You are on page 1of 10

Journal of Industrial and Engineering Chemistry 20 (2014) 19421951

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


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

Fixed-bed reactor modeling for methanol to dimethyl ether (DME)


reaction over g-Alumina using a new practical reaction rate model
Mohammad Ghavipour *, Reza Mosayebi Behbahani
Gas Engineering Department, Petroleum University of Technology, 63431 Ahwaz, Iran

A R T I C L E I N F O

Article history:
Received 11 June 2013
Accepted 12 September 2013
Available online 19 September 2013
Keywords:
Methanol dehydration
DME
g-Alumina
Fixed-bed reactor modeling
Reaction rate

A B S T R A C T

Dimethyl ether (DME) synthesis reaction rate was studied over a commercial sample of g-Alumina to
investigate the accuracy of the most applicable rate models. Due to the deviation of the former proposed
models especially at temperatures below 593 K, a new simple empirical rate model was proposed.
Besides, previous proposed correlations for methanol dehydration equilibrium constant were examined
experimentally and a new equation was developed. Subsequently, one dimensional unsteady state
heterogeneous model was applied to simulate adiabatic and non-adiabatic xed-bed reactors.
Temperature prole and methanol conversion along the reactors were predicted while varying feed
rate and feed temperature.
2013 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

1. Introduction
Considering environmental pollution, energy security and
future oil supplies, the global community is seeking new
alternative fuels. A promising alternative fuel could be dimethyl
ether (DME) due to low NOx and SOx emission and because it
does not have large issues with toxicity, production, infrastructure, and transportation [13]. It also can be used as hydrogenrich feed of fuel-cells [4,5] or as a substitute fuel in domestic
appliances [6]. DME synthesis can be performed through two
routes: direct synthesis (i.e. syngas to DME over hybrid
catalysts) and indirect synthesis (i.e. methanol dehydration to
DME) [7]. g-Alumina and HZSM-5 are the most common
catalysts for DME indirect synthesis reaction. g-Alumina tends
to adsorb water on its surface and thereby loses its activity
because of its hydrophilic nature. When pure methanol is used
as the process feed, the catalyst deactivation occurs very slowly
and this hydrophilic nature is not a considerable problem [8].
The g-Alumina modied by silica or phosphorous has shown
better performance compared to the untreated one. The best
operating temperatures are from 523 to 673 K and the amounts
of coking and by-products are very low [911]. The activity of
HZSM-5 is higher than the g-Alumina because of more strong
acidic sites. Unlike the g-Alumina, which exhibits only Lewis
acidity, HZSM-5 has both Lewis (due to the extra-framework

* Corresponding author. Tel.: +98 6115550868; fax: +98 6115550868.


E-mail addresses: Ghavipour@put.ac.ir, M.Ghavipour@gmail.com, lghavipour@put.ac.ir (M. Ghavipour).

aluminum) and Brnsted acid sites. The reaction takes place at


the temperature range of 423523 K. The HZSM-5 catalyst is
deactivated sooner than the g-Alumina because a layer of coke
covers its strong acid sites and it has lower DME selectivity
especially at higher temperatures. Modication of the strong
acidic sites by Na, Si or P inhibits hydrocarbon formation (such
as methane) and thereby enhances the catalyst stability [1014].
By taking all of the above facts into account, it seems that the gAlumina is a better choice and its commercial use in large extent
for this reaction proves this opinion.
The mechanism of methanol adsorption, dehydration and
specially, formation of the rst C-C bond and the nature of the
intermediates involved, are still not fully understood, but they
have been studied extensively and different rate equations have
been developed till now (see Table 1) [15]. Most of these rate
equations have been derived from experiments conducted at
conditions far away from an industrial reactor. Almost all of these
experiments were performed with diluted methanol by nitrogen
and/or water, as the reactor feed. In spite of that, to meet the
highest production in the industrial reactors, pure methanol is
consumed as the feed. However, water in the reactor feed will
deactivate the g-Alumina by covering the catalyst active sites [8]
and also will retard the DME production reaction (i.e. forward
reaction as indicated below) and speed up the DME consumption
reaction (i.e. backward reaction) according to Le Chateliers
principle in equilibrium reactions

2CH3 OH ? CH3 OCH3 H2 O

1226-086X/$ see front matter 2013 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jiec.2013.09.015

