Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/284009491

Dehydrogenation–hydrogenation of methylcyclohexane-
toluene system on 1.0 wt% Pt/zeolite beta catalyst

Article  in  Progress in Reaction Kinetics and Mechanism · November 2015


DOI: 10.3184/146867815X14413752286029

CITATIONS READS

3 792

3 authors:

Muhammad Rashid Usman Faisal Alotaibi


University of the Punjab Saudi Aramco
30 PUBLICATIONS   451 CITATIONS    6 PUBLICATIONS   19 CITATIONS   

SEE PROFILE SEE PROFILE

Rabya Aslam
University of the Punjab
27 PUBLICATIONS   103 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Liquid-Liquid Extraction View project

Not project View project

All content following this page was uploaded by Rabya Aslam on 21 December 2015.

The user has requested enhancement of the downloaded file.


Dehydrogenation of methylcyclohexane–toluene

Progress in Reaction Kinetics and Mechanism, 2015, 40(4), 353–366


doi:10.3184/146867815X14413752286029 Paper: 1400337

Dehydrogenation–hydrogenation of
methylcyclohexane-toluene system on 1.0 wt%
Pt/zeolite beta catalyst
Muhammad R. Usmana,b*, Faisal M. Alotaibic and Rabya Aslamb
a School of Chemical Engineering and Analytical Science, University of Manchester,
Manchester, UK
bInstitute of Chemical Engineering and Technology, University of the Punjab, Lahore, Pakistan
cKing Abdulazeez City for Science & Technology, Saudi Arabia

E-mail: mrusman.icet@pu.edu.pk

ABSTRACT
The dehydrogenation of methylcyclohexane is a reversible reaction where
methylcyclohexane is dehydrogenated to toluene and hydrogen in the forward reaction
while the hydrogenation of toluene occurs is the reverse reaction. In the present work,
both the forward and the reverse reactions were studied over the same catalyst: 1.0
wt% Pt/zeolite beta. The catalyst was prepared using the ion exchange method and
it was characterised using ICP-OES, XRD, SEM, XPS, N2-BET, and NH3-TPD.
The catalytic reactions were studied in a 1.02 cm internal diameter fixed bed tubular
reactor under integral conditions and the effect of reaction operating conditions were
studied on both the forward and reverse reactions. The kinetic modelling of the
overall reactions, as well as of the by-products formed in the forward reaction, was
carried out using simple power law models.

KEYWORDS: dehydrogenation, hydrogenation, zeolite beta, organic


hydride, power law kinetics

1. INTRODUCTION
The dehydrogenation of methylcyclohexane (MCH) is a reversible reaction in which
MCH is dehydrogenated to toluene and hydrogen in the forward reaction while
toluene is hydrogenated in the reverse reaction. The dehydrogenation–hydrogenation
of MCH–toluene system is an important reaction system in the reforming of naphtha
and in hydrogen storage applications as in the methylcyclohexane–toluene-hydrogen
(MTH) system [1–6]. The dehydrogenation of MCH has already been studied
mostly using Pt/Al2O3 catalysts. However, only a few studies have been carried out
on Pt/zeolite catalysts. Corma et al. [7] and García de la Banda et al. [8] studied
the dehydrogenation of MCH over Pt/NaY zeolite catalyst. Coughlan and Keane
[9] studied the dehydrogenation of MCH over Ni-supported zeolite-Y catalysts.
The obvious reason for using alumina as a support is its lower acidity compared
353
354 Muhammad R. Usman, Faisal M. Alotaibi and Rabya Aslam

to zeolites and its large surface area (for γ and η phases) though the latter is not
comparable to zeolites. It is true that zeolites are well-known for supporting cracking
and isomerisation reactions, a zeolite with low acidity and relatively large surface
area such as zeolite beta is worth studying for the dehydrogenation of MCH. In the
present study, 1.0 wt% Pt/zeolite beta was prepared in-house and well characterised.
The dehydrogenation of MCH was carried out for a range of operating conditions
and both the activity and selectivity of the catalyst were studied. Moreover, a few
experiments were also carried out over the same catalyst to study the activity and
selectivity of the catalyst for the reverse reaction, the hydrogenation of toluene.
Kinetic modelling of the overall reaction (including by-products) for each of the
forward (dehydrogenation) and reverse (hydrogenation) reactions was carried out.
Separately, kinetic modelling for the yields of the by-products formed during only
the dehydrogenation reaction was also carried out. To our knowledge, study of the
dehydrogenation-hydrogenation of the MCH–toluene system over Pt/zeolite beta
catalyst has never been reported in the open literature.

2. EXPERIMENTAL
The catalyst, 1.0 wt% Pt over zeolite beta, was prepared by the ion-exchange
technique. The zeolite beta support was provided by Zeolyst International and was
used without any modification. For the catalyst preparation, each gram of the support
was mixed with 0.0175 g of tetraammine platinum(II) chloride [Pt(NH3)4Cl2.
H2O] (Aldrich, 99.5% purity) and dissolved in the required amount of deionised
water. The desired pH was achieved by adding ammonia solution (32%, Aldrich).
The mixture was stirred for 24 h, then washed, centrifuged, and dried in an oven
at 110 °C. The catalyst was subjected to various characterisation techniques such
as inductively-coupled plasma-optical emission spectroscopy (ICP-OES), X-ray
photoelectron spectroscopy (XPS), X-ray diffraction (XRD), ammonia–temperature
programmed desorption (NH3-TPD), scanning electron microscopy (SEM), and
Brunaeur–Emmett–Teller (BET) surface area analysis by nitrogen adsorption. The
dehydrogenation and hydrogenation reactions of MCH and toluene, respectively,
were studied in a 10.2 mm fixed bed reactor under integral conditions. Experiments
were performed under varying operating conditions of space time, reactor wall
temperature, operating pressure, and feed compositions. The products of the catalytic
reactions were analysed by a Varian gas chromatograph using a 100 m fused silica
capillary column (CP-Sil PONA CB: 100% dimethylpolysiloxane).

3. CATALYST CHARACTERISATION
Table 1 gives a summary of the results obtained for the characterisation of the
catalyst, while Figures 1 and 2 show the SEM image and XRD patterns of the
catalyst, respectively. The XRD patterns shown in Figure 2 are comparable to those
found in the literature [10] and confirm the presence of the zeolite beta phase. The
www.prkm.co.uk
Dehydrogenation of methylcyclohexane–toluene 355
Table 1 Results of characterisation of the catalyst
Before ion exchange After ion exchange with Pt
Elemental analysis
(bulk)
Si/Al mole
164.32
Si/Al mole
164.32
Pt 1.15
ratio fraction (wt%) (ICP)
XPS
Si = 25.57, Al = 1.17, Si/Al = 21.85, Pt = 0.05
Atomic %conc.
XRD Crystallinity = 96%
Si/Al mole ratio for the framework = 12.07
NMR
Lewis acid sites are not present
NH3-TPD 0.08 (±0.01) mmol g–1
Surface area = 509.81 m2 g-1, Pore volume = 0.38 cm3 g-1
N2-BET
Pore size = 33.30 Å

Figure 1 SEM image of the catalyst.

Figure 2 XRD patterns for the


catalyst.

SEM image of the catalyst suggests the occurrence of uniform and virtually spherical
crystallites with size ranging between 0.5 and 2 µm.

4. RESULTS AND DISCUSSION


As the dehydrogenation reaction of MCH to produce toluene is endothermic, while
the hydrogenation reaction of toluene to produce methylcyclohexane is exothermic,
the temperature of the reactor bed decreased during the course of the dehydrogenation
reaction and the temperature of the reactor bed increased for the hydrogenation
www.prkm.co.uk
356 Muhammad R. Usman, Faisal M. Alotaibi and Rabya Aslam

Figure 3 Axial temperature profiles: (a) for the dehydrogenation of MCH at various temperatures,
(b) for the hydrogenation of toluene at various operating pressures.

reaction. The average bed temperatures were therefore not the same as the reactor
wall temperatures. Figure 3 shows a few temperature profiles measured at the centre
line along the axial position of the bed using a sliding wire thermocouple while Tables
2 and 3 give the average temperature for its corresponding wall temperature under
the given conditions of experimentation. An average temperature was measured
using the following equation:

1 i = N  Ti + Tw 
T= ∑
N i =1  2 
 (1)

Both the forward (dehydrogenation of MCH) and reverse (hydrogenation of toluene)


reactions were found to be not clean under the conditions studied and in each case the
reaction products contained a large proportion of by-products. Generally, the isomers
of dimethylcyclopentane (DMCP), ethylcyclopentane (ECP), and 3-methylhexane
(MXN) were among the important by-products obtained in each of the reactions.
Usman et al. [11] and Alhumaidan et al. [12] have also found these by-products in
their reaction products while studying the dehydrogenation reaction over a Pt/Al2O3
catalyst. Corma et al. [13] while working on HNaY zeolite (without metal loading)
have also found dimethylcyclopentanes in their products with MCH as the feed. The
presence of these by-products in appreciable amounts suggested the occurrence of
isomerisation reactions caused by the acidic character of the zeolite beta support.
Figure 4 (a–f) shows the effect of various operating conditions on the conversion of
MCH and the yields of by-products formed in the forward reaction. It is observed
that an increase in the reaction pressure generally has a moderate effect on the overall
conversion of MCH to products. The MCH conversion slightly decreased initially
and then remained virtually constant. The yield of toluene (main reaction product)
also decreased with the reactor pressure, however, the change was much higher than
the change in MCH conversion. This is because the ratio of the rate of formation of
by-products to the rate of formation of toluene increases with an increase in pressure.
In contrast to the yield of toluene, the yields of by-products generally increased with
an increase in pressure and the effect was much more pronounced for the combined
www.prkm.co.uk
Table 2 Experimental data for the dehydrogenation reaction
Tw T p W/FA0 ×10 –5
F H20 /FA0 yA0 yH20 %X Ae %X A %Y T %YA %YE %YD %YM
(°C) (°C) (bar) (s g-cat/mol A)
359.9 337.5 1.013 4.214 0.1914 0.8067 0.6219 91.15 77.21 59.81 22.79 3.466 12.25 0.893
278.8 269.7 1.013 4.214 0.1914 0.8067 0.6219 77.52 31.12 22.99 68.88 2.493 4.743 0.306
299.0 286.6 1.013 4.214 0.1914 0.8067 0.6219 82.28 40.52 29.94 59.48 3.422 6.480 0.396
319.6 303.6 1.013 4.214 0.1914 0.8067 0.6219 86.03 53.87 40.38 46.13 3.525 9.046 0.532
340.1 319.9 1.013 4.214 0.1914 0.8067 0.6219 88.82 68.42 53.28 31.58 3.551 10.41 0.679
359.9 332.3 1.013 4.214 0.1914 0.8067 0.3109 90.53 51.74 38.08 48.26 3.695 8.907 0.532
359.9 336.5 1.013 8.428 0.1059 0.8930 0.6219 90.77 82.81 70.90 17.19 2.881 7.809 0.623
359.9 340.2 1.013 1.054 0.4846 0.5105 0.6219 92.33 72.83 50.14 27.18 3.939 14.87 1.049
359.9 338.9 2.0 4.214 0.1914 0.8067 0.6219 89.33 73.36 53.88 26.64 3.488 13.65 0.922
359.9 345.6 6.0 4.214 0.1914 0.8067 0.6219 86.46 73.80 38.19 26.20 5.608 26.03 3.048
359.9 348.9 8.0 4.214 0.1914 0.8067 0.6219 85.87 72.18 28.03 27.82 6.830 30.98 5.302
380.3 357.7 1.013 4.214 0.1914 0.8067 0.6219 93.16 83.32 64.62 16.68 3.407 12.28 1.186

www.prkm.co.uk
Table 3 Experimental data for the hydrogenation reaction
Tw T p W/FT0 ×10 –5
F H20 /FT0 yT0 yH20 %XTe %XT %Y T %YA %YE %YD %YM
(°C) (°C) (bar) (s g-cat/mol T)
219.8 262.6 8 4.344 0.1871 0.8129 0.6410 99.77 0.6310 99.37 53.97 7.909 29.40 5.412
219.8 257.2 4 4.344 0.1871 0.8129 0.6410 98.75 18.58 81.42 58.44 7.131 13.58 2.155
Dehydrogenation of methylcyclohexane–toluene

219.8 259.2 4 8.688 0.1032 0.8968 0.6410 99.62 10.03 89.97 64.59 8.494 14.52 2.255
219.8 233.1 1.013 4.344 0.1871 0.8129 0.6410 92.08 72.65 27.35 25.40 1.122 0.722 0.1110
240.5 251.0 1.013 4.344 0.1871 0.8129 0.6410 75.86 78.15 21.85 19.63 1.192 0.900 0.1290
281.3 285.3 1.013 4.344 0.1871 0.8129 0.6410 23.04 90.93 9.074 6.274 0.9330 1.209 0.1600
357
358 Muhammad R. Usman, Faisal M. Alotaibi and Rabya Aslam

Figure 4 Effect of change in variables on the dehydrogenation of MCH: (a) Effect of pressure on
the overall conversion; (b) effect of pressure on by-products formation; (c) effect of temperature
on the overall conversion; (d) effect of temperature on by-products formation; (e) effect of
hydrogen concentration in feed on the overall conversion; (f) effect of hydrogen concentration
in feed on by-products formation.

yield of isomers of dimethylcyclopentane compared to the yields of the other by-


products, and at 8 bar the yield of dimethylcyclopentanes actually exceeds that of
toluene.
The MCH conversion was found to be a strong function of temperature. It increased
with increase in temperature. Again, the yields of the by-products formed were
increased, however, in contrast to the effect of pressure, the yield of toluene increased
with increase in temperature while the yield of dimethylcyclopentanes did not
www.prkm.co.uk
Dehydrogenation of methylcyclohexane–toluene 359

Figure 5 Effect of change in variables on the hydrogenation of toluene. (a) Effect of pressure on
overall conversion; (b) effect of pressure on by-products formation; (c) effect of temperature on
overall conversion; (d) effect of temperature on by-products formation.

increase at the same rate. Increasing the H2/MCH molar ratio increased both the
MCH conversion and the toluene yield, however, the yields of by-products decreased
with an increase in the H2/MCH molar ratio. This observation may suggest that
hydrogen addition in the feed may cause some kind of hydrogenation of the coke
precursors and stabilise the activity of the Pt metal that performs the hydrogenation–
dehydrogenation reaction and/or it is helping in the rapid adsorption of MCH or
in the rapid desorption of some strongly adsorbed intermediates that lead to the
formation of toluene and the desorption of toluene itself. This type of observation has
also been reported by other researchers in the field. Rohrer and Sinfelt [14], Jossens
and Petersen [15], Usman et al. [16], and Usman et al. [17] have also observed, at low
pressures, an increase in the rate of the dehydrogenation reaction when hydrogen is
present in the feed.
Figure 5 (a–d) shows the effect of pressure and temperature on the conversion of
toluene to its corresponding products and the yields of by-products formed during the
hydrogenation of toluene. Similar by-products were obtained for the reverse reaction
and again an appreciable amount of by-products was observed. It was observed that
the reactor pressure has a profound effect on the rate of consumption of toluene and
an increase in pressure greatly increases the toluene conversion and at 8 bar pressure,
www.prkm.co.uk
360 Muhammad R. Usman, Faisal M. Alotaibi and Rabya Aslam

a reactor wall temperature of 220 °C, and a H2/TOL molar ratio of 4.34, the toluene
conversion virtually reaches the equilibrium value. This suggests that under the
conditions described, any pressure greater than 8 bar would not help in the conversion
of toluene to the reaction products. The yields of by-products also increased with an
increase in pressure where the combined yield of isomers of dimethylcyclopentanes
increased the most significantly. On increasing temperature (at low pressure), for a
given temperature range, the toluene conversion decreased while the yields of the by-
products were found generally to be not affected by temperature.
The observations in Figures 4 and 5 may not be generalised, as under any other
reaction conditions, the effects on the conversions and yields of the principal products
and by-products may be different.

5. KINETIC MODELLING
The laboratory experimental reactor was assumed to behave as a one-dimensional
plug flow reactor and the following differential equation was used to describe the
reactor performance:
dX i
= (−ri ) (2)
d (W / Fi 0 )

Based on the formation of major by-products, a reaction sequence was written for
each of the dehydrogenation and hydrogenation reactions as shown in Table 4. For
each of the reaction schemes an overall reaction was written, based on which a rate
equation was defined and solved along with Eqn (2) to fit against the experimental
reaction data. A Fortran routine was used to solve the differential equation [(Eqn (2)].
Table 4 Reaction sequences for the dehydrogenation and hydrogenation reactions
For dehydrogenation
MCH ↔ TOL + 3H2
MCH ↔ DMCP1
MCH ↔ DMCP2
MCH ↔ DMCP3
MCH ↔ ECP
MCH + H2 ↔ MXN
1 1 1 1 1 1 1
Overall: A ↔ TOL + DMCP1 + DMCP 2 + DMCP3 + ECP + MXN + H 2
6 6 6 6 6 6 3
For hydrogenation
TOL + 3H2 ↔ MCH
TOL + 3H2 ↔ DMCP1
TOL + 3H2 ↔ DMCP2
TOL + 3H2 ↔ DMCP3
TOL + 3H2 ↔ ECP
TOL + 3H2 ↔ MXN
19 1 1 1 1 1 1
Overall: TOL + H 2 ↔ MCH + DMCP1 + DMCP 2 + DMCP3 + ECP + MXN
6 6 6 6 6 6 6

MCH, methylcyclohexane; TOL, toluene; DMCP1, 1-1, dimethylcyclopentane; DMCP2, 1-2,


dimethylcyclopentanes; DMCP3, 1-3, dimethylcyclopentane; ECP, ethylcyclopentane; MXN,
3-methylhexane.
www.prkm.co.uk
Dehydrogenation of methylcyclohexane–toluene 361
Simple power law rate models were used and the experimental data were subjected
to regression and the SSE (the sum of squares of the errors) was used as the objective
function to be minimised.
Based on the overall balanced equation for the dehydrogenation reaction, the
following rate equation was written:

( )1/ 6 nA
 p ⋅ p ⋅ p ⋅ p ⋅ p ⋅ p ⋅ p2 
(−rA ) = k A ⋅  p A − T D1 D 2 D 3 E M H 2  (3)
 KA 
 
while Eqn (4) was developed for the overall hydrogenation reaction:
nT

(−rT ) = kT ⋅  pT ⋅ p19 /6 ( p A ⋅ pD1 ⋅ pD 2 ⋅ pD3 ⋅ pE ⋅ pM )1 / 6 
H2 −  (4)
 KT 
In each case, the rate constant k was given the Arrhenius temperature dependency
as shown below:

  T 
ki = kri ⋅ exp Bi 1 − ri   (5)
  T 
where,
Ei (6)
Bi =
RTri
where, Tr is the reference temperature which lies at the centre of all the reaction
temperatures in a set of experimental data. The partial pressure of each component
was described by the equations given in Table 5. The equilibrium constants for the
dehydrogenation and hydrogenation reactions were calculated from the Gibbs free
energies of formation of the reaction components involved. The values of the Gibbs
free energy of formation were taken from Stull et al. [18]. Eqns (7) and (8) describe
the equilibrium constants for the dehydrogenation and hydrogenation reactions,
respectively.
1000
ln K A = −5.3515 × + 9.2364 (7)
T

1000
ln KT = 24.72 × − 46.49 (8)
T
The results of the kinetic treatment are shown in Table 6 and Figure 6 shows the
scatter diagram between the observed conversions and model conversions for both
the dehydrogenation and hydrogenation reactions. As the hydrogenation reaction is an
exothermic reaction so the net rate of the forward reaction decreases with temperature
(the dehydrogenation reaction becomes more pronounced at high temperature). The
negative value of the activation energy, though not possible, is only to show that the
www.prkm.co.uk
362 Muhammad R. Usman, Faisal M. Alotaibi and Rabya Aslam
Table 5 Mole fraction of the components involved at any conversion Xi for the overall reactions
defined in Table 4
Mole fraction at any conversion Xi
Component Representation For dehydrogenation For hydrogenation
y A0 − X A y A0 y A0 + 0.16667 X T yT 0
MCH A yA = yA =
1 + 0.33333 X A y A0 1 − 3.16667 ⋅ X T yT 0
yT 0 + 0.16667 X A y A0 yT 0 − X T yT 0
TOL T yT = yT =
1 + 0.33333 X A y A0 1 − 3.16667 ⋅ X T yT 0
1 + 0.16667 X A y A 0
y D10 y D10 + 0.16667 X T yT 0
DMCP1 D1 y D1 = y D1 =
1 + 0.33333 X A y A0 1 − 3.16667 ⋅ X T yT 0
y D20 + 0.16667 X A y A0 y + 0.16667 X T yT 0
DMCP2 D2 yD 2 = y D 2 = D20
1 + 0.33333 X A y A0 1 − 3.16667 ⋅ X T yT 0
y D30 + 0.16667 X A y A0 y D30 + 0.16667 X T yT 0
DMCP3 D3 yD3 = yD3 =
1 + 0.33333 X A y A0 1 − 3.16667 ⋅ X T yT 0
y E 0 + 0.16667 X A y A0 y E 0 + 0.16667 X T yT 0
ECP E yE = yE =
1 + 0.33333 X A y A0 1 − 3.16667 ⋅ X T yT 0
yM 0 + 0.16667 X A y A0 yM 0 + 0.16667 X T yT 0
MXN M yM = yM =
1 + 0.33333 X A y A0 1 − 3.16667 ⋅ X T yT 0
y H 20 + 0.33333 X A y A0 y − 3.16667 ⋅ X T yT 0
H2 H2 yH 2 = y H 2 = H 20
1 + 0.33333 X A y A0 1 − 3.16667 ⋅ X T yT 0

