalysts on ceria perhaps
contains fascinating information. However, so far few reports have
been made on the employment of ceria as support for VPO catalyst.
Wachs et al. [33] investigated the oxidation of n-butane to maleic
anhydride over a series of model-supported vanadia catalysts and
observed that the MA selectivity closely followed the lewis acid
strength of the oxide support cations, Al > Nb > Ti > Si > Zr > Ce. Herron et al. [34] employed ceria-supported VPO catalyst in n-butane
oxidation and found that the catalyst exhibited low n-butane conversion and especially low MA selectivity. No detailed studies have
been done to understand the effect of this support on the structure
and properties of VPO catalysts. Especially, no any attempt has been
made to modify ceria and thus improve the catalytic performance
of ceria-supported VPO catalysts for the reaction.
It has been reported that the addition of phosphorus to some
of the catalysts can enhance their catalytic performance in ODH
of light alkanes [14,29,35,36]. For chromium oxide supported on
phosphorus-loaded alumina, the addition of phosphorus inuenced the nature of the Cr oxide species and thus heightened the
selectivity to propylene [35]. For P-containing ceria, the phosphorus decreased the lattice oxygen mobility through the formation
of Ce O P bonds and increased the propylene selectivity [29]. In
case of VPO catalysts supported on H3 PO4 -treated ZrO2 , the way
of H3 PO4 treatment conducted on ZrO2 had a great inuence on
the state and structure of the VPO component loaded on it [14]. In
case of phosphorus-modied silica supported vanadium catalysts,
phosphorus modication improved the yield of MA in the selective
oxidation of n-butane [36].
The present study is the rst to adopt phosphorus-modied
ceria as a support material for VPO catalysts and to investigate
the inuence of the support on the structure and nature of supported VPO species. The physicochemical properties of the VPO
catalysts supported on ceria and on phosphorus-modied ceria
were systematically examined by XRD, TEM, Raman, XPS and H2 TPR techniques.

2. Experimental details
2.1. Materials and preparation
2.1.1. Synthesis of pure ceria
Ceria was fabricated according to the method described elsewhere [37]. Briey, a solution of (NH4 )2 CO3 (0.9 mol L1 , 50 mL)
was added dropwise to the cerium nitrate solution (0.3 mol L1 ,
50 mL) and white powders were produced immediately. After vigorous stirring for 30 min, the mixture was transferred to a reux
device and heated at 100 C for 12 h. The powder obtained after ltration was washed with water and ethanol. Subsequently, it was
dried at 80 C for 12 h and then calcined at 500 C for 4 h.

2.1.2. Synthesis of P-modied CeO2


To achieve phosphorus-modied ceria, the as-prepared ceria
powder (1 g) was added to solutions of NH4 H2 PO4 (10 mL) at concentrations of 1, 1.5 and 2 mol L1 . The above suspension was
stirred at room temperature for 6 h and then ltered by Buchner
funnel. The obtained solid was dried at 80 C for 12 h and calcined

at 500 C for 4 h. The phosphorus-modied samples were marked


as 1P-CeO2 , 1.5P-CeO2 and 2P-CeO2 , respectively.
2.1.3. Synthesis of supported VPO catalysts
Typically, supported VPO precursor was prepared in the following way: V2 O5 (0.6 g, 3.3 mmol) was added to a mixture of
iso-butanol and benzyl alcohol (26 mL, volume ratio of 1/1) and
reuxed at 135 C for 5 h, then a suitable amount of support materials was introduced. After one hour, phosphoric acid (0.6 mL, 85%)
was added dropwise to reach a P/V atomic ratio of 1.2. After reuxing for another 6 h, the suspension was ltered, and the solid was
washed with iso-butanol and ethanol, and then dried in air at 110 C
for 12 h. The catalysts are denoted as xVPO/CeO2 or xVPO/P-CeO2 ,
where x = 9, 14, 18 and 23%, which refer to the vanadium content
of 9, 14, 18 and 23 wt%, respectively.
The precursor, previously pressed into pellets and sieved to
4060 mesh, was activated under the reaction conditions (1.5%
n-butane in air, GHSV = 1500 h1 ) from ambient temperature to
400 C at a rate of 2 C min1 , holding it at this temperature for 16 h.
After that, the precursor was transformed into the nal catalyst.
2.2. Catalyst characterization
The structural properties of the catalysts were characterized by
using the Rigaku Ultima IV X-ray diffractometer with monochromatized CuK radiation. Raman spectra were recorded by using
Renishaw UVvis Raman 1000 system equipped with a CCD detector and a Leica DMLM microscope. The line at 532 nm of Ar+ laser
was used for excitation. H2 -TPR experiments were carried out in
a quartz tube of 4 mm diameter with 50 mg sample by raising the
temperature from 100 to 800 C at a rate of 10 C min1 under a
5%H2 /Ar mixture owing at 30 mL min1 . The thermogravimetric
(TG) analysis was conducted on a TG209F1 thermal analyzer. The
surface compositions of the samples were obtained from X-ray
photoelectron spectroscopy (XPS) using a VG ESCALAB/Auger. All
binding energies were corrected with reference to the C1s signal
located at 284.6 eV. The morphologies of the samples were investigated by TECNAI F-30 FEG transmission electron microscope.
2.3. Catalytic reaction
The selective oxidation of n-butane was executed in a xed-bed
ow microreacter using 0.5 g of catalyst with a gas hourly space
velocity (GHSV) of 1500 mL g1 h1 . The reaction mixture (1.5% nbutane in air) was fed into the reactor via calibrated mass ow
controller. The test temperature was in the range of 340440 C. The
outlet efuents were analyzed with an on-line gas chromatography
and the carbon balance was usually >95%.
3. Results
3.1. Catalyst characterization
Shown in Fig. 1 are the results of thermogravimetric (TG)
analysis and differential thermogravimetry (DTG) analysis on the
thermal decomposition of bulk VPO and 18%VPO/1.5P-CeO2 . Three
distinct stages in weight loss curve are observed for both catalysts.
The rst weight loss step occurring below 200 C originates from
the desorption of physically adsorbed water, the second step from
200 to 430 C arises from the conversion of VOHPO4 to (VO)2 P2 O7
and from the removal of alcohols trapped in the layers of precursor, while further weight loss above 430 C may be due to the
progressive dehydration and to the formation of crystal defects
[38,39]. The results show that the catalysts have similar thermal
decomposition process, except the second stage where the central decomposition temperature of the 18%VPO/1.5P-CeO2 is 342 C,

H.-Y. Wu et al. / Journal of Molecular Catalysis A: Chemical 414 (2016) 18

Fig. 1. TG-DTG curves of (a) bulk VPO and (b) 18%VPO/1.5P-CeO2 .

which is 25 C lower than that of bulk VPO. This difference is understandable, because the alcohols trapped in the layers of precursor
need to react with oxygen to form COx during the thermal treatment [38] and ceria is just a material that has high lattice oxygen
mobility and high oxygen storage capacity. So, for the supported
catalyst, ceria can enhance the oxidation of alcohols and then make
the transformation occur at a low temperature.
The TEM images of the samples are shown in Fig. 2. As depicted
in Fig. 2a, the ceria particles of 200400 nm can be observed.
NH4 H2 PO4 treatment has no effect on the size of the support
(Fig. 2b). For the 18%VPO/1.5P-CeO2 sample, a thin layer is formed

on the surface of the support (Fig. 2c). To determine the composition of this layer, EDX analysis was performed (Fig. S1). The result
revealed that the layer mainly contains vanadium, oxygen, phosphorus and ceria.
The XRD patterns of various samples are depicted in Fig. 3. As
shown in Fig. 3a, 1P-CeO2 and 1.5P-CeO2 present the same XRD
pattern as that of pure CeO2 , where no peaks corresponding to
phosphorus-containing phases can be detected. For the 2P-CeO2
sample, CePO4 peaks are also observed. Fig. 3b shows the XRD
patterns of VPO and supported VPO catalysts. Differing from unsupported VPO catalyst that gives only the characteristic reection of

H.-Y. Wu et al. / Journal of Molecular Catalysis A: Chemical 414 (2016) 18

Fig. 2. TEM images of (a) CeO2 , (b) 1.5P-CeO2 ,and (c) the activated 18%VPO/1.5P-CeO2.

(VO)2 P2 O7 with main peaks at 2 = 18.5, 22.7, 28.2, 29.8 and 43.2 ,
the supported VPO catalysts show patterns exhibiting different VPO
phases depending on the supports. Over 18%VPO/CeO2 , the signal
of (VO)2 P2 O7 phase is very weak and almost undetectable. Besides
the intense peaks belonging to CeO2 , three weak peaks appear at
2 = 12.3, 22.1 and 24.9 , which can be ascribed to VOPO4 2H2 O, VOPO4 and II -VOPO4 phase, respectively [26,40]. The diffraction
peaks corresponding to (VO)2 P2 O7 phase become detectable when
the VPO catalysts are supported on NH4 H2 PO4 -treated ceria and
the signals of (VO)2 P2 O7 phase increase with increasing NH4 H2 PO4
concentration. This observation indicates that NH4 H2 PO4 treatment can enhance the formation of (VO)2 P2 O7 phase. Fig. 3C
illustrates the XRD patterns of VPO/1.5P-CeO2 catalysts with different VPO loadings. All the samples show the presence of ceria
and crystalline (VO)2 P2 O7 , even at VPO loading as low as 9%, and
the XRD signals of (VO)2 P2 O7 increase with the increase of VPO
loading.
The curves reported in Fig. 4a are the Raman spectra of ceria
and phosphorus-modied ceria. For all the samples, a strong band
at 461 cm1 and two weak bands at 252 and 1172 cm1 can be
observed and assigned to the cubic CeO2 phase [41]. On close
inspection, a weak band at 968 cm1 , which is associated with the
CePO4 phase [29], starts to become visible in the 1P-CeO2 sample. The absence of the signal of CePO4 phase in the XRD patterns
of 1P-CeO2 and 1.5P-CeO2 samples may be due to that this phase
is highly dispersed on the support or its amount below XRD detection limit. As indicated in Fig. 4b, the spectrum of bulk VPO exhibits
four bands at 923, 1023, 1134 and 1180 cm1 . The strong band at
923 cm1 can be attributed to the asymmetric P O P stretches of
(VO)2 P2 O7 , the weak bands at 1134 and 1180 cm1 fall in the range
of V O P stretches of (VO)2 P2 O7 and the band at 1023 cm1 is
perhaps due to -VOPO4 [14,41]. For the 18%VPO/CeO2 sample,
a remarkably different spectrum is obtained. No obvious bands
ascribed to (VO)2 P2 O7 could be detected. The observed bands are
assignable to II -VOPO4 (939, 990, 1090 cm1 ) and -VOPO4 (939,
970,1016, 1066 cm1 ), respectively [40,42]. When the VPO catalysts are supported on NH4 H2 PO4 -treated CeO2 , however, the
bands of VOPO4 phases almost vanish, whereas those of (VO)2 P2 O7
appear and grow stronger with the increase of NH4 H2 PO4 concentration. For the 18%VPO/2P-CeO2 sample, the spectrum is almost
the same with that of pure VPO.
Shown in Fig. 5 are the TPR proles of the samples. Pure ceria has
two major reduction peaks. The peak at lower temperature can be
ascribed to the reduction of surface oxygen and the peak at higher
temperature can be attributed to the removal of bulk oxygen from
ceria structure [31]. When modied with phosphorus, the support
shows notable decline in its reduction peaks, especially in the peak
at lower temperature, indicating a reduction of the surface oxygen of ceria due to phosphorus modication. Note that the area
of the reduction peaks for bare supports are very small compared
with those for supported VPO catalysts. So, the contribution of the

Table 1
XPS results of bulk VPO and supported VPO catalysts.
Catalyst

VPO
9%VPO/1.5P-CeO2
14%VPO/1.5P-CeO2
18%VPO/1.5P-CeO2
23%VPO/1.5P-CeO2
18%VPO/CeO2
18%VPO/1 P-CeO2
18%VPO/2P-CeO2

Binding energy (eV)

Average oxidation state of


vanadium (Vox )a

V2p3/2

P2p

O1s

517.2
517.0
517.2
517.3
517.2
518.0
517.3
517.2

133.9
133.7
134.0
134.1
133.9
133.6
133.8
133.9

531.4
531.2
531.4
531.5
531.4
531.2
531.2
531.3

4.16
4.16
4.16
4.16
4.16
4.84
4.37
4.16

a
The vanadium oxidation state is shown to be related to the splitting between
O(1s) and V(2p3/2 ) transition centroids, Vox = 13.82 0.68 [O(1s) V(2p3/2 )] [46].

supports to the reducibility of the catalysts can be neglected. For


the bulk VPO, two main peaks are observed. The strong peak at
approximately 745 C can be attributed to the reduction of lattice
oxygen related to V4+ phase, while the diffuse and dwarf peak in the
range of 400600 C is assignable to the reduction of lattice oxygen
associated with V5+ phases [43]. For supported VPO samples, the
main pattern with two major peaks is still observed, but the relative
areas of the peaks are signicantly inuenced by the supports. On
ceria supported VPO sample, the peak at lower temperature shows
a larger area than the peak at higher temperature, indicating a high
ratio of V5+ /V4+ in this catalyst. In the case of NH4 H2 PO4 -treated
ceria supported VPO catalysts, however, the peak related to V4+
phase becomes predominant and its area increases with increasing
NH4 H2 PO4 concentration. This means that NH4 H2 PO4 treatment of
ceria is conducive to the formation of (VO)2 P2 O7 in the supported
VPO catalyst, which is in agreement with the results of XRD and
Raman.
XPS analysis was used to investigate the oxidation state and
surface composition. As illustrated in Table 1, the V2p3/2 binding
energy of the 18% VPO/CeO2 sample is 518.0 eV, which approximates the value reported for VOPO4 [44,45]. Therefore, the
oxidation state of vanadium in this sample should be close to 5+. The
V2p3/2 binding energies of bulk VPO and NH4 H2 PO4 -treated ceria
supported VPO catalysts are all around 517.2 eV and approach to the
value reported for (VO)2 P2 O7 [13,44]. Hence, it can be concluded
that the valence states of vanadium in those samples are close to
4+ and the vanadium species in the samples are mainly in the form
of (VO)2 P2 O7 . The P2p binding energies of supported VPO samples
are almost the same as those of the VPO sample. As for the binding
energies of oxygen,the values are very similar (531.3 eV) for all the
samples. The average oxidation state of vanadium was calculated
according to the method reported before [46]. From Table 1, we can
see that the average oxidation states of vanadium decrease with
increasing NH4 H2 PO4 concentration, but they remain unchanged
with increasing VPO loading.

H.-Y. Wu et al. / Journal of Molecular Catalysis A: Chemical 414 (2016) 18

Fig. 4. Raman spectra of (a) ceria and phosphorus-modied ceria, and (b) supported
VPOs (18% loading) and bulk VPO.

Fig. 3. The XRD patterns of different samples. (a) ceria and phosphorus-modied
ceria; (b) pure VPO and supported VPOs (18% loading); (c) VPO/1.5P-CeO2 samples
with different VPO loadings.

3.2. Catalytic performance


Table 2 lists the catalytic performance of the pure VPO and
the supported VPO catalysts with 18% loading. The blank test
results indicate that under the reaction conditions, non-catalytic
auto-oxidation of n-butane in the empty reactor is neglectable. All

the catalysts show that the n-butane conversion increases with


increasing temperature while the MA selectivity declines. This is
because higher temperature can result in the consecutive oxidation of MA to carbon oxides. For the supported catalysts, it is found
that the phosphorus modication by impregnating the support
with NH4 H2 PO4 solution markedly improves the catalytic performance. In the case of NH4 H2 PO4 -treated ceria supported VPO
catalysts, both the n-butane conversion and MA selectivity under
a certain temperature rstly increase with the NH4 H2 PO4 concentration and then decrease, attaining the highest values when the
NH4 H2 PO4 concentration reaches 1.5 mol L1 . The VPO/1.5P-CeO2
catalyst exhibits catalytic performance similar to that of bulk VPO,
and shows MA selectivity signicantly higher than those of VPO
catalysts supported on fumed SiO2 [9], TiO2 [10], Al-MCM-41 [12],
SBA-15 [13], SiO2 [47] and ZrO2 [48] and close to that of VPO species
on H3 PO4 -treated ZrO2 [14] at comparable n-butane conversions. In
order to compare the performance more accurately, specic activity (normalized to per unit mass of VPO loading) is usually adopted
as a measure of catalytic behavior of supported VPO catalyst in the
literature [9,13,14,24]. From Table 2, we nd that all the supported

H.-Y. Wu et al. / Journal of Molecular Catalysis A: Chemical 414 (2016) 18

Table 2
The catalytic performances of pure VPO and supported VPO (18% loading) catalysts. Reaction conditions: GHSV = 1500 mL g1 h1 , C4 H10 /air = 1.5/98.5.
Sample

T ( C)

CC4 (mol%)

SMA (mol%)

YMA (mol%)

Specic activitya (mol MA g1 min1 )

Blank
Pure
VPO

420
380
400
420
380
400
420
380
400
420
380
400
420
380
400
420

0
76.4
87.1
98.6
56.5
74.0
87.9
54.5
74.8
90.5
74.0
90.8
98.6
72.1
89.0
95.5

74.2
65.9
57.5
36.8
33.0
27.6
46.8
43.8
38.6
70.5
63.3
51.9
71.1
62.5
50.7

56.7
57.4
56.7
20.8
24.4
24.3
25.5
32.7
34.9
52.1
57.5
51.2
51.2
55.6
48.4

17.6
17.9
17.6
16.2
19.0
18.9
19.9
25.5
27.2
40.6
44.8
39.9
39.9
43.4
37.7

VPO/CeO2

VPO/1PCeO2
VPO/1.5PCeO2
VPO/2PCeO2
a

The activity is dened as MA formation rate and normalized to per unit mass of VPO loading.

Fig. 5. TPR proles of the samples.

catalysts show higher specic activity than bulk VPO, except the
VPO/CeO2 catalyst. Among the catalysts studied, theVPO/1.5P-CeO2
sample exhibits the highest specic activity for selective oxidation of n-butane to MA. The specic activity of this catalyst is also
notably higher than those of VPO catalysts supported on SBA-15
[13] and fumed SiO2 [24] at the same temperature (400 C).
The inuence of VPO loading on the catalytic performance of
the VPO/1.5P-CeO2 samples was also investigated. As shown in
Fig. 6, the n-butane conversion slightly increases with increasing
VPO loading from 9% to 23%, while the MA selectivity rst increases
and then decreases with the increase of VPO loading and achieves
the maximum value (63.3%) at a VPO loading of 18%.
The stability of a catalyst is very important in the view of
potential industrial application. So the stability of the activated
catalyst 18%VPO/1.5P-CeO2 was examined by recording the catalytic performance for 60 h on-line at 400 C. As shown in Fig. 7,
both n-butane conversion and MA yield maintain their initial values during a long running time, suggesting the high stability of the
VPO/1.5P-CeO2 catalyst. The catalyst before and after the reaction
was characterized by XRD and the results are shown in Fig. 8. It
can be seen that the structure of the catalyst remains essentially
unchanged after the 60 h reaction.

Fig. 6. The inuence of VPO loading on the catalytic performance of


VPO/1.5P-CeO2 catalyst. Reaction conditions: T = 400 C, GHSV = 1500 mL g1 h1 ,
C4 H10 /air = 1.5/98.5.

Fig. 9 presents the Arrhenius plots based on the oxidation rate


of n-butane over bulk VPO and 18%VPO/1.5P-CeO2 . The apparent
activation energies on the bulk VPO catalyst and the 18%VPO/1.5PCeO2 sample are estimated from the data obtained at a temperature
range of 340440 C. The Ea values calculated from the Arrhenius
plots are 61 kJ mol1 for bulk VPO and 53 kJ mol1 for 18%VPO/1.5PCeO2 . It is evident that the 18%VPO/1.5P-CeO2 sample has lower
activation energy than the bulk VPO, which indicates that the introduction of the support can reduce the value of the barrier in the
selective oxidation of n-butane.
4. Discussion
In this work, ceria and P-modied ceria are adopted as the supports for the dispersion of VPO components. All the VPO catalysts
supported on P-modied ceria have higher specic activity than
the bulk VPO catalyst. The best MA yield obtained among the supported catalysts at 400 C is 57.5% over 18%VPO/1.5P-CeO2 , which
is at a high level among the supported VPO catalysts reported in
literature [914,17,47,48].

H.-Y. Wu et al. / Journal of Molecular Catalysis A: Chemical 414 (2016) 18

Fig. 7. Alteration of n-butane conversion, MA selectivity and MA yield as a function


of time on stream over 18%VPO/1.5P-CeO2 catalyst. Reaction condition: T = 400 C,
GHSV = 1500 mL g1 h1 , C4 H10 /air = 1.5/98.5.

Fig. 9. The Arrhenius plots over (a) bulk VPO and (b) 18%VPO/1.5P-CeO2 .

Fig. 8. XRD patterns of the fresh activated 18%VPO/1.5P-CeO2 catalyst before (a)
and after (b) reaction at 400 C for 60 h.

As mentioned in the XRD, Raman and XPS sections, the VPO/xPCeO2 catalysts have similar valence states of vanadium and phase
structures with bulk VPO. However, for the supported VPO catalyst,
VPO is dispersed on the support in the form of a thin layer. Such
a manner of dispersion could improve the surface area to volume
ratio of active component and increase the number of active sites
available for per unit mass of catalyst. In addition, compared with
bulk VPO, the high-temperature reduction peak of supported VPO
catalyst shifts to a low temperature (Fig. 5), which indicates that the
support can enhance the reducibility of V4+ phase and thus improve
the activity of the catalyst. This is consistent with the results of the
activation energy measurements. From Fig. 9, it can be observed
that the apparent activation energy on the bulk VPO catalyst is
about 10 kJ mol1 higher than that on the 18%VPO/1.5P-CeO2 catalyst, suggesting that the nature of the active phases of the supported
VPO catalyst is quite different from that of bulk VPO [49].

From Table 2, we can observe that the treatment of ceria with


NH4 H2 PO4 solution has a signicant impact on the catalytic performance of the catalyst. In contrast to VPO/CeO2 , the VPO/xP-CeO2
catalysts exhibit good catalytic performance, especially superior
MA selectivity and yield. It is well known that the phase structure can determine the catalytic performance of VPO catalyst. As
discussed in the XRD and Raman sections, the vanadium species
formed on ceria and phosphorus-modied ceria are rather different. Over ceria supported VPO catalyst, VOPO4 is the main phase
and the (VO)2 P2 O7 phase is almost undetectable. However, over
P-CeO2 supported VPO catalysts, (VO)2 P2 O7 becomes the major
phase. In the process of precursor formation, V2 O5 rstly reacted
with the iso-butanol to produce vanadium (V) alkoxide and then
the vanadium (V) alkoxide species were reduced by benzyl alcohol
to form insoluble V2 O4 [50]. The V2 O4 intermediate absorbed on
the support would react with H3 PO4 to generate VOHPO4 0.5H2 O.
During this process, the H3 PO4 could interact either with the vanadium species or with the support, depending on the nature of
support [35]. In the case of VPO/CeO2 samples, a certain amount
of the H3 PO4 would interact with ceria, thus causing a shortage
of H3 PO4 that reacted with vanadium species. This could account

H.-Y. Wu et al. / Journal of Molecular Catalysis A: Chemical 414 (2016) 18

for that (VO)2 P2 O7 was hardly formed on these catalysts, since a


slight excess of phosphorous is required to stabilize the V4+ oxidation state and inhibit the overoxidation of V4+ -containing phase
[20]. For VPO/xP-CeO2 samples, however, most of the H3 PO4 would
coordinate to vanadium species to form VOHPO4 0.5H2 O while little H3 PO4 react with the support. The pretreatment of ceria with
NH4 H2 PO4 solution and the formation of CePO4 phase on it, which
is evidenced by Raman results as shown in Fig. 4a, inhibited the
reaction of H3 PO4 with the support during the synthesis of precursor and conduced to the formation of (VO)2 P2 O7 in VPO/xP-CeO2
catalysts.
Generally speaking, the nal catalyst contains a complex mixture of vanadium (IV) and (V) phosphates, such as vanadyl
pyrophosphate ((VO)2 P2 O7 ), vanadyl polyphosphate (VO(PO3 )2 )
and vanadyl orthophosphates (I -, II -, -, - and -VOPO4 ) [7,51].
Crystalline (VO)2 P2 O7 is generally accepted as the active phase for
MA formation [2,52], whereas the role of V5+ species is still a controversial issue. Guliants et al. [51] indicated that - and -VOPO4 in
the VPO catalyst were found to be detrimental to their catalytic performance. Centi et al. [2] considered that V5+ species could play an
important role in the formation of MA, but excess V5+ species would
result in the consecutive oxidation of MA. This can explain the very
low MA selectivity observed for CeO2 -supported VPO catalyst. With
increasing NH4 H2 PO4 concentration adopted in treatment of support CeO2 , the catalysts show an increase in the ratio of V4+ /V5+
and enhancement of MA selectivity. The highest MA selectivity
and MA yield are obtained with the catalyst supported on 1.5PCeO2 . The best performance of this catalyst could be related to its
more suitable phase composition, since an active VPO catalyst for
this reaction requires the coexistence of V4+ species and a suitable
amount of V5+ species [53].
5. Conclusions
In this work, a series of ceria- and P-modied ceria-supported
VPO catalysts were prepared for the partial oxidation of n-butane
to MA. It was found that the nature of the support has a great
inuence on the catalytic performance. When ceria is used as
the support, a rather poor result is obtained and the MA yield is
just 24.4% at 400 C, whereas an excellent catalytic performance is
achieved when phosphorus-modied ceria is used as the support.
The best yield of MA achieved at 400 C could reach up to 57.5% over
the 18%VPO/1.5P-CeO2 catalyst, which was at a high level among
the supported VPO catalysts reported so far. The characterization
results showed that the pretreatment of ceria with NH4 H2 PO4 solution can modulate the nature of the support and conduce to the
formation of (VO)2 P2 O7 loaded on it, thus leading to a higher catalytic performance. Although VPO/P-CeO2 and bulk VPO catalysts
had the similar VPO phase structure, the former catalysts exhibited well-dispersed VPO species and the enhanced reducibility of
V4+ phase, which is responsible for their higher specic activities
(normalized to per unit mass of VPO loading).
Acknowledgements
This work was supported by the National Basic Research Program of China (2010CB732303, 2013CB933102) and the National
Natural Science Foundation of China (21073148, 21033006)
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.molcata.2015.12.
024.

References
[1] E. Bordes-richards, A. Shekari, G.S. Patience, Handbook of Advanced Methods
and Process in Oxidation Catalysis, in: D. Duprez, F. Cavani (Eds.), World
Scientic, 2014, pp. 549585.
[2] G. Centi, F. Trir, J.R. Ebner, V.M. Franchetti, Chem. Rev. 88 (1988) 5588.
[3] L.K. Leong, K.S. Chin, Y.H. Tauq-Yap, Catal. Today 164 (2011) 341346.
[4] Y.H. Tauq-Yap, C.K. Goh, G.J. Hutchings, N. Dummer, J.K. Bartley, Catal. Lett.
141 (2011) 400407.
[5] X.B. Fan, N.F. Dummer, S.H. Taylor, J.K. Bartley, G.J. Hutchings, Catal. Today
183 (2012) 5257.
[6] R.M. Blanco, A. Shekari, S.G. Carrazn, E.B. Richard, G.S. Patience, P. Ruiz, Catal.
Today 203 (2013) 4852.
[7] G.J. Hutchings, A.D. Chomel, R. Olier, J.C. Volta, Nature 368 (1994) 4145.
[8] G.J. Hutchings, J. Mater. Chem. 19 (2009) 12221235.
[9] C.Y. Xiao, X. Chen, Z.Y. Wang, W.J. Ji, Y. Chen, C.T. Au, Catal. Today 9395
(2004) 223228.
[10] R.A. Overbeek, P.A. Warringa, M.J.D. Crombag, A.J. van Dillen, J.W. Geus, Appl.
Catal. A 135 (1996) 209230.
[11] J. Marc, C.C. Ledoux, P.H. Cuong, T. Vincent, K. Kostantinos, M. Patrick, L. Jan, J.
Catal. 203 (2001) 495508.
[12] W.Y. Nie, X.S. Wang, W.J. Ji, Q.J. Yan, Y. Chen, C.T. Au, Catal. Lett. 76 (2001)
201206.
[13] X.K. Li, W.J. Ji, J. Zhao, Z.B. Zhang, C.T. Au, J. Catal. 238 (2006) 232241.
[14] R.M. Feng, X.J. Yang, W.J. Ji, Y. Chen, C.T. Au, J. Catal. 246 (2007) 166176.
[15] G.C. Behera, K. Parida, P.K. Satapathy, RSC Adv. 3 (2013) 48634866.
[16] N.P. Rajan, G.S. Rao, V. Pavankumar, K.V.R. Chary, Catal. Sci.Technol. 4 (2014)
8192.
[17] X.K. Li, W.J. Ji, J. Zhao, Z.B. Zhang, C.T. Au, Appl. Catal. A 306 (2006) 816.
[18] V.V. Guliants, M.A. Carreon, In: J.J. Spivey (Eds.), Catalysis, The Royal Society of
Chemistry (2005) pp. 145.
[19] G.J. Hutchings, J. Mater. Chem. 114 (2004) 33853395.
[20] B.K. Hodnett, Catal. Rev. Sci. Eng. 27 (1985) 373424.
[21] E.A. Lombardo, C.A. Sanchez, L.M. Cornaglia, Catal. Today 15 (1992) 407418.
[22] D. Ye, A. Satsuma, T. Hattor, Y. Mumkami, Catal. Today 16 (1993) 113121.

[23] E. Mikolajska, S.B. Rasmussen, A.E. Lewandowska, M.A. Banares,


Phys. Chem.
Chem. Phys. 14 (2012) 21282136.
[24] Z.Q. Zhou, H.Y. Xu, W.J. Ji, Y. Chen, Catal. Lett. 96 (2004) 221225.
[25] A. Caldarelli, F. Cavani, F. Folo, S. Luciani, C. Cortelli, R. Leanza, Catal.Today 157
(2010) 204210.
[26] J.M.C. Buena, G.K. Bethke, M.C. Kung, H.H. Kung, Catal. Today 43 (1998)
101110.
[27] I. Moreno, N. Navascues, S. Irusta, J. Santamaria, J. Catal. 329 (2015) 479489.
[28] L.Y. Yang, X. Yang, S.Y. Lin, R.X. Zhou, Catal. Sci. Technol. 5 (2015) 26882695.
[29] I.T. Troutus, C.M. Teodorescu, V.I. Prvulescu, I.C. Marcu, ChemCatChem. 5
(2013) 757765.
[30] R. Si, M.F. Stephanopoulos, Angew. Chem. 120 (2008) 29262929.
[31] H.C. Yao, Y.F. Yu Yao, J. Catal. 86 (1984) 254265.
[32] E. Aneggi, M. Boaro, C.D. Leitenburg, G. Dolcetti, A. Trovarelli, J. Alloys Compd.
408412 (2006) 10961102.
[33] I.E. Wachs, J.M. Jehng, G. Deo, B.M. Weckhuysen, V.V. Guliants, J.B. Benziger, S.
Sundaresan, J. Catal. 170 (1997) 7588.
[34] N. Herron, D.L. Thorn, US Patent 5,932,746 (1999).
[35] S.J. Khatib, J.L.G. Fierro, M.A. Banares, Top. Catal. 52 (2009) 14591469.
[36] W.D. Harding, K.E. Birkeland, H.H. Kung, Catal. Lett. 28 (1994) 17.
[37] M.Y. Cho, K.C. Roh, S.M. Park, H.J. Choi, J.W. Lee, Mater. Lett. 64 (2010)
323326.
[38] V.N. Kalevaru, N. Madaan, A. Martin, Appl. Catal. A 391 (2011) 5262.
[39] F. Wang, J.L. Dubois, W. Ueda, Appl. Catal. A 376 (2010) 2532.
[40] F.B. Abdelouahab, R. Olier, N. Guilhaume, F. Lefebvre, J.C. Volta, J. Catal. 134
(1992) 151167.
[41] M.L. Granados, F.C. Galisteo, P.S. Lambrou, R. Mariscal, J. Sanz, I. Sobrados,
J.L.G. Fierro, A.M. Efstathiou, J. Catal. 239 (2006) 410421.
[42] Z.Y. Xue, G.L. Schrader, J. Phys. Chem. B 103 (1999) 94599467.
[43] B.T. Pierini, E.A. Lombardo, Mater. Chem. Phys. 92 (2005) 197204.
[44] S. Luciani, F. Cavani, V.D. Santo, N. Dimitratos, M. Rossi, C.L. Bianchi, Catal.
Today 169 (2011) 200206.
[45] H.Y. Wu, H.B. Wang, X.H. Liu, J.H. Li, M.H. Yang, C.J. Huang, W.Z. Weng, H.L.
Wan, Appl. Surf. Sci. 351 (2015) 243249.
[46] G.W. Coulston, E.A. Thompson, N. Herron, J. Catal. 163 (1996) 122129.
[47] R.A. Overbeek, P.A. Warringa, M.J.D. Crombag, A.J. van Dillen, J.W. Geus, Appl.
Catal. A 135 (1996) 231248.
[48] A. Caldarelli, F. Cavani, F. Folco, S. Luciani, C. Cortelli, R. Leanza, Catal. Today
157 (2010) 204210.
[49] Y. Schuurman, J.T. Gleaves, Catal. Today 33 (1997) 2537.
[50] L.M. Cornaglia, C.A. Snchez, E.A. Lombardo, Appl. Catal. A 95 (1993) 117130.
[51] V.V. Guliants, J.B. Benziger, S. Sundaresan, I.E. Wachs, J.-M. Jehng, J.E. Roberts,
Catal. Today 28 (1996) 275295.
[52] H. Imai, Y.C. Kamiya, T. Okuhara, J. Catal. 225 (2008) 213219.
[53] A.A. Rownaghi, Y.H. Tauq-Yap, F. Rezaei, Ind. Eng. Chem. Res. 48 (2009)
75177828.

You might also like