You are on page 1of 6

Microporous and Mesoporous Materials 116 (2008) 210215

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

Comparison of four catalysts in the catalytic dehydration of ethanol to ethylene


Xian Zhang, Rijie Wang, Xiaoxia Yang *, Fengbao Zhang
School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, PR China

a r t i c l e

i n f o

Article history:
Received 29 November 2007
Received in revised form 18 March 2008
Accepted 2 April 2008
Available online 6 April 2008
Keywords:
Dehydration of ethanol
c-Al2O3
HZSM-5
SAPO-34
NiAPSO-34

a b s t r a c t
The aim of this study was to compare the activity and stability of c-Al2O3, HZSM-5 (Si/Al = 25), silicoaluminophosphate (SAPO-34) and Ni-substituted SAPO-34 (NiAPSO-34) as catalysts in the dehydration of
ethanol to ethylene. c-Al2O3- and HZSM-5 were commercial catalysts. SAPO-34 and NiAPSO-34 molecular
sieves had been synthesized with hydrothermal method in the laboratory, characterized with X-ray powder diffraction (XRD), Infrared Spectroscopy (FT-IR), H2 temperature-programmed reduction (H2TPR)
technique and NH3 temperature-programmed desorption (NH3TPD) technique. The incorporation of
Ni2+ into the SAPO-34 framework generated in NiAPSO-34 sample was proved by XRD, FT-IR and H2
TPR techniques. NH3TPD study had revealed that substitution of Ni2+ for Al3+ in the SAPO-34 framework
led to increase the weak and moderately strong acid strength and give rise to weak acid sites. Dehydration of ethanol was carried out over four catalysts and the results showed that conversion of ethanol and
selectivity to ethylene decreased in the order HZSM-5 > NiAPSO-34 > SAPO-34 > c-Al2O3. As to the stability of catalyst, NiAPSO-34 and SAPO-34 were better than other two catalysts. Considering the activity and
stability of the four catalysts comprehensively, NiAPSO-34 was the suitable catalyst in the dehydration of
ethanol.
2008 Elsevier Inc. All rights reserved.

1. Introduction
Ethylene is an important material for the organic chemistry
industry used in the preparation of polyethylene, ethylene oxide,
ethylene dichloride, etc. Conventionally, it has been commercially
produced by the thermal cracking of liqueed petroleum gas
(LPG) or naphtha [1]. This process is an endothermic reaction
requiring high temperature of 6001000 C. Compared to the conventional route, catalytic dehydration of ethanol to ethylene is
proving attractive as it requires lower temperature and offers higher ethylene yield [2,3]. Moreover, ethanol can be produced from
renewable sources, production of ethylene, therefore, does not depend on petroleum source.
Dehydration of ethanol can take place by two competitive paths.
One is the intramolecular dehydration of ethanol to ethylene and
the other is intermolecular dehydration of ethanol to diethyl ether.
At lower temperature, diethyl ether is produced in signicant quantities, while, at the higher temperature, ethylene is the dominant
product. Ethanol dehydrogenation to produce acetaldehyde can
also occur as a side reaction at the higher temperature [4].
One of the early catalysts employed for the dehydration of ethanol was c-alumina (Al2O3) [5]. With Al2O3 catalyst, dehydration of
ethanol required higher reaction temperature (450 C) and offered
lower ethylene yield (80%). At present, researchers studied on
dehydration of ethanol over HZSM-5 zeolite. Numerous papers
* Corresponding author. Tel./fax: +86 22 27401018.
E-mail addresses: xxytju@hotmail.com, gsarah@eyou.com (X. Yang).
1387-1811/$ - see front matter 2008 Elsevier Inc. All rights reserved.
doi:10.1016/j.micromeso.2008.04.004

had already been published on the subject [6,7]. With HZSM-5 catalyst, at lower temperature (300 C), 95% of ethylene was reported
in the reaction produce at a conversion level of 98% of ethanol. But
because of strong acidity of HZSM-5, catalyst easily occurred on
coking deactivation. Therefore, stability of HZSM-5 was worse.
In 1982, SAPO-34 had been synthesized by UCC [8]. Due to narrow
pores, extending in three dimensions and mild acidity, SAPO-34 was
successfully applied in the MTO (Methanol to Olens) process [9
11]. In previous studies, SAPO-34 was believed to be the best catalyst
in terms of activity and selectivity to light olens [12]. Detailed research revealed that Ni-substituted SAPO-34 (NiAPSO-34) showed
excellent performance for ethylene production from methanol
[1315]. However, for catalytic dehydration of ethanol to ethylene,
application of SAPO-34 and NiAPSO-34 was no reported.
In this work, we had devoted ourselves to investigations to look
for a catalyst capable of exhibiting high efciency as well as stability in the formation of the nal product ethylene. The selectivity
for catalytic dehydration of ethanol to ethylene of four catalysts,
Al2O3, HZSM-5, SAPO-34 and NiAPSO-34, was studied to nd a suited catalyst in the activity and stability.

2. Experimental
2.1. Preparation of catalyst
Commercial Al2O3 and HZSM-5 (Si/Al = 25) were obtained from
Tianjin Research and Design Institute of Chemical Industry and the

211

X. Zhang et al. / Microporous and Mesoporous Materials 116 (2008) 210215


Table 1
Physical properties of the four catalysts
Property
2

Surface area (m /g)


Mean pore radius ()
Cumulative pore volume (ml/g)

Al2O3

HZSM-5

SAPO-34

NiAPSO-34

190
45.5
0.48

295
5
0.22

377
4.7
0.49

358
4.5
0.46

Catalyst Plant of Naikai University, respectively. SAPO-34 and NiAPSO-34 samples were prepared with hydrothermal method in the
laboratory. Pseudoboehmite (Tianjin Research and Design Institute
of chemical Industry), orthophoshporic acid (85wt%, Tianjin Benchmark Chemical Reagent Co. Ltd.), colloidal silica (Beijing Hongxingguangsha Chemical Architectural Material Co. Ltd.), Ni
(NO3)26H2O (Tianjin Yaohua Chemical Reagent Co. Ltd.) and morpholine (Tianjin Guangfu Fine Chemical Research Institute) were
used as the sources of aluminum, phosphorus, silicon, nickel and
template, respectively. The synthesis of SAPO-34 and NiAPSO-34
followed the procedure reported in the Literature [16]. The chemical composition of the starting gel was 1.0Al2O3:1.0P2O5:1.0SiO2:2.0R:50H2O for SAPO-34 and 1.0Al2O3:1.0P2O5:1.0SiO2:
0.0150NiO: 2.0R:50H2O for NiAPSO-34. The precursor gels were
transferred to a stainless-steel autoclave and heated in static condition for 60 h at 200 C under autogeneous pressure. The products
were ltrated, washed, and dried at 110 C for 8 h. The physical
properties of the four catalysts were given in Table 1.

perature-programmed desorption (TPD) proles of pre-adsorbed


NH3, determined by home-made apparatus equipped with a thermal conductivity detector (TCD). The reducibility of samples was
estimated by the temperature-programmed reduction (TPR) technique using home-made apparatus with a TCD. The elemental
analyses of the samples were done by X-ray uorescence (XRF)
in a SHIMADZU XRF-1800 spectrometer with Rh tube as excitation
source. The physical properties of samples were conducted using a
Quantachrome Autosorb-1 physisorption machine, where the surface area, mean pore radius and cumulative pore volume were
calculated.
2.3. Catalytic reaction
Experimental setup used for this study is presented in Fig. 1.
Catalytic reaction studies were performed in electrically heated
xed bed reactor (ID 10 mm, length 420 mm). Ethanol was changed through a micro pump. Usually the reactor charged 3 ml of
catalyst.
The products were analyzed on gas chromatograph (Agilent
6820) with FID using DB-WAX capillary column by programming
the oven temperature between 60 C and 150 C at 15 C/min.
3. Results and discussion
3.1. Characterization of SAPO-34 and NiAPSO-34 samples
The elemental analyses of SAPO-34 and NiAPSO-34 samples
were measured by XRF spectrometer as given in Table 2. The ob-

2.2. Characterization
X-ray diffraction analysis was carried out in a PANLytical XPery
PRO diffractometer in the scan range of 2h between 5 and 50
using Co Ka as source. The peaks were identied with reference
to JCPDS les (Joint Committee on Powder Diffraction Standards).
Unit cell parameters and volume were calculated using a standard
least square renement technique. FT-IR spectra of the samples
were recorded on a Thermo Nicolet NEXUS FT-IR spectrophotometer using KBr pellet. The acidity of samples was estimated by tem-

Table 2
Elemental analyses for SAPO-34 and NiAPSO-34 sample
Sample

Composition in crystals (atomic ratio)


Al

Si

Ni

SAPO-34
NiAPSO-34

1.00
1.00

0.27
0.28

0.72
0.73

0.0084

N2

Reactor

Micro pump
Ethanol

Gas chromatograph

Fig. 1. Catalyst experimental setup.

X. Zhang et al. / Microporous and Mesoporous Materials 116 (2008) 210215

Relative Intensity (a.u.)

212

(b)

(a)
SAPO-34(PDF#47-0429)

10

20

30
CoK / 2 (deg)

40

50

Fig. 2. XRD patterns for as-synthesized samples (a) SAPO-34 and (b) NiAPSO-34.

served framework chemical compositions for SAPO-34 and NiAPOS-34 samples suggested Si substitution for P. In NiAPOS-34 sample, a portion of Ni2+ in gel mixtures was incorporated into the
SAPO-34 framework. This result showed that incorporation of Ni
into SAPO-34 framework was very difcult.
XRD patterns of as-synthesized SAPO-34 and NiAPSO-34 samples are shown in Fig. 2. The position and the intensity of the diffraction peaks of both samples are well-matched with JCPDS les
(PDF#47-0429). The patterns indicated high crystallinity without
obvious impurity. Unit cell parameters and volume were calculated for as-synthesized samples and are listed in Table 3. It was
found that the unit cell parameters and volume increased for NiAPSO-34 sample. The replacement of Al3+ by the larger Ni2+ in the

Table 3
Unit cell parameters and unit cell volume of samples

SAPO-34
NiAPSO-34

Unit cell parameters

Unit cell volume

a/

v/3b/

c/

13.763
13.786

13.763
13.786

14.823
14.851

2431.91
2444.34

Relative Absorbance (a.u.)

Sample (as-synthesized)

tetrahedral SAPO-34 framework causes a slight expansion in the


unit cell parameters and volume which is one of the evidences that
Ni2+ is probably incorporated into the SAPO-34 framework.
FT-IR spectra of the as-synthesized samples in the framework
vibration frequency region are shown in Fig. 3. The vibration frequencies had been assigned in Table 4. As shown in Table 4, framework vibrations were characteristically similar to relevant reports
[17,18]. IR band around 480 cm 1, 530 cm 1, 570 cm 1, 640 cm 1,
730 cm 1, 870 cm 1, 1100 cm 1 exhibited by SAPO-34 sample can
be attributed to stretching mode of [SiO4], (Si,Al)O4, [PO4], D-6
rings, symmetric, protonated morpholine, asymmetric, respectively. The framework vibration frequencies of NiAPSO-34 sample
shifted to position of lower wavenumber, compared with those
of SAPO-34 sample. This is the supporting evidence for the presence of Ni2+ in the SAPO-34 framework.
The H2TPR proles of SAPO-34 and NiAPSO-34 samples are
shown in Fig. 4. It should be noticed that SAPO-34 sample presented one reduction of peak. This peak starting at 1100 C, which
was ascribed to the reduction of Al3+ in the SAPO-34 framework,
formed after the collapse of molecular sieves structures [19]. NiAPSO-34 sample was reduced in two regions, which started at
1000 C and 1100 C, respectively. The appearance of the rst peak

(b)

(a)
0

500

1000

1500

2000

2500

3000

3500

4000

Wavenumber / cm-1
Fig. 3. FT-IR spectra of as-synthesized samples (a) SAPO-34 and (b) NiAPSO-34.

4500

213

X. Zhang et al. / Microporous and Mesoporous Materials 116 (2008) 210215


Table 4
Assignment of the FTIR bands in the framework vibration region of the as-synthesized samples
Sample

IR peaks wavenumber/cm

TO bond

SAPO-34
NiAPSO-34

SiO4

(Si, Al)O4

PO4

D-6 rings

480
453

530
503

570
554

640
614

Symmetric stretch PO (AlO)

Protonated morphiline stretch

Asymmetric stretch OPO

730
714

870
846

1110
1086

220 C

Relative Intensity (a.u.)

Relative Intensity (a.u.)

430 C

(b)

197 C
426 C
407 C

(d)

185 C

(c)
(b)

(a)

(a)
150

400

600

800
1000
Temperature / C

1200

Fig. 4. H2TPR proles of samples (a) SAPO-34 and (b) NiAPSO-34.

attributed to the presence of Ni2+. Reduction temperature of Ni2+ in


the extra framework is lower 800 C. Therefore, the rst peak was
associated to reduction of Ni2+ in the framework. This indicated
that Ni2+ existed in the framework and was not easily reduced.
The TPR proles of samples conrmed the results of XRD and FT-IR.
3.2. Acidic property of the four catalysts
NH3TPD was used to characterize the acidic properties of the
four samples. NH3TPD proles of samples are shown in Fig. 5.
The desorption temperature indicates the acid strength of sample.
The higher desorption temperature is, the stronger the acid
strength is. The acid amounts corresponding to the amounts of adsorbed NH3 were estimated from the areas underlying the NH3
TPD curves. Al2O3 sample showed two peaks appearing from
150 C to 400 C and 400 C to 600 C. They are typically known
as Lewis acid sites [20]. The desorption characteristics of zeolites
are quite different from those of Al2O3. They are known to possess
mainly the Brnsted acid sites. For SAPO-34, two distinctly resolved maxima NH3 desorption peaks sample at 185 C and
407 C, which were indication of the presence of acid sites in two
different strengths, were observed. The desorption peaks at low
and high temperature corresponded to the weak and moderately
strong acid sites in the molecular sieve framework, respectively.
The relative distribution of these two types of acid sites 1:5, as estimated from the integrated areas of the two peaks, suggesting the
presence of a substantial amount of moderately strong acid sites
in the SAPO-34 sample. The rst peak arose due to the weak acid

300
450
Temperature / C

600

750

Fig. 5. NH3TPD proles of samples (a) Al2O3 (b) SAPO-34 (c) NiAPSO-34 and
(d) HZSM-5.

sites present as surface hydroxyl groups. The moderately strong


peak at 407 C arose possibly from structural acidity [17]. As
shown in Fig. 5, compared to SAPO-34 sample, NiAPSO-34 sample
exhibited higher desorption temperature of both weak and strong
acid sites and possessed more weak acid sites due to incorporation
of Ni2+ into the SAPO-34 framework. It was obvious that HZSM-5
sample showed highest desorption temperature of both weak
and strong acid sites and most weak acid sites. This result indicated
that the acidic property of HZSM-5 sample is stronger than that of
SAPO-34 and NiAPSO-34 samples. In general, the acid site was
important as well as the acid strength to catalytic activity for dehydration of ethanol to ethylene.
3.3. Catalytic performance of dehydration of ethanol
Catalytic testing of ethanol dehydration was carried our over
Al2O3, HZSM-5, SAPO-34 and NiAPSO-34. The ethanol conversion
and selectivity to ethylene were measured at different temperature. The results obtained are graphically illustrated in Fig. 6 and
Fig. 7, respectively. Product distribution data at the optimal reaction condition of each catalyst are shown in Table 5. As shown in
Fig. 6, at lower temperature, dehydration of ethanol was incomplete and ethanol conversion was lower. With the increase in temperature, conversion of ethanol from each catalyst increased. At
475 C, the conversion of ethanol was 90.1% for Al2O3 sample. At
375 C, the conversion of ethanol was 96.5% for NiAPSO-34 and
93.5% for SAPO-34. In contrast, HZSM-5 sample possessing strong
acidic property displayed the highest conversion of ethanol, up to

214

X. Zhang et al. / Microporous and Mesoporous Materials 116 (2008) 210215

100

100

80

90

70
60
50

HZSM-5
NiAPSO-34
SAPO-34
Al2O3

40
30
200

250

300

350

400

450

Yield of Ethylene/%

Conversion of Ethanol / %

90

80
HZSM-5
NiAPSO-34

70

SAPO-34
Al2O3

500

Temperature / C

60

Fig. 6. Effect of reaction temperature on conversion of ethanol over different catalysts. (reaction condition: LHSV 3 h 1, ethanol partial pressure 0.7 atm).

20

40

60

80

100

Reaction time /h
Fig. 8. Comparison of the stability of catalysts. (HZSM-5: reaction temperature
300 C, LHSV 3 h 1, ethanol partial pressure 0.7 atm SAPO-34 and NiAPSO-34:
reaction temperature 350 C, LHSV 3 h 1, ethanol partial pressure 0.7 atm Al2O3:
reaction temperature 450 C, LHSV 3 h 1, ethanol partial pressure 0.7 atm).

100

Selectivity to Ethylene / %

90
80
70
60
50
HZSM-5
NiAPSO-34
SAPO-34
Al2O3

40
30
200

250

300

350

400

450

500

Temperature / C
Fig. 7. Effect of reaction temperature on selectivity to ethylene over different catalysts (reaction condition: LHSV 3 h 1, ethanol partial pressure 0.7 atm).

Table 5
Product distribution data at the optimal reaction condition of each catalyst*
Product/%

Catalyst
Al2O3

HZSM-5

SAPO-34

NiAPSO-34

Optimal reaction temperature/C

Ethanol
Ethylene
Diethyl ether
Acetaldehyde
Unknown
*

450

300

350

350

14.4
78.7
6.9
0.1

4.9
93.7
0.5
0.5
0.4

8.8
86
4.9
0.3

6.1
92.3
1.0
0.4
0.2

Reaction condition: LHSV 3 h

, ethanol partial pressure 0.7 atm.

97.3%, at lower temperature (325 C). As shown in Fig. 7, with four


catalysts, the selectivity to ethylene showed the same change tendency. An increase in temperature increased the selectivity to ethylene up to a certain temperature, after which it decreased. At
lower temperature, selectivity to ethylene was lower due to producing outgrowth diethyl ether. As the temperature increased,
the selectivity to ethylene increased with a concomitant decrease

in the selectivity to diethyl ether. At higher temperature, the selectivity to ethylene begun to decrease due to the formation of outgrowth acetaldehyde produced by ethanol dehydrogenation.
With Al2O3, HZSM-5, SAPO-34, NiAPSO-34, selectivity to ethylene
was maximal (91.9%, 98.5%, 94.3%, 98.3%) at 450 C, 300 C,
350 C, 350 C, respectively. HZSM-5 exhibited greater selectivity
to ethylene than any of the catalysts at lower temperature. In summary, the conversion of ethanol and the selectivity to ethylene decreased in the order HZSM-5 > NiAPSO-34 > SAPO-34 > Al2O3.
3.4. The stability of catalysts in the dehydration of ethanol
The stability of Al2O3, HZSM-5, SAPO-34 and NiAPSO-34 was
studied at the optimal reaction condition of each catalyst. As
shown in Fig. 8, Al2O3 catalyst deactivated after reaction for 80 h
and yield of ethylene decreased from 82.8% to 62.1%. HZSM-5 catalyst, which showed the highest initial activity (93.1%), presented
noticeable decrease of yield of ethylene after reaction for 60 h.
The deactivation of H-ZSM catalyst was often reported in the literature [21], and is generally assigned to the possible strong acidic
property of HZSM-5. In contrast, SAPO-34 and NiAPSO-34 exhibited effective catalytic activity and stability, and provided 86.0%
and 92.3% yield of ethylene, respectively, which remained unchanged for 100 h of time on stream. Compared to HZSM-5 catalyst, NiAPSO-34 molecular sieve showed similar activity and
better stability.
4. Conclusions
Dehydration of ethanol to ethylene was studied in this work.
Al2O3, HZSM-5, SAPO-34 and NiAPSO-34 were employed as catalyst in the dehydration of ethanol. Al2O3 and HZSM-5 were commercial catalysts. SAPO-34 and NiAPSO-34 were synthesized in
the laboratory. For NiAPSO-34 sample, unit cell parameters and
volume increased and the framework vibration frequencies shifted
to position of lower wavenumber. In NH3TPD proles of samples,
NiAPSO-34 sample exhibited higher desorption temperature of
both weak and strong acid sites and possessed more weak acid
sites. Incorporation of Ni2+ into the framework generated in NiAPSO-34 sample was proved by XRD, FT-IR and TPR techniques.

X. Zhang et al. / Microporous and Mesoporous Materials 116 (2008) 210215

A comparative study of Al2O3, HZSM-5, SAPO-34 and NiAPSO-34


revealed that for conversion of ethanol and selectivity to ethylene,
the decreasing order of preference of the catalysts was HZSM5 > NiAPSO-34 > SAPO-34 > Al2O3. NiAPSO-34 and SAPO-34 appeared better stability than other catalysts. Synthetically, taking
the activity and stability of catalyst into account, of the four catalysts, NiAPSO-34 was the suitable catalyst in the dehydration of
ethanol.

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]

A. Okagami, S. Matsuoka, US Patent 3541179, 1970.


S. Bun, S. Nishiyama, S. Tsuruya, M. Masai, Appl. Catal. 59 (1990) 13.
T. Zaki, J. Colloid Interf. Sci. 284 (2005) 606613.
P. Berteau, M. Runet, B. Delmon, Acta. Chim. Hung. 124 (1987) 2533.
O. Winter, E. Ming-Teck, Hydrocarb. Process Nov (1976) 125133.
C.B. Phillips, R. Datta, Ind. Eng. Chem. Res. 36 (1997) 44664475.
T.M. Nguyen, R.L.V. Mao, Appl. Catal. 58 (1990) 119129.
J. Haggin, Chem. Eng. News 20 (1983) 36. June.

215

[9] P. Dutta, S.C. Roy, L.N. Nandi, P. Samuel, S.M. Pillai, B.D. Bhat, M.
Ravindranathan, J. Mol. Catal. A: Chem. 223 (2004) 231235.
[10] L. Yun-Jo, B. Seung-Chan, J. Ki-Won, Appl. Catal. A: General 329 (2007) 130
136.
[11] D.R. Dubois, D.L. Obrzut, J. Liu, J. Thundimadathil, P.M. Adekkanattu, J.A. Guin,
A. Punnoose, M.S. Seehra, Fule Process. Technol. 83 (2003) 203218.
[12] S. Wilson, P. Barger, Micropor. Mesopor. Mater. 29 (1999) 117126.
[13] M. Inoue, P. Dhupatemiya, S. Phatanasri, T. Inui, Micropor. Mesopor. Mater. 28
(1999) 1924.
[14] M. Kang, J. Mol. Catal. A: Chem. 150 (1999) 205212.
[15] M. Kang, J. Mol. Catal. A: Chem. 160 (2000) 437444.
[16] B.M. Lok, C.A. Messina, Ossining, R.L. Patton, Katonah, R.T. Gajek, T.R. Cannan,
US Patent 4,440871, 1984.
[17] A.M. Parakash, S. Unnikrishnau, J. Chem. Soc. Faraday Trans. 90 (15) (1994)
22912296.
[18] S. Ashtekar, S.U.V. Chilukuri, D.K. Chakrabarty, J. Phys. Chem. 98 (1994) 4878
4883.
[19] R. Roque-Malherbe, R. Lopez-Cordero, J.A. Gonzales-Morles, J. Onate-Martinez,
M. Carreras-Gracial, Zeolite 13 (1993) 481484.
[20] F.S. Ramos, A.M. Duarte de Farias, L.E.P. Borges, J.L. Monteiro, M.A. Fraga, E.F.
Sousa-Aguiar, L.G. Appel, Catal. Today 101 (2005) 3944.
[21] A.T. Aguayo, A.G. Gayubo, A. Atutxa, B. Valle, J. Bilbao, Catal. Today 107108
(2005) 410416.

You might also like