Table 6 Results of the kinetic modelling based on the overall reactions defined in Table 4
Dehydrogenation Hydrogenation
Parameter Value Units Parameter Value Units
N 12 – N 6 –
krA 1.143×10 –5 mol s–1 g-cat–1 bar–nA krT 0.5252×10 –5 mol s–1 g-cat–1 bar–nT
BA 6.424 – BT –29.24 –
(E A) 6.0 kJ mol–1 (ET) –129.14 kJ mol–1
nA 0.0740 – nT 0.3088 –
Statistics
SSE 0.0555 SSE 0.0668
F-value 20.79 F-value 16.16
Tr = 600.79 K for the overall dehydrogenation reaction and Tr = 531.21 K for the overall
hydrogenation reaction.

rate constant is decreasing with temperature. A simple power law model may not
be suitable for the data and therefore a detailed mechanistic study is required and a
more rigorous kinetic model is needed to evaluate the proper activation energy for
the reaction which is beyond the scope of this study.
The data for the yields of toluene and the major by-products of the dehydrogenation
reaction were modelled separately. As 3-MXN was obtained generally in low
concentrations, only the combined yield of the isomers of dimethylcyclopentane and
www.prkm.co.uk
Dehydrogenation of methylcyclohexane–toluene 363

Figure 6 Scatter diagram between overall observed (experimental) and model (calculated)
conversions: (a) for the dehydrogenation of MCH; (b) for the hydrogenation of toluene. Any
model value greater than 1.0 was constrained to 1. There was only one such value for the
modelling of hydrogenation of toluene.

yield of ethylcyclopentane were kinetically modelled. The main reaction and the side
reactions were considered to be occurring in parallel (Table 4) and the following
sequence of reactions as shown in Eqns (9)–(11) was proposed.
dYT
= (rT ) (9)
d (W / FA0 )

dYD
= (rD ) (10)
d (W / FA0 )

dYE
= (rE ) (11)
d (W / FA0 )
The following rate equations, when used with Eqns (9), (10) and (11), respectively,
were found the best fitting rate models.
nT ′
 p ⋅ p3 
(rT ) = kT ′ ⋅  p A − T H 2  (12)
 KT ′ 

pD
(rD ) = k D ⋅ ( p A − ) (13)
KD

(rE ) = k E ⋅ p A (14)
where, kT´, kD, and kE were subjected to Arrhenius temperature dependency as
described in Eqns (5) and (6). The equilibrium constant for Eqn (12) was as given
by Schildhaur et al. [19] and the equilibrium constant for Eqn (13) was calculated
from the Gibbs free energies of formation of the components involved [18]. The
equilibrium constant for dimethylcyclopentane formation from methylcyclohexane
is given by Eqn (15). The equilibrium constant was calculated based on the free
www.prkm.co.uk
364 Muhammad R. Usman, Faisal M. Alotaibi and Rabya Aslam
Table 7 Mole fractions in the reaction mixture for the modelling of the yields of toluene and
selected by-products
Components Symbol Mole fraction
y A0 ⋅ (1 − (YT + YD + YE )
Methylcyclohexane A
1 + 3 ⋅ YT ⋅ y A0
yT 0 + YT ⋅ y A0
Toluene T
1 + 3 ⋅ YT ⋅ y A0
yH 20 + 3 ⋅ YT ⋅ y A0
Hydrogen H2
1 + 3 ⋅ YT ⋅ y A0
yD 0 + YD ⋅ y A0
Dimethylcylcopentanes D
1 + 3 ⋅ YT ⋅ y A0
yE 0 + YE ⋅ y A0
Ethylcyclopentane E
1 + 3 ⋅ YT ⋅ y A0

Figure 7 Scatter diagrams for the yields of toluene and selected by-products obtained in
the dehydrogenation reaction: (a) yield of toluene; (b) yield of dimethylcyclopentanes;
(c) ethylcyclopentane.

energies of cis-1-1, dimethylcyclopentane. The partial pressures in the reaction


mixture that are to be used with Eqs (12)–(14) are given in Table 7.

1000
ln K D = −2.0723 × + 2.4175 (15)
T
The results of regression for the modelling of the by-products are shown in Table 8
and the scatter diagrams relating the observed and model yields of the toluene and
by-products are shown in Figure 7 (a–c).
Table 8 Results of the kinetic modelling for the yields of toluene and selected by-products
Parameter Value Units Parameter Value Units
k rT´ 0.4135×10 –5 mol s –1 g-cat–1 bar ­–nT BT´ 1.299 ―
k rD 6.816×10 –5 mol s –1 g-cat–1 bar ­–1 BD 17.64 ―
k rE 0.1536×10 –5 mol s –1 g-cat–1 bar ­–1 BE 22.22 ―
nT´ ‒0.0965 ―
Statistics
N 36 SSE 0.1252 F-value 52.81
Tr = 600.79 K.

www.prkm.co.uk
Dehydrogenation of methylcyclohexane–toluene 365
6. CONCLUSION
For the conditions utilised in the present work, the variation in reactor temperature
strongly affected the overall conversion of MCH to products. The effect of pressure
and hydrogen was generally less significant. On the other hand, the overall
conversion of tolune to reaction products was strongly affected by change in the
reaction pressure. Both the dehydrogenation of MCH as well as the hydrogenation
of toluene over the 1.0 wt% Pt/zeolite beta were found to be not clean and in each
case ethylcyclopentane and isomers of dimethylcyclopentane were found to be the
principal by-products. In both cases, the yields of by-products were found to be an
important function of reaction pressure. For the dehydrogenation reaction, at 1.013
bar pressure, with the addition of hydrogen in the feed, the overall conversion was
increased while the yields of by-products decreased with an increase in the hydrogen
concentration in the feed. The power law kinetics for the dehydrogenation of MCH
suggested zero-order kinetics with respect to methylcyclohexane, however, as the
hydrogenation reaction is exothermic, a negative value of the activation energy,
though not possible, showed that the rate constant decreased with the temperature. A
simple power law model therefore may not be suitable for the hydrogenation kinetics
and therefore a detailed mechanistic study is required and a more rigorous kinetic
model is needed to evaluate the proper activation energy for the reaction.

7. NOMENCLATURE
Dimensionless activation energy or Arrhenius number: defined for Eqn (3) = BA; defined for Eqn
(13) = BD; defined for Eqn (14) = BE; for an ith reaction = Bi; defined for Eqn (4) = BT; defined for
Eqn (12) = BT´.
Activation energy (kJ mol-1): defined for Eqn (3) = EA; for an ith reaction = Ei; for Eqn (4) = ET.
Initial molar flowrate (mol s-1): of methylcyclohexane = FA0; hydrogen = FH2O; ith component =
Fi0.
Rate constant (mol kg-1 s-1): for dehydrogenation of methylcyclohexane as defined in Eqn (3) =
kA; of an ith reaction = ki; for Eqn (3) at reference temperature = krA; for an ith reaction at the
corresponding reference temperature = kri; for Eqn (4) at reference temperature = krT; for Eqn (12) =
krT´; for hydrogenation of toluene as defined in Eqn (4) = kT; for the formation of toluene as defined
in Eqn (12) = kT´.
Reaction equilibrium constant: for dehydrogenation of MCH as defined in Eqn (3) = KA; formation
of dimethylcyclopentanes as shown in Eqn (13), Pa0 = KD; hydrogenation of toluene as defined in
Eqn (4) = KT; for the formation of toluene as defined in Eqn (12), Pa3 = KT´.
Order of the: dehydrogenation reaction as defined in Eqn (3) = nA; hydrogenation reaction as defined
in Eqn (4) = nT; toluene formation reaction as defined in Eqn (12) = nT´.
No. of data points = N.
Reaction pressure (Pa) = p.
Partial pressure (Pa) of: methylcyclohexane = pA; dimethylcyclopentanes = pD; 1-1,
dimethylcyclopentane = pD1; 1-2, dimethylcyclopentane = pD2; 1-3, dimethylcyclopentane = pD3;
ethylcyclopentane = pE; hydrogen = pH2; 3-methylhexane = pM; toluene = pT.