M. Ghavipour, R.M. Behbahani / Journal of Industrial and Engineering Chemistry 20 (2014) 19421951

Nomenclature

Table 1
Methanol dehydration reaction rate models.
Bercic and Levec (1992) [15]

av
Ci
Cp
dp
dt
hf
kgi
K
Keq
L
Q
r
R
Re
Sci
T
Tin
Tr
t
us
U
W
WHSV
X
z

specic catalyst surface area (m2/m3)


molar concentration of component i (mol/m3)
heat capacity of gas mixture at constant pressure
(J/mol K)
catalyst particle diameter (m)
reactor diameter (m)
overall heat transfer coefcient between gas and
solid (J/m2 K s)
mass transfer coefcient for component i (m/s)
thermal conductivity (J/m K)
equilibrium constant for methanol dehydration
reaction
reactor length (m)
liquid methanol ow rate (mL/min)
rate of methanol dehydration reaction (mol/kgcat s)
reactor diameter (m)
Reynolds number
Schmidt number of component i
temperature (K)
temperature of the reactor inlet feed (K)
temperature of the reactor outside wall (K)
time (s)
supercial velocity of gas phase (m/s)
overall heat transfer coefcient (J/m2 K)
catalyst weight (g)
weight hourly space velocity (gMeOH/gcat h)
methanol conversion (dimensionless)
axial reactor coordinates (m)

Greek letters
m
viscosity of uid phase (kg/m s)
r
density (kg/m3)
e
catalyst bed void fraction
h
effectiveness factor (dimensionless)
DHr
heat of reaction (J/molMeOH)
Subscripts
bulk of solid phase (catalyst bed)
B
in bulk of gas phase
g
gaseous chemical species involved in the reaction
i
methanol
M
solid phase
s
water
W
Superscripts
at the catalyst surface
s

From another point of view, it should be mentioned that in the


previous works, to nd the rate parameters, differential reactors
were used and because they did not use a partially converted feed
(i.e. a feed that contains DME and water as well as methanol) and
the reactor outlet conversion could not exceed 10% due to the
restriction of differential reactor assumption; the reaction rates
were measured at average conversions from zero to less than 5%,
causing low accuracy in the prediction of the rate parameters
especially at higher conversions where the industrial reactors
operate. During the present study, integral reactor has been used
and the methanol dehydration rate model has been investigated

1943

Gates and Johanson (1971) [37]


Figueras et al. (1971) [38]
Kallo and Knozinger (1967) [39]

rM
rM

2 C D C W
ks k2
CM
K
M

I
4

12kM C M 0:5 kW C W
ks k2
C2
M M

II

0:5
k s k M CM
0:5
M CM kW C W

III

C 0:5

IV

1kM C M kW C W 2

rM k 1k

M
rM k C 0:5 k
M

2 CW

Rubio et al. (1980) [40]

0:5
0:5
 k 2  CW
rM k1  CM

Schmitz (1978) [41]


This work

rM k1 k2  C M

VI
VII



C
rM k0 expEa =RT C MeOH  KW
eq

based on the realistic experimental data (i.e. more similar to the


industrial conditions).
Among the proposed kinetic equations for DME direct synthesis
from syngas over a bifunctional catalyst of CuOZnOAl2O3 and gAlumina in a xed-bed reactor [16,17], one of the most common
kinetic models [1820] is the combination of the methanol
synthesis model proposed by Graaf and Stamhuis [21] and the
methanol dehydration model proposed by Bercic and Levec [15].
Therefore, the new proposed model at present work can also be
used in DME direct synthesis investigations as the methanol
dehydration model.
Some studies have been performed on the modeling of xedbed reactors for the methanol dehydration. Nasehi et al. [22]
simulated an industrial adiabatic xed-bed reactor for DME
production from methanol dehydration at steady state conditions
and found that the difference between one and two dimensional
modeling for adiabatic xed bed reactor is negligible. Farsi et al.
[23] simulated an industrial adiabatic reactor of DME synthesis
with accompanying feed pre heater and controlled it in dynamic
conditions. Fazlollahnejad et al. [24] investigated methanol
dehydration in a bench scale adiabatic packed bed reactor. They
assessed the effects of weight hourly space velocity and temperature on the methanol conversion. Bercic and Levec [25] employed
one-dimensional heterogeneous and pseudo-homogeneous plug
ow models to assess an adiabatic xed bed reactor for the
catalytic dehydration of methanol to dimethyl ether. They found
that intraparticle mass transfer was the rate-controlling step while
using 3 mm g-Alumina pellets as the catalyst. Farsi et al. [26]
modeled a shell and tube xed-bed reactor and optimized it for
maximum DME production via adjusting the optimal temperature
distribution along the reactor using genetic algorithm. None of
these studies has performed an unsteady state modeling to show
the progress of the reaction along the reactor from the start up to
the steady state conditions and none of them has discussed in
detail the temperature prole through the reactor at different
weight hourly space velocities (WHSVs) and feed temperatures.
Moreover, a clear comparison between adiabatic and nonadiabatic xed bed reactors for this reaction has not been
accomplished yet.
In the present study, the methanol conversion to DME over a
commercial sample of g-Alumina at different temperatures and
WHSVs was measured experimentally. Due to the considerable
difference between the equilibrium constant models proposed
recently, a new reliable correlation was proposed that matched our
experimental data satisfyingly. The Bercic and Levec rate model
which is the most applicable rate model in the literature, shows
lower conversions than our experimental data. Therefore, various
reaction rate models were examined via non-linear regression to
nd a new reliable model and a new empirical model based on the
experimental data was proposed. Finally, a one dimensional
heterogeneous model was used to study the behavior of xed bed

M. Ghavipour, R.M. Behbahani / Journal of Industrial and Engineering Chemistry 20 (2014) 19421951

1944
Table 2
Catalyst characterizations.
Particle size
distribution

Dp > 0.060 mm 98%


Dp > 0.20 mm 4%

Bulk
density
(g/cm3)

0.869

BET surface
area (m2/gcat)

183.2

NH3-TPD

Temp (K)

mmolNH3 =g cat

402
666

0.295
0.452

reactor for the DME synthesis via methanol dehydration. The


experimental data that had been acquired of our biggest size xed
bed reactor was employed to validate the simulation results of
non-adiabatic xed bed reactor. Two types of xed bed reactors
(adiabatic and non-adiabatic) were simulated and the methanol
conversion and the temperature prole of the reactor were
described in detail.
2. Experiments
In our experiments, a commercial sample of g-Alumina from
Engelhard Corporation in powder form was used. Several catalyst
characterization tests have been performed such as BET, NH3-TPD,
Bulk density and particle size distribution that are listed in Table 2.
A multipurpose micro reactor and catalyst test setup was
employed for this work. Pure liquid methanol was injected to a pre
heater by means of a HPLC metering pump (working range of 0.1
99.9 mL/min), then evaporated in the pre heater at the constant
temperature of 453 K. The reactor consisted of two heating zones.
First zone was to raise the feed temperature to a desirable level and
the other one was to maintain the reactor surrounding at a proper
temperature to minimize the heat losses. The thermocouples at the
second zone were arranged in a way that the operator could x the
outer surface of the stainless steel reactor at a constant
temperature (non-adiabatic reactor). There were other thermocouples to report the temperatures of the pre heater, the feed, the
reactor center and other critical sections.
The experimental data were obtained from three different
sizes of reactors (i.e. outside diameter of 3/8, 1/2 and 3/4 inches
with heights of 0.4, 1 and 1.5 inches respectively). Catalyst
weights of 1.05, 6.32, 12.64 g with the constant liquid methanol
ow rate of 2 mL/min were used (weight hourly space velocities
of 90, 15 and 7.5, respectively). The temperatures of the reactor
zones were adjusted in a way that the reactor center temperature
varied from 523 to 623 K. All of these reactors were assumed to
be isothermal at their center temperatures which were
measured through the submerged thermocouple. For the
assessment of our xed bed reactor modeling, the biggest size
reactor was used (i.e. outside diameter of 1/2 with height of 8
inches), 40 g catalyst was loaded and with different liquid
methanol ow rates, various WHSVs were adjusted. Reactor
pressure was maintained constant at 3 bar via a back pressure
regulator installed at the reactor exit. Activity of the commercial
catalyst decreased during the rst day (activity depletion of 1
3% that was attributed to the strong unstable acidic sites) and the
conversion remained almost constant during 120 h. Therefore,
the experimental data were recorded during 3060 h after
reactor startup. For every specic temperature or WHSV, fresh
catalyst was used and data gathering started after 30 h of
reaction. The effect of coke formation on the catalyst activity was
neglected, because the experiments were not long enough that
the deactivation of the catalyst could be monitored. Consequently, the proposed model of the present work is applicable to
predicting the activity of the fresh catalyst.
Gas sample analysis was performed by means of Agilent 6890
Gas Chromatograph equipped with ame ionization and thermal

conductivity detectