Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Chemical Engineering Journal 171 (2011) 1387–1398

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Kinetic modeling of ethylbenzene dehydrogenation over hydrotalcite catalysts


Luqman Atanda, Nabil Al-Yassir, Sulaiman Al-Khattaf ∗
KAUST Center in Development, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia

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

Article history: Kinetics of ethylbenzene dehydrogenation to styrene was investigated over a series of quaternary mixed
Received 30 January 2011 oxides of Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni) catalysts prepared by calcination of hydrotalcite-
Received in revised form 13 April 2011 like compounds and compared with commercial catalyst. The study was carried out in the absence
Accepted 12 May 2011
of steam using a riser simulator at 400, 450, 500 and 550 ◦ C for reaction times of 5, 10, 15 and 20 s.
Mg3 Fe0.25 Mn0.25 Al0.5 afforded the highest ethylbenzene conversion of 19.7% at 550 ◦ C. Kinetic parame-
Keywords:
ters for the dehydrogenation process were determined using the catalyst deactivation function based
Hydrotalcite
on reactant conversion model. The apparent activation energies for styrene production were found to
Dehydrogenation
Ethylbenzene
decrease as follows: E1-Ni > E1-Co > E1-Mn .
Kinetic modeling © 2011 Elsevier B.V. All rights reserved.

1. Introduction for certain base-catalyzed reactions [8,9], thus they are activated
by thermal decomposition. Therefore, the resultant oxides after
The increasing application of styrene monomer in the produc- calcinations offer improved catalytic performance due to the for-
tion of polystyrene, ABS resins, synthetic rubbers, etc. has made mation of small crystal size stable to thermal treatments. The
the demand for styrene increase tremendously with current pro- oxides also possess high surface area, basic properties and further
duction capacity exceeding 25 × 106 tonnes/year [1]. Styrene is form small and thermally stable metal crystallites by reduction
manufactured commercially through two process routes – cat- [10]. Moreover, they are potentially recoverable and recyclable
alytic dehydrogenation of ethylbenzene and coproduction with [11]. Mixed oxides of thermally decomposed HT precursors are
propylene oxide from the reaction of propene with oxidized ethyl- potential catalysts with growing applications in various reactions
benzene to ethylbenzenedroperoxide, with the former accounting as reviewed by Cavani et al. [10]. HTs of several reducible biva-
for 90% of world production capacity. Catalytic dehydrogenation lent (Ni, Cu, Co) and trivalent (Fe, Cr) cations in their structure
of ethylbenzene (EB) is carried out over supported and unsup- together with the classical ones (Mg, Zn, Al) serves as precur-
ported Fe-based catalyst in the presence of superheated steam sors for the preparation of different mixed oxides active for
at temperature of 600–700 ◦ C. The process is carried out at high oxidation and hydrogenation/dehydrogenation reactions [12]. Sup-
steam-to-hydrocarbon ratio to increase activity, selectivity and sta- ported iron catalyst prepared from Mg–Fe–Al HTlcs as precursor
bility of the catalyst [2]. However, the use of excess steam consumes has been successfully applied in the EB dehydrogenation to styrene
high energy which has been estimated to be 1.5 × 109 cal/styrene [6]. Upon calcinations, mixed oxides with a high surface area as
tonne [3]. In addition to that, it was found that most examined cata- well as mesoporous character were formed. XRD analysis indicated
lysts suffered from low stability, fast deactivation and subsequently mainly the presence of periclase Mg(Fe, Al)O phase. Reduction pre-
low catalytic activities [4–7]. treatment afforded the catalyst with somewhat higher activity as
Therefore, there is need to develop new catalyst systems that well as higher stability. EB conversion of 60% and a styrene selec-
allows for uniform dispersion of the active phase as well as high tivity of 95% were stably observed over Mg3.0 Fe0.5 Al0.5 at 550 ◦ C. A
stability. Hydrotalcite-like compounds (HTlcs) as precursors for recent study on the activity of various FeOx –MeOy /Mg(Al)O cat-
mixed oxide catalysts offer unique basic properties that make alysts derived from HTs showed FeOx –CoOy /Mg(Al)O have high
them very attractive for catalytic applications. Fresh hydrotalcites activity due to the formation of Fe–Co bimetallic species [13]. Addi-
(HTs) have high water content which often makes them inactive tion of Co further facilitated the reduction–oxidation of Fe3+ /Fe2+ ,
and the active Fe3+ species was stabilized. The role of MgO as a
base support has been reported effective to initiate dehydrogena-
tion by H+ abstraction on the basic site, and accelerated by the
Abbreviations: EB, ethylbenzene; cat, catalyst; DEB, diethylbenzene; HT, hydro-
talcite; HTlc, hydrotalcite-like compound.
reduction-oxidation of Fe3+ active species [14].
∗ Corresponding author. Tel.: +966 3 860 1429; fax: +966 3 860 4234. Carrà and Forni [15] performed kinetic studies in the tem-
E-mail address: skhattaf@kfupm.edu.sa (S. Al-Khattaf). perature range of 770–900 K over the industrial catalyst, Shell

1385-8947/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.05.053
1388 L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398

increases styrene selectivity and consequently reduces the yield


Nomenclature of toluene due to the removal of hydrogen through the mem-
branes. It was shown that increase in styrene production over that
Ai pre-exponential factor for the ith reaction (m3 /kg of of the industrial fixed bed unit is achievable, if design and operating
catalyst s) parameters are properly chosen. In view of this, the present study
Ci concentration of specie i in the riser simulator is aimed at investigating the kinetics of EB dehydrogenation over
(mol/m3 ) quaternary mixed metal oxide catalysts of Mg3 Fe0.25 Me0.25 Al0.5
CL confidence limit (Me = Co, Mn and Ni) obtained from HT precursors in a CREC flu-
Ei apparent activation energy of the ith reaction idized riser simulator operated at temperatures below that of the
(kJ/mol) industrial process, as previous studies [6] have shown that opti-
Kc equilibrium constant mum condition was achieved at 550 ◦ C. Injection of the reactant and
ki apparent rate constant for the ith reaction (m3 /kg of withdrawal of products were for short period of time. It was envis-
catalyst s) aged that for short reaction time, the potential of side reactions
k0i pre-exponential factor for the ith reaction after re- taking place will be limited. Kinetic modeling of the dehydrogena-
parameterization (m3 /kg of catalyst s) tion process was done using the catalyst activity decay function
MWi molecular weight of specie i based on reactant conversion.
r2 correlation coefficient
ri rate of reaction for species i
2. Experimental
R universal gas constant (kJ/kmol K)
t reaction time (s)
2.1. Materials
T reaction temperature (K)
To average temperature of the experiment
All the precursors used in the synthesis of the HT quater-
V volume of the riser (45 cm3 )
nary mixed oxides Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni) were
Wc mass of the catalyst (0.81 g)
obtained from Sigma–Aldrich and used as supplied. The chemical
Whc total mass of the hydrocarbon injected the riser
feedstock (ethylbenzene, 99%) was also supplied by Sigma–Aldrich.
(0.162 g)
yi mass fraction of ith component
2.2. Catalyst preparation

Greek letters
The HT precursors of Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and
ϕ apparent deactivation function
Ni) catalysts were prepared by co-precipitation of metal nitrates,
 catalyst deactivation constant (RC model)
following the method by Miyata and Okada [22] after minor modi-
fications. An aqueous solution containing the nitrates of Mg2+ , Fe3+ ,
Me2+ (Co2+ , Mn2+ or Ni2+ ) and Al3+ (200 ml) was added slowly with
105. The intrinsic rate of styrene formation was developed, based vigorous stirring into an aqueous solution of sodium carbonate
upon Langmuir–Hinshelwood kinetics with activation energy of (400 ml). The pH of the solution was adjusted at 10.0 by dropping
45 kcal/mol. Coulter et al. [16] studied the kinetics using unpro- a 1 M aqueous solution of sodium hydroxide, leading to a precipi-
moted and K-promoted polycrystalline catalysts. The unpromoted tation of heavy slurry. After the solution was aged at 60 ◦ C for 24 h,
catalyst yielded apparent activation energy of 39 kcal/kmol while the solution was filtrated and the precipitates were washed with
for K-promoted catalyst, the activation energy was 36 kcal/mol. de-ionized water (1000 ml), dried in air at 100 ◦ C for 4 h, and cal-
The authors concluded that the active sites of unpromoted and cined at 550 ◦ C for 12 h in a muffle furnace in a static air atmosphere.
promoted catalysts are identical which was later confirmed by The concentration of Na2+ in the catalysts after the calcination was
Shekhah et al. [17] As found in the work of Addiego et al. [18], the confirmed to be less than 10 ppm by atomic absorption.
increase of potassium loading initially decreases the apparent acti-
vation energy to 21 kcal/mol while further addition of potassium 2.3. Catalyst characterization
led to an increase of the apparent activation energy to 34 kcal/mol.
Shekhah et al. [17] suggested that the increase of potassium loading All catalyst characterizations were performed for calcined cata-
leads to the decrease in initial conversion rate due to the cov- lysts, which were calcined at 550 ◦ C for 12 h.
erage of active sites by excess potassium. Miyakoshi et al. [19] The chemical compositions of synthesized HT quaternary mixed
studied the effect of some transition metals on Fe-K catalyst pre- oxides of Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni) were deter-
pared through sol–gel method; Mn was reported to enhance the mined by atomic absorption spectroscopy, using the Perkin-Elmer
catalytic activity and suppressed coke formation. The activation equipment (Model AAnalyst 100).
energies determined from the Arrhenius plot are 22.4 kcal/mol and Surface areas were calculated by the Brunauer–Emmett–Teller
21.9 kcal/mol for 20% Mn-substituted Fe-K oxide and Fe-K oxide, (BET) method. N2 adsorption–desorption isotherms at −196 ◦ C
respectively. The authors concluded that the increased catalytic were measured using a conventional volumetric apparatus (Bel
activity resulted from the increased number of active sites as it Japan, BELSORP Mini). Adsorption measurements were done after
was observed that the difference in the activation energies was samples (ca. 0.1 g) were heated at 400 ◦ C for 10 h under nitrogen
not appreciable. Lee and Froment [20] developed a set of intrinsic flow.
rate equations for a commercial potassium-promoted iron cata- XRD was recorded on a Mac Science MX18XHF-SRA pow-
lyst based on Hougen–Watson model. Apparent activation energy der diffractometer with monochromatized Cu K␣ radiation
of 41.9 kcal/mol was obtained. The model was applied in the sim- ( = 0.154 nm) at 40 kV and 30 mA. The diffraction pattern was iden-
ulation of the dehydrogenation in industrial multibed adiabatic tified through comparison with those included in the JCPDS (Joint
reactors with either axial or radial flow, or accounting also for Committee of Powder Diffraction Standards) database and litera-
thermal radical-type reactions, internal diffusion limitations, coke tures.
formation, and gasification. H2 -TPR of the calcined catalysts was carried out by flow-
The use of fluidized bed has been suggested to improve EB ing 5 vol.% H2 /N2 (30 cm3 min−1 ) in the temperature range of
conversion [21]. The application of selective membranes further 100–900 ◦ C. The sample temperature increased with a rate of
L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398 1389

Table 1
Specific surface area, metal composition, and XPS analytical data of Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni).

Catalyst SBET (m2 gcat −1 ) Molar ratioa Binding energyb (eV) Surface molar ratiob,c

Mg/Fe Mg/Al Mg/Me Mg 2p3/2 Fe 2p3/2 Me 2p3/2 Mg/Fe Mg/Me

Mg3 Fe0.25 Co0.25 Al0.5 187 11.5 5.5 11.5 48.9 710.6 779.6 6.5 5.0
Mg3 Fe0.25 Mn0.25 Al0.5 158 11.1 5.1 11.6 49.0 710.6 641.5 7.1 8.3d
Mg3 Fe0.25 Ni0.25 Al0.5 197 10.2 5.8 11.3 49.0 710.6 853.8 8.0 2.4
a
By atomic absorbance analyses.
b
By XPS analyses.
c
After ASF correction.
d
The data scattered between 6.25 and 12.5.

10 ◦ C min−1 and the amount of H2 consumed was monitored by a increased to ∼35 psi due to increased number of molecules, which
thermal conductivity detector of a gas chromatograph (Shimadzu, is as a result of dehydrogenation and cracking reaction. A sudden
GC-8AIT). drop in reactor pressure was noticed after the stipulated time of
XPS measurements were performed on a Perkin Elmer 1600E reaction. The quantitative analysis of the reaction products were
spectrometer using Mg K␣ radiation as excitation source. In charge- carried out online using a Agilent GC with FID (Agilent Chromato-
up correction, the calibration of binding energy (BE) of the spectra graph Model 6890N), equipped with an HP-INNOWax capillary
was referenced to the C 1s electron bond energy corresponding to column (polyethylene glycol (PEG)) (length 60 m × internal diam-
graphitic carbon at 284.5 eV. In addition, relative atomic sensitivity eter 0.32 mm × film thickness 0.50 ␮m). During the course of the
factors (ASF) were corrected to determine practically more accurate investigation, a number of runs were repeated to check for repro-
chemical compositions on the surface. ducibility in the experiment results, which were found to be
excellent. Typical errors were in the range of ±2%.
2.4. Catalytic experiments The ethylbenzene (EB) conversion (mol% or wt%) is expressed as
follows:
2.4.1. Experimental set-up (riser simulator) moles of converted ethylbenzene
All of the experimental runs were carried out in the riser simu- XEB = × 100
moles of ethylbenzene fed
lator reactor. This reactor is novel bench-scale equipment with an
internal recycle unit invented by de Lasa [23]. The detailed descrip- The selectivity (to product i) is expressed as follows:
tion of the simulator riser has been previously described [24]. number of carbon atoms of product i
Si = × 100
number of carbon atoms of converted products
2.4.2. Testing procedures and product identifications
In a typical experiment, the reactor was charged with 0.8 g of The yield (to product i) is expressed as follows:
calcined catalyst previously crushed and sieved to a particle size Si
of ∼ 60 ␮m diameter. The catalyst/reactant ratio was 5 (weight Yi = × 100
XEB
of catalyst = 0.81 g, weight of reactant injected = 0.162 g). The reac-
tion temperature was varied in the range of 400–550 ◦ C, and the
reaction time was varied from 5 to 20 s. For all experimental runs 2.4.3. Coke analysis
at pre-determined temperature, pressure changes are monitored The amount of coke deposited on the spent catalysts was deter-
through pressure transducers installed in both the reactor and the mined by a common combustion method using a carbon analyzer
vacuum box chambers. The feedstock was injected into the reac- multi EA 2000 (Analytikjena). The multi EA 2000 with CS module
tor at ∼14.7 psi (atmospheric pressure). The feedstock vaporizes is a specially developed system to permit simultaneous or separate
almost instantaneously, causing an abrupt increase in pressure to determination of the total carbon and total sulfur in samples of
∼26 psi. After vaporization of the reactant, the pressure further solids, pastes, and liquids by means of high temperature oxidation

A B

c c
Intensity (a.u.)
Intensity (a.u.)

b b

a a

0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100

2 theta (°) 2 theta (°)

Fig. 1. XRD patterns of Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni) after (A) drying and (B) calcination at 550 ◦ C (: hydrotalcites; : unknown and : periclase).
1390 L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398

Toluene + CH 4

-1
50 µmol g cat +H2 k3
k1
EB Styrene + H 2
e
k-1
k2
H2 consumed

d
Benzene + C 2H 4

c Scheme 1. Catalytic reaction scheme of ethylbenzene (EB) dehydrogenation.

b the ability of Mg–Al and Mg–Fe to form the hydrotalcite structure,


resulting in large surface area after the calcination. The relatively
low SBET (158 m2 g−1 ) of Mg3 Fe0.25 Mn0.25 Al0.5 compared to other
catalysts is probably due to plugging of the pores with segregated
a bulk manganese oxide particles. This was suggested by XRD and
H2 -TPR, which are discussed in the following paragraphs. It is also
0 100 200 300 400 500 600 700 800 900 1000
worth mentioning that the metal compositions are in complete
Temperature / (°C) agreement with the nominal values. This is due to the formation of
Fig. 2. H2 -TPR of Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn, and Ni) catalysts. (a)
hydrotalcite structure as the precursors during the coprecipitation,
Mg3 Fe1 ; (b) Mg3 Fe0.5 Al0.5 ; (c) Mg3 Fe0.25 Mn0.25 Al0.5 ; (d) Mg3 Fe0.25 Co0.25 Al0.5 ; and (e) which accommodates many metal cations in the structure.
Mg3 Fe0.25 Ni0.25 Al0.5 . XRD patterns of Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni) cat-
alysts after drying are shown in Fig. 1A. It can be noted that all
Mg3 Fe0.25 Me0.25 Al0.5 showed typical hydrotalcite reflections, indi-
in a current of oxygen supplied to the unit directly. A small amount
cating that the second metal components were incorporated into
of the spent catalyst (0.09–0.1 g) is used for the analysis. The coke
the hydrotalcite structure except Mg3 Fe0.25 Mn0.25 Al0.5 . Mn2+ pos-
laid out on the sample during reaction experiments is burned com-
sesses a larger ionic radii of 0.80 Å than the other Me2+ [25] and
pletely, converting the carbonaceous deposit into carbon dioxide.
requires the highest pH for its complete precipitation as hydrox-
The amount of coke formed is determined by measuring the num-
ides among the Me cations used [10]. After the calcination at 550 ◦ C
ber of moles of CO2 released. It is worth noting that hydrogen is also
(Fig. 1B), all samples showed the periclase reflections together with
produced and converted into water. However, this method does not
weak spinel reflections. This indicates that iron was incorporated
detect hydrogen, and thus it is difficult to determine the C/H and
in the periclase and the spinel.
consequently comment on the nature of coke.
H2 -TPR data of Mg3 Fe0.25 Me0.25 Al0.5 (Me = Mn, Ni, and Co) are
shown in Fig. 2. Mg3 Fe1 showed a reduction peak at 418 ◦ C and
3. Results and discussions a broad peak above 650 ◦ C (Fig. 2a). The first peak is attributed to
the reduction of spinel (MgFe3+ 2 O4 ) to periclase (Mg1−x Fe2+ x O) and
3.1. Physico-chemical properties the latter is to the reduction of Mg1−x Fe2+ x O to Fe0 , as reported
by Shen et al. [26] Mg3 Fe0.25 Ni0.25 Al0.5 showed a weak reduction
The textural properties along with metal composition of peak at 353 ◦ C and broad reduction peak above 800 ◦ C. The latter
Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni) catalysts after the cal- is attributed to the reduction of Ni2+ in the Mg(Ni)O periclase [27].
cination at 550 ◦ C are shown in Table 1. It can be clearly seen Therefore, the one observed at 353 ◦ C can be attributed the reduc-
from Table 1 that all Mg3 Fe0.5 Me0.5 Al0.5 catalysts exhibit large SBET tion of Fe3+ species. Mg3 Fe0.25 Mn0.25 Al0.5 exhibited rather strong
between 150 and 200 m2 gcat −1 . This can be mainly attributed to two reduction peaks at 257 ◦ C and 428 ◦ C (Fig. 2c). This catalyst

Table 2
Product distribution (wt%) at various reaction conditions for the dehydrogenation of ethylbenzene over Mg3 Fe0.25 Co0.25 Al0.5 mixed oxides catalyst.

Temp. (◦ C) Reaction time (s) EB conv. (%) Styrene Benzene Toluene Gases DEB

400 5 3.18 2.58 0.44 0.09 – 0.12


10 4.66 3.66 0.74 0.08 – 0.18
15 5.19 3.85 0.89 0.11 0.06 0.22
20 5.51 3.79 1.10 0.12 0.06 0.38

450 5 5.48 4.14 0.85 0.16 0.09 0.13


10 6.47 4.82 1.14 0.23 0.12 0.16
15 6.61 4.48 1.49 0.19 0.13 0.25
20 7.12 4.44 1.58 0.18 0.14 0.41

500 5 6.99 4.89 1.45 0.23 0.19 0.15


10 8.55 5.67 1.85 0.24 0.24 0.21
15 10.71 6.71 2.69 0.33 0.44 0.24
20 11.7 7.35 2.99 0.37 0.39 0.37

550 5 8.65 5.83 1.95 0.24 0.33 0.22


10 11.38 7.11 2.83 0.36 0.48 0.17
15 12.83 7.84 3.39 0.42 0.56 0.25
20 15.46 8.84 4.56 0.60 0.88 0.34
L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398 1391

Table 3
Product distribution (wt%) at various reaction conditions for the dehydrogenation of ethylbenzene over Mg3 Fe0.25 Mn0.25 Al0.5 mixed oxides catalyst.

Temp. (◦ C) Reaction time (s) EB conv. (%) Styrene Benzene Toluene Gases DEB

400 5 4.06 3.76 0.36 – – –


10 5.82 5.19 0.49 – – 0.04
15 6.25 5.65 0.55 – – 0.05
20 6.64 5.93 0.65 – – 0.06

450 5 5.64 4.97 0.55 0.05 – –


10 9.00 7.58 1.11 0.11 0.08 0.07
15 8.98 7.44 1.12 0.09 0.08 0.13
20 10.19 8.25 1.28 0.12 0.07 0.19

500 5 8.70 6.96 1.40 0.12 0.14 –


10 10.65 8.33 1.77 0.18 0.15 0.07
15 12.93 9.69 2.39 0.21 0.20 0.15
20 17.25 12.01 3.03 0.39 0.25 0.24

550 5 11.46 8.13 2.59 0.25 0.30 –


10 15.02 10.27 3.58 0.39 0.41 0.13
15 18.63 11.50 4.66 0.58 0.54 0.16
20 19.70 12.28 4.95 0.70 0.54 0.19

exhibited unknown reflections together with those of hydrotalcite 3.2. Catalytic activity
after drying (Fig. 1B), suggesting that manganese oxides were sep-
arated from periclase after calcination. It was reported that a mixed The results from the dehydrogenation of EB over
oxide of Mg0.2 Mn1.8 Al1 composition derived from hydrotalcite and Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni) catalysts at differ-
calcined at 500 ◦ C showed two reduction peaks at 333 ◦ C and 460 ◦ C ent reaction temperatures and reaction times are presented in
[28]. This also suggests that manganese oxides were separated from Tables 2–4. The experimental results showed styrene to be the
periclase structure in Mg3 Fe0.25 Mn0.25 Al0.5 and showed the reduc- main reaction product, although small amounts of benzene were
tion peaks at 257 ◦ C and 428 ◦ C. obtained as by product (Fig. 3). Traces of toluene, gaseous hydro-
The XPS analytical results of Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn carbons, diethylbenzene (DEB) were also observed; however the
and Ni) after calcination at 550 ◦ C are summarized in Table 1. The yields of these products were neglected in subsequent analysis. A
surface molar ratios of Mg/Fe and Mg/Me calculated from the XPS possible reaction scheme is shown in Scheme 1. In all the catalysts
results exhibited lower values than the bulk molar ratios obtained investigated, EB conversion increased with both reaction temper-
by AA analyses (Table 1). This indicates that both Fe and second ature and time and maximum conversion of approximately 15.5,
component metal (Me) species are localized in the surface layer of 19.7, and 18.0% was attained at 20 s of time on stream and temper-
the catalyst particles. On all Mg3 Fe0.25 Me0.25 Al0.5 catalysts except ature of 550 ◦ C for Mg3 Fe0.25 Co0.25 Al0.5 , Mg3 Fe0.25 Mn0.25 Al0.5 , and
Mg3 Fe0.25 Ni0.25 Al0.5 , Fe species showed similar surface enriched Mg3 Fe0.25 Ni0.25 Al0.5 catalysts respectively as shown in Fig. 4(A–C).
localization, i.e., almost 1.4 times higher concentration in the sur- The mixed oxides catalyst consisting of Mn gave the highest activ-
face layer than the bulk phase. Surface enrichment of Fe was not ity regardless of its low surface area. This may be accounted for by
so significant for Mg3 Fe0.25 Ni0.25 Al0.5 . Contrarily, some of the sec- the isolation of manganese oxides from the periclase structure, as
ond component metals exhibited far enriched distribution in the suggested by XRD, XPS and H2 -TPR. Therefore, the inhomogeneous
surface layer, i.e., the highest concentration in the surface layer dispersion of Mn on the catalyst (i.e. presence of segregated (bulk)
was obtained by Ni (5 times), followed by Co (2.2 times) and Mn Mn oxide) is assumed to increase the active sites of the catalyst.
(1.3 times). The scattered experimental values for Mn clearly sug- One may suggest that the segregated Mn oxide further contributes
gest that Mn dispersion on the catalyst was not homogeneous. to the basicity of MgO which has been reported to initiate H+
Detailed synthetic approaches and physico-chemical properties of abstraction [14]. The investigation of iron oxide catalysts contain-
these hydrotalcite based materials are given in Balasamy et al. [13]. ing various transition metal oxides by Hirano [29] also showed

Table 4
Product distribution (wt%) at various reaction conditions for the dehydrogenation of ethylbenzene over Mg3 Fe0.25 Ni0.25 Al0.5 mixed oxides catalyst.

Temp. (◦ C) Reaction time (s) EB conv. (%) Styrene Benzene Toluene Gases DEB

400 5 2.14 1.70 0.35 0.05 – 0.06


10 3.23 2.39 0.58 0.07 – 0.21
15 3.86 2.64 0.76 0.08 – 0.27
20 4.21 2.78 0.89 0.12 0.05 0.32

450 5 4.08 3.17 0.65 0.07 0.08 0.07


10 4.82 3.36 0.88 0.13 0.07 0.23
15 6.25 4.24 1.23 0.14 0.12 0.32
20 7.01 4.59 1.52 0.17 0.17 0.52

500 5 5.75 4.45 0.99 0.12 0.15 0.06


10 8.14 5.58 1.68 0.19 0.25 0.24
15 9.74 6.26 2.23 0.25 0.31 0.40
20 10.87 7.05 2.46 0.47 0.35 0.36

550 5 10.04 6.86 2.12 0.35 0.39 0.13


10 13.27 8.74 3.03 0.44 0.55 0.20
15 16.17 10.18 3.82 0.67 0.67 0.31
20 18.04 11.16 4.31 0.86 0.76 0.37
1392 L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398

Fig. 3. Reactions occurring during EB dehydrogenation over Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni).

that increase in basicity due to reduction in electronegativity of the to be as a result of increase in acidity of the catalyst due to loss of
promoter gave rise to the dehydrogenation activity of the catalyst. potassium from the catalyst surface.
As a control, a commercial catalyst was also tested. It showed a low It is also worth mentioning that Mn based catalyst exhibited
activity as shown in Fig. 5 which can be attributed to the absence much lower selectivity to diethylbenzene (DEB, three isomers) than
of steam. the other two catalysts (Co and Ni). Therefore, it can be inferred
from these results that Mn does not favor the disproportionation
reaction. For instance, the selectivity to DEB (temperature = 400 ◦ C
and residence time = 20 s) over Mn, Co, and Ni based oxides were
3.2.1. Styrene/benzene selectivity 1.0 wt%, 6.9 wt% and 7.6 wt%, respectively (Tables 2–4).
The product selectivity during the dehydrogenation of EB over
Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni) and commercial cata-
lysts were compared in Fig. 6 at constant conversion level of ∼6%. 3.2.2. Styrene/benzene yield
The results show that styrene is obtained as the most predominant The yield of styrene was found to increase with both
product on all the HT catalysts with the best selectivity observed temperature and time, in a similar manner with EB con-
in Mg3 Fe0.25 Mn0.25 Al0.5 . The selectivity towards styrene exhibits version. Maximum styrene yield of approximately 8.8, 12.3,
a trend opposite to EB conversion for all the HT catalysts inves- and 11.2% for Mg3 Fe0.25 Co0.25 Al0.5 , Mg3 Fe0.25 Mn0.25 Al0.5 and
tigated. It decreased when the reaction temperature was raised Mg3 Fe0.25 Ni0.25 Al0.5 catalysts, respectively were attained at
due to cracking reaction. Fig. 7 shows selectivity as a function of reaction temperature of 550 ◦ C and 20 s time on stream.
the EB conversion for all the reaction temperatures for Mn substi- Mg3 Fe0.25 Mn0.25 Al0.5 gave the maximum styrene yield, and its vari-
tuted catalyst, showing a maximum styrene selectivity of 92.7% at ation with temperature is shown in Fig. 8A. Similarly, benzene yield
reaction temperature of 400 ◦ C which corresponds to 4.9% EB con- was equally increasing with both temperature and time with a max-
version. However, benzene selectivity increased as the temperature imum of approximately 4.6, 4.9, and 4.3% for Mg3 Fe0.25 Co0.25 Al0.5 ,
was raised. This confirmed that cracking of EB to benzene becomes Mg3 Fe0.25 Mn0.25 Al0.5 , and Mg3 Fe0.25 Ni0.25 Al0.5 catalysts, respec-
pronounced at high temperature as shown in Fig. 7. A maximum tively at 550 ◦ C for reaction time of 20 s. Fig. 8B shows the variation
benzene selectivity of 25.1% was attained at 550 ◦ C corresponding of benzene yield with temperature for Mg3 Fe0.25 Mn0.25 Al0.5 . It can
to 19.7% EB conversion. On the contrary, the commercial catalyst thus be concluded that for longer reaction time, EB undergoes both
has low selectivity towards styrene which was due to significant dehydrogenation and dealkylation reactions to give styrene and
cracking and such an occurrence has been suggested by Lee [30] benzene, respectively. It can also be seen from Tables 2–4 that both
L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398 1393

18 25
400 C 400 C
450 C A B
450 C
16 500 C 500 C
550 C
Ethylbenzene Conversion (%)

20 550 C
14 400 C Pred

Ethylbenzene Conversion (%)


400 C Pred
450 C Pred 450 C Pred
12 500 C pred 500 C Pred
550 C Pred 550 C Pred
15
10

8
10
6

4
5
2

0 0
0 5 10 15 20 25 0 5 10 15 20 25
Reaction Time (s) Reaction Time (s)

20
400 C C
18 450 C
500 C
550 C
Ethylbenzene Conversion (%)

16 400 C Pred
450 C Pred
14 500 C Pred
550 C Pred
12

10

0
0 5 10 15 20 25
Reaction Time (s)
Fig. 4. Conversion of ethylbenzene with respect to time at various temperatures over different hydrotalcites mixed oxides: (A) Mg3 Fe0.25 Co0.25 Al0.5 , (B) Mg3 Fe0.25 Mn0.25 Al0.5 ,
and (C) Mg3 Fe0.25 Ni0.25 Al0.5 .

the yield of DEB (disproportionation product) and toluene (cracking approach is more sound than time-on-stream approach which has
product) increased with temperature. been widely used in literature to account for catalyst deactivation.
Full description of the development of this mathematical model
3.2.3. Coke content measurement is shown in Ref. [31]. In developing the kinetic model, only the
Table 5 shows the amount of coke deposited for the catalysts most important reactions taking place were considered. Toluene
investigated at 550 ◦ C. The data revealed low coke yield, and the formation was not considered owing to its very low concentra-
ratio of coke weight percent to percent conversion is small. Simi- tion in the reaction product. It can thus be assumed that due to
larly, coking on the commercial catalyst was low. This implies that the short reaction time, the styrene produced did not undergo suc-
dehydrogenation of ethylbenzene over the catalysts is not accom- cessive cracking to toluene. Hence, reaction in Scheme 1 has been
panied by substantial coke deposition. It can thus be inferred that adapted, neglecting toluene formation, in determining the kinetic
loss of activity of this commercial catalyst in the absence of steam parameters.
was not due to coking but as a result of solid state transformation, The following set of species balances and catalytic reactions can
though the exact deactivation mechanism was not studied. be written:

3.3. Kinetic modeling Rate of EB disappearance, rEB

3.3.1. Model formulation V dCEB


− = (k1 CEB − k−1 CST CH + k2 CEB )ϕ (1)
Generally, the accepted catalytic reaction scheme for EB dehy- Wc dt
drogenation is shown in Scheme 1. The experimental results were
modeled using catalyst deactivation function model based on Rate of styrene appearance, rST
reactant conversion developed by Al-Khattaf and de Lasa [31].
According to this model, the deactivation is based on the conversion V dCST
= (k1 CEB − k−1 CST CH )ϕ (2)
of ethylbenzene which is directly related to coke formation. This Wc dt
1394 L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398

Table 5
Coke formation and properties for ethylbenzene dehydrogenation reaction at different reaction conditions.

Catalysts Temp. (◦ C) Reaction time (s) Conversion (%) Coke wt (% C) Coke wt/conversion

Co 550 10 11.38 0.503 0.044


20 15.46 0.488 0.032

Mn 550 10 15.02 0.618 0.041


20 19.70 0.694 0.035

Ni 550 10 13.27 0.582 0.044


20 18.04 0.452 0.025

Comm. 550 10 7.87 0.382 0.049


20 11.15 0.446 0.040

20 100
400 C
18 comm ercial catalyst 90 450 C
Co
Mn 500 C
16 80 styrene
Ni 550 C

Selectivity (%)
14 70
EB conversion (%)

12 60

10 50

8 40

6 30

4 20

2 10 benzene

0 0
0 5 10 15 20 25 0 5 10 15 20 25
Reaction Time (sec) Ethy lbenzene Conversion (%)

Fig. 5. Conversion of ethylbenzene with respect to time at 500 C. Comparison Fig. 7. Selectivity with respect to conversion and temperature for
between commercial and HT catalysts. Mg3 Fe0.25 Mn0.25 Al0.5 .

Rate of formation of benzene, rB

V dCB
= k2 CEB ϕ (3)
Wc dt

where Ci is molar concentration of each of the species in the riser


simulator, V is the volume of the riser (45 cm3 ), Wc is the mass of the
catalyst (0.81 g cat), t is time (s), k is the rate constant (cm3 /(g cat s)),
and ϕ, the catalyst decay function, which accounts for the loss of
catalytic activity as a result of deactivation due to coking. For the
reactant conversion model, the catalyst deactivation function is
given as:

ϕ = exp(−(1 − yEB )) (4)

where  is catalyst deactivation constant.


By definition, the molar concentration Ci of each species can
be written in terms of its mass fraction yi , which are the measur-
able variables from our chromatographic analysis, by the following
relation:
yi Whc
Ci = (5)
MWi V

where Whc is the weight of feedstock injected into the reactor,


Fig. 6. Product selectivity of ethylbenzene dehydrogenation over the different cat- MWi is the molecular weights of the individual species and V is
alysts at ∼6% ethylbenzene conversion.
the volume of the riser simulator.
L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398 1395

14 6
400 C A 400 C B
450 C
12 450 C
500 C 5 500 C
550 C
550 C
10 400 C Pred
400 C Pred
450 C Pred 4
Styrene yield (%)

450 C Pred

Benzene yield (%)


500 C Pred
8 500 C Pred
550 C Pred
550 C Pred
3
6

2
4

1
2

0 0
0 5 10 15 20 25 0 5 10 15 20 25

Ethylbenzene Conversion (%) Ethylbenzene Conversion (%)

Fig. 8. Styrene (A) and benzene (B) yield vs. ethylbenzene conversion at various temperatures Mg3 Fe0.25 Mn0.25 Al0.5 .

To ensure thermodynamic consistency at equilibrium, the dyB Wc


= G4 k2 yEB exp(−(1 − yEB )) (10)
temperature dependent equilibrium constant for the reversible dt V
reaction was found to be:
G1 , G2 , G3 , and G4 are lumped constants given as:
k1
KC = (6) MWEB Whc
k−1 G1 =
MWST MWH V
which has been calculated from literature as [32]:
   MWST
122700 − 126.3T − 2.194 × 10−3 T 2 G2 =
KC = exp − × 105 Pa MWEB
RT
(7)
Whc
G3 =
MWH V
Substituting Eqs. (4)–(6) into Eqs. (1)–(3) and simplifying gives
the following:
MWB
dyEB yST yH
   W G4 =
c MWEB
= − k1 yEB − G1 + k2 yEB exp(−(1 − yEB ))
dt KC V
The activation energy, Ei , for the dehydrogenation reaction is
(8)
related to temperature dependent rate constants ki according to
the Arrhenius equation given below:
dyST yST yH
 W  E 
c i
= k1 G2 yEB − G3 exp(−(1 − yEB )) (9) ki = Ai exp − (11)
dt KC V RT

Table 6
Estimated kinetic parameters for hydrotalcite mixed oxides catalyst based on reactant conversion model.

Parameters Values

k1 k2 

Mg3 Fe0.25 Co0.25 Al0.5


Ei-Co (kJ/mol) 76.37 93.07 30.86
95% CL 7.67 9.99 3.52
k0i a [×103 m3 /(kg of catalyst s)] 1.38 0.64
95% CL 0.30 0.15
Mg3 Fe0.25 Mn0.25 Al0.5
Ei-Mn (kJ/mol) 65.2 111.8 17.18
95% CL 6.41 13.87 2.10
k0i a [×103 m3 /(kg of catalyst s)] 1.24 0.32
95% CL 0.20 0.07
Mg3 Fe0.25 Ni0.25 Al0.5
Ei-Ni (kJ/mol) 87.46 94.74 16.51
95% CL 6.24 9.63 1.9
k0i a [×103 m3 /(kg of catalyst s)] 0.52 0.23
95% CL 0.06 0.03
a
Pre-exponential factor as obtained from Eq. (12); unit for second order (m6 /(kg of catalyst) s).
1396 L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398

Table 7
Correlation matrix for ethylbenzene dehydrogenation over hydrotalcite mixed oxides catalyst.

k1 E1 k2 E2 

Mg3 Fe0.25 Co0.25 Al0.5


k1-Co 1.0000 0.8342 0.9290 0.6689 0.9788
E1-Co 0.8342 1.0000 0.8147 0.6937 0.8645
k2-Co 0.9290 0.8147 1.0000 0.4810 0.9251
E2-Co 0.6689 0.6937 0.4810 1.0000 0.6714
 0.9788 0.8645 0.9251 0.6714 1.0000
Mg3 Fe0.25 Mn0.25 Al0.5
k1-Mn 1.0000 0.7903 0.6898 0.4180 0.9696
E1-Mn 0.7903 1.0000 0.6074 0.4035 0.8374
k2-Mn 0.6898 0.6074 1.0000 −0.2421 0.6962
E2-Mn 0.4180 0.4035 −0.2421 1.0000 0.4154
 0.9696 0.8374 0.6962 0.4154 1.0000
Mg3 Fe0.25 Ni0.25 Al0.5
k1-Ni 1.0000 0.6215 0.6349 0.5320 0.9260
E1-Ni 0.6215 1.0000 0.6500 0.3865 0.8041
k2-Ni 0.6349 0.6500 1.0000 −0.0659 0.7173
E2-Ni 0.5320 0.3865 −0.0659 1.0000 0.5097
 0.9260 0.8041 0.7173 0.5097 1.0000

where Ai is called the pre-exponential factor. Agarwal and Brisk promoted Fe2 O3 . Dehydrogenation of EB over Mg3 Fe0.25 Co0.25 Al0.5
have suggested re-parameterization of equation above during to styrene was conducted using a continuous gas-flow fixed bed
kinetic modeling to help reduce parameter interaction. Therefore, reactor at atmospheric pressure [13]. The reaction was carried
ki constants were re-parameterized by centering the values around out for 3 h of time-on-stream at 550 ◦ C with the catalyst after
k0i which is the value of the rate constant for reaction i at the been pre-treated with a He gas. Activation energy of 29.1 kcal/mol
average temperature T0 of the investigated temperatures [33]: was estimated for Mg3 Fe0.25 Co0.25 Al0.5 using the Arrhenius plot. It
 E 1 1
 can be assumed that the extensive bed mixing and heat distribu-
i
ki = k0i exp − − (12) tion within the fluidized bed enhanced reactant–catalyst contact
R T T0
and uniform product quality, hence reduction of activation energy
Since the experimental runs were done at 400, 450, 500 and towards styrene formation. The active specie is still identified to be
550 ◦ C, To was calculated to be 475 ◦ C. the metastable Fe3+ .
The above reaction model equations were derived based on the The main difference between the three catalysts presented is the
following assumptions: nature of the divalent cation involved in the matrix composition.
Mn2+ has been reported to possess a higher ionic radius than the
1. Catalyst deactivation was assumed to be a function of reactant other Me2+ which may lead to incomplete incorporation of Mn2+
conversion. And, a single deactivation function was defined for into the HT structure [13]. From the estimated activation ener-
all the reactions. gies, the addition of Mn2+ best improved the catalyst activity even
2. The model assumes only catalytic reactions and neglects thermal though both Co2+ and Ni2+ are well incorporated into the HT struc-
conversion. ture. It can thus be suggested that the segregation of Mn2+ serves as
3. The reactor operates under isothermal conditions, justified by
the negligible temperature change observed during the reaction.
20
4. Toluene yield is negligible.
5. The model also neglects the disproportionation reaction. Ethylbenzene Conversion (%)
18
Styrene yield (%)

16 Benzene yield (%)


3.3.2. Determination of model parameters
The kinetic parameters k0i , Ei and  for the reactions were 14
determined by fitting experimental results into the rate Eqs.
(8)–(10) using non-linear regression analysis with the aid of MAT- 12
LAB package. The values of the model parameters along with
Model

their corresponding 95% confidence limits (CLs) are shown in 10


Table 6 while the resulting cross-correlation matrices are also
given in Table 7 for Mg3 Fe0.25 Co0.25 Al0.5 , Mg3 Fe0.25 Mn0.25 Al0.5 and 8
Mg3 Fe0.25 Ni0.25 Al0.5 catalysts respectively. Table 7 shows the corre-
lations between k1 and E1 , E1 and  and k1 and  for all the estimated 6
parameters of dehydrogenation reaction to produce styrene. Sim-
ilarly, correlations between k2 and E2 , E2 and  and k2 and  are 4
shown for parameters of cracking reaction. It can be observed that
the cross-correlation matrices presented in this study revealed 2
low and moderate level of parameter interaction with only a few
exceptions. From the results of the kinetic parameters, the esti- 0
mated apparent energies of 18.3, 15.6, and 20.9 kcal/mol for Co, 0 5 10 15 20
Mn, Ni substituted catalysts, respectively are found to be lower
Experimental
compared to those reported in most literatures, however, the result
obtained for Ni is essentially the same as that reported by Addiego Fig. 9. Overall comparison between the experimental results and model predictions
et al. [18], an apparent activation energy of 21 kcal/mol for 30% K- of all HTs catalysts for Scheme 1.
L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398 1397

additional active sites which increased the basicity of the catalyst, References
responsible for initiating dehydrogenation by H+ abstraction [14].
The activation energies estimated for dealkylation reactions [1] S.J. Liao, T. Chen, C.X. Miao, W.M. Yang, Z.K. Xie, Q.L. Chen, Effect of TiO2 on the

which are 22.3, 26.7, and 22.7 kcal/mol for Mg3 Fe0.25 Co0.25 Al0.5 , structure and catalytic behavior of iron-potassium oxide catalyst for dehydro-
Mg3 Fe0.25 Mn0.25 Al0.5 , and Mg3 Fe0.25 Ni0.25 Al0.5 catalysts respec- genation of ethylbenzene to styrene, Catal. Commun. 9 (2008) 1817–1821.
[2] F. Cavani, F. Trifiò, Alternative processes for the production of styrene, Appl.
tively, are found to be higher than that for dehydrogenation Catal. A 133 (1995) 219–239.
reaction. Similar trend has been reported by Hirano [4,34] both for [3] N. Mimura, I. Takahara, M. Saito, T. Hattori, K. Ohkuma, M. Ando, Dehydrogena-
unpromoted and promoted iron oxide catalysts. This indicates that tion of ethylbenzene over iron oxide-based catalyst in the presence of carbon
dioxide, Catal. Today 45 (1998) 61–64.
high activation energy must be overcome for the cracking reaction [4] T. Hirano, Roles of potassium in potassium-promoted iron-oxide catalyst for
to proceed, which was found to be highest in Mg3 Fe0.25 Mn0.25 Al0.5 . dehydrogenation of ethylbenzene, Appl. Catal. 26 (1986) 65–79.
This depicts that Mn2+ gave the best selectivity towards styrene [5] K. Shibata, T. Kiyoura, Effect of potassium promoter on iron oxide catalysts for
dehydrogenation of ethylbenzene to styrene, Bull. Chem. Soc. Jpn. 42 (1969)
formation which is consistent with the experimental data. 871–874.
The estimated kinetic parameters for the fitted parameters sub- [6] Y. Ohishi, T. Kawabata, T. Shishido, K. Takaki, Q. Zhang, Y. Wang, K. Nomura, K.
stituted into the developed model for the reaction scheme and Takehira, Mg–Fe–Al mixed oxides with mesoporous properties prepared from
hydrotalcite as precursors: catalytic behavior in ethylbenzene dehydrogena-
equations were validated through numerical solution using fourth-
tion, Appl. Catal. A 288 (2005) 220–231.
order-Runge–Kutta routine. The numerical results were compared [7] B.D. Herzog, H.F. Rase, In situ catalyst reactivation: used ethylbenzene dehy-
with experimental data as shown in Fig. 4(A–C). The figure shows drogenation catalyst with agglomerated potassium promoter, Ind. Eng. Chem.
the model predictions have a very good match with the experi- Prod. Res. Dev. 23 (1984) 187–196.
[8] J.C.A.A. Roelofs, D.J. Lensveld, A.J. van Dillen, K.P. de Jong, On the structure of
mental values. Further comparisons between model predictions activated hydrotalcites as solid base catalysts for liquid-phase aldol condensa-
and experimental data are presented in Fig. 8(A and B). This tion, J. Catal. 203 (2001) 184–191.
demonstrates that the proposed kinetic model fits well into our [9] P. Kuśtrowski, D. Sułkowska, L. Chmielarz, A. Rafalska-Lasocha, B. Dudek,
R. Dziembaj, Influence of thermal treatment conditions on the activity of
experimental observations. The reconciliation plot (Fig. 9) shows hydrotalcite-derived Mg–Al oxides in the aldol condensation of acetone, Micro-
the overall agreement of experimental data and model predictions. por. Mesopor. Mater. 78 (2005) 11–22.
The two agrees very well as suggested by the value of regression [10] F. Cavani, F. Trifiro, A. Vaccari, Hydrotalcite-type anionic clays: preparation,
properties and applications, Catal. Today 11 (1991) 173–301.
coefficient, r2 (0.99). [11] Y. Liu, E. Lotero, J.G. Goodwin, X. Mo, Transesterification of poultry fat with
methanol using Mg–Al hydrotalcite derived catalysts, Appl. Catal. A 331 (2007)
138–148.
4. Conclusion [12] D.P. Debecker, E.M. Gaigneaux, G. Busca, Exploring, tuning, and exploiting the
basicity of hydrotalcites for applications in heterogeneous catalysis, Chem. Eur.
The kinetics of ethylbenzene has been carried out over J. 15 (2009) 3920–3935.
[13] R.J. Balasamy, A. Khurshid, A.A.S. Al-Ali, L.A. Atanda, K. Sagata, H. Yahiro, K.
Mg3 Fe0.25 Me0.25 Al0.5 (Me = Co, Mn and Ni) quaternary mixed oxide Nomura, T. Sano, K. Takehira, S.S. Al-Khattaf, Ethylbenzene dehydrogenation
catalysts prepared from hydrotalcites as precursors. The following over binary FeOx –MeOy /Mg(Al)O catalysts derived from hydrotalcites, Appl.
conclusions were drawn from the study: Catal. A 390 (2010) 225–234.
[14] R.J. Balasamy, A. Khurshid, A.A.S. Al-Ali, L.A. Atanda, K. Sagata, H. Yahiro, K.
Nomura, T. Sano, K. Takehira, S.S. Al-Khattaf, Ethylbenzene dehydrogenation
(1) The EB conversion increases significantly with the increase of over FeOx /(Mg,Zn)(Al)O catalysts derived from hydrotalcites: role of MgO as
reaction temperature and time, but the selectivity to styrene the catalyst support, Appl. Catal. A 398 (2011) 113–122.
[15] S. Carrà, L. Forni, Kinetics of catalytic dehydrogenation of ethylbenzene to
decreased. styrene, Ind. Eng. Chem. Process Des. Dev. 4 (1965) 281–285.
(2) The amount of benzene and toluene in the reaction product [16] K. Coulter, D.W. Goodman, R.G. Moore, Kinetics of the dehydrogenation of
increased as the temperature was raised, showing that cracking ethylbenzene to styrene over unpromoted and K promoted model iron-oxide
catalysts, Catal. Lett. 31 (1995) 1–8.
reaction is favored at higher temperature. However, the yield [17] O. Shekhah, W. Ranke, R. Schlögl, Styrene synthesis: in situ characterization and
of toluene was not appreciable and hence was neglected. reactivity studies of unpromoted and potassium-promoted iron oxide model
(3) Incorporation of Mn2+ was the most effective giving superior catalysts, J. Catal. 225 (2004) 56–68.
[18] W.P. Addiego, C.A. Estrada, D.W. Goodman, M.P. Rosynek, R.G. Windham, An
activity and selectivity towards styrene. A maximum EB con- infrared study of the dehydrogenation of ethylbenzene to styrene over iron-
version of 19.7% at 550 ◦ C, and styrene selectivity of 92.7% at based catalysts, J. Catal. 146 (1994) 407–414.
400 ◦ C were attained. [19] A. Miyakoshi, A. Ueno, M. Ichikawa, Mn-substituted Fe–K mixed oxide catalysts
for dehydrogenation of ethylbenzene towards styrene, Appl. Catal. A 216 (2001)
(4) Kinetic parameters for the ethylbenzene dehydrogenation
137–146.
showed that apparent energies for styrene formation were [20] W.J. Lee, G.F. Froment, Ethylbenzene dehydrogenation to styrene: kinetic mod-
found to decrease as follows: E1-Ni > E1-Co > E1-Mn . Also, activa- eling and reactor simulation, Ind. Eng. Chem. Res. 47 (2008) 9183–9194.
[21] B.K. Abdalla, S.S.E.H. Elnashaie, Fluidized bed reactors without and with selec-
tion energies for the cracking reaction were found to decrease
tive membranes for the catalytic dehydrogenation of ethylbenzene to styrene,
as follows: E2-Mn > E2-Ni > E2-Co . J. Membr. Sci. 101 (1995) 31–42.
(5) Activation energy for cracking was found to be higher than [22] S. Miyata, A. Okada, Synthesis of hydrotalcite-like compounds and their physic-
dehydrogenation to styrene for all the HT catalyst, i.e., ochemical properties—the systems Mg2+ –Al3+ –SO4 2− and Mg2+ –Al3+ –CrO4 2− ,
Clays Clay Miner. 25 (1977) 14–18.
E2-Me > E1-Me (Me = Co, Mn and Ni). This trend agrees with lit- [23] H.I. de Lasa, Riser simulator for catalytic cracking studies. U.S. Patent 5,102,628
erature. (1992).
[24] D.W. Kraemer, H.I. de Lasa, Catalytic cracking of hydrocarbons in a riser simu-
lator, Ind. Eng. Chem. Res. 27 (1988) 2002–2008.
Acknowledgments [25] R.D. Shannon, Revised effective ionic radii and systematic studies of interatomic
distances in halides and chalcogenides, Acta Crystallogr. A 32 (1976) 751–767.
[26] X. Ge, M. Li, J. Shen, The reduction of Mg–Fe–O and Mg–Fe–Al–O complex
This publication was based on work supported in part by Award oxides studied by temperature-programmed reduction combined with in situ
No. K-C1-019-12 made by King Abdullah University of Science Mössbauer spectroscopy, J. Solid State Chem. 161 (2001) 38–44.
and Technology (KAUST). In addition, the authors express their [27] D. Li, I. Atake, T. Shishido, Y. Oumi, T. Sano, K. Takehira, Self-regenerative activity
of Ni/Mg(Al)O catalysts with trace Ru during daily start-up and shut-down
appreciation for the financial support from the Centre for Research operation of CH4 steam reforming, J. Catal. 250 (2007) 299–312.
Excellence in Petroleum Refining and Petrochemicals at King Fahd [28] A. Chen, H. Xu, Y. Yue, W. Hua, W. Shen, Z. Gao, Hydrogenation of methyl
University of Petroleum & Minerals. The authors also acknowl- benzoate to benzaldehyde over manganese oxide catalysts prepared from
Mg/Mn/Al hydrotalcite-like compounds, Appl. Catal. A 274 (2004) 101–109.
edge the contribution of Prof. Emeritus K. Takehira (Hiroshima [29] T. Hirano, Dehydrogenation of ethylbenzene on potassium-promoted iron-
University, Higashi-Hiroshima, Japan). Mr. Mariano Gica is also oxide catalysts containing various transitional-metal oxides, Bull. Chem. Soc.
acknowledged for his help during the experimental work. Jpn. 59 (1986) 1653–1655.
1398 L. Atanda et al. / Chemical Engineering Journal 171 (2011) 1387–1398

[30] E.H. Lee, Iron oxide catalysts for dehydrogenation of ethylbenzene in the pres- [33] A.K. Agarwal, M.L. Brisk, Sequential experimental design for precise parameter
ence of steam, Catal. Rev. Eng. Sci. 8 (1973) 285–305. estimation. 1. Use of reparameterization, Ind. Eng. Chem. Process Des. Dev. 24
[31] S. Al-Khattaf, H. de Lasa, Catalytic cracking of cumene in a riser simulator: a (1985) 203–207.
catalyst activity decay model, Ind. Eng. Chem. Res. 40 (2001) 5398–5404. [34] T. Hirano, Dehydrogenation of ethylbenzene over potassium-promoted iron
[32] R. Dittmeyer, V. Höllein, P. Quicker, G. Emig, G. Hausinger, F. Schmidt, oxide containing cerium and molybdenum oxides, Appl. Catal. 28 (1986)
Factors controlling the performance of catalytic dehydrogenation of ethylben- 119–132.
zene in palladium composite membrane reactors, Chem. Eng. Sci. 54 (1999)
1431–1439.

You might also like