www.prkm.co.uk
366 Muhammad R. Usman, Faisal M. Alotaibi and Rabya Aslam

Rate of depletion (mol kg-1 s-1) of: methylcyclohexane = (–rA); an ith component = (–ri); toluene
= (–rT);
Rate of formation (mol kg-1 s-1) of: dimethylcyclopentanes = (rD); ethylcyclopentane = (rE); toluene
as defined in Eqn (12) = (rT).
R Universal gas constant, Pa m3 mol-1 K-1
Temperature (K): average reaction temperature = T; at the ith reactor position = Ti; reference
temperature = Tr; ith reference temperature = Tri; reactor wall temperature = Tw.
W Weight of the catalyst, kg.
XA Fractional conversion of methylcyclohexane.
XAe Equilibrium fractional conversion of methylcyclohexane.
Fractional conversion in: an ith overall reaction = Xi; of toluene = XT.
XTe Equilibrium fractional conversion of toluene.
Initial mole fraction of: methylcyclohexane in the vapor phase = yA0; 1-1, dimethylcyclopentane
in the vapour phase = yD10; 1-2, dimethylcyclopentane in the vapour phase = yD20; 1-3,
dimethylcyclopentane in the vapour phase = yD30; ethylcyclopentane in the vapour phase = yE0;
hydrogen in the vapour phase = yH20; 3-methylhexane in the vapour phase = yM0; toluene in the
vapour phase = yT0.
Yield of: methylcyclohexane = YA; dimethylcyclopentanes (combined yield of the isomers) = YD;
ethylcyclopentane = YE; 3-methylhexane = YM; toluene = YT.

REFERENCES
[1] Taube, M., Rippin, D. W. T., Cresswell, D. L. and Knecht, W. (1983) Int. J. Hydrogen Energy, 8, 213-225.
[2] Alhumaidan, F., Cresswell, D. and Garforth, A. (2011) Ind. Eng. Chem. Res., 50, 2509–2522.
[3] Zhang, C., Liang, X. and Liu, S. (2011) Int. J. Hydrogen Energy, 36, 8902-8907.
[4] Shukla, A., Pande, J. V. and Biniwale, R. B. (2012) Int. J. Hydrogen Energy, 37, 3350-3357.
[5] Usman, M., Cresswell, D. and Garforth, A. (2012) Ind. Eng. Chem. Res., 51, 158–170.
[6] Usman, M. R. and Cresswell, D. L. (2013) Int. J. Green Energy., 10, 177–189.
[7] Corma, A., Cid, R. and Lopez Agudo, A. (1979) Can. J. Chem. Eng., 57, 638-642.
[8] García de la Banda, J. F., Corma, A. and Melo, F. V. (1986) Appl. Catal., 26, 103-121.
[9] Coughlan, B. and Keane, M. A. (1990) Catal. Lett., 5, 89–100.
[10] Treacy, M. M. J. and Higgins, J. B. Collection of simulated XRD powder patterns for zeolites. Elsevier,
Amsterdam, 2001.
[11] Usman, M. R., Cresswell, D. L. and Garforth, A. A. (2011) Pet. Sci. Tech., 29, 2247-2257.
[12] Alhumaidan, F., Tsakiris, D., Cresswell, D. and Garforth, A. (2013) Int. J. Hydrogen Energy, 38, 14010–
14032.
[13] Corma, A. and Lopez Agudo, A. (1981) React. Kinet. Catal. Lett., 16, 253–257.
[14] Rohrer, J. C. and Sinfelt, J. H. (1962) J. Phys. Chem., 66, 1193-1194
[15] Jossens, L. W. and Petersen, E. E. (1982) J. Catal., 73, 377-386.
[16] Usman, M. R., Aslam. R. and Alotaibi, F. (2011) Energy Sources, Part A., 33, 2264–2271.
[17] Usman, M. R., Cresswell, D. L. and Garforth, A. A. (2013) ISRN Chem. Eng., Article ID 152893, 2247-
2257.
[18] Stull, D. R., Westrum, Jr., E. F. and Sinke, G. C. The chemical thermodynamics of organic compounds.,
John Wiley and Sons, New York, 1969.
[19] Schildhauer, T. H., Newson, E. and Müller, S. (2001) J. Catal., 198, 355-358.

www.prkm.co.uk

View publication stats

You might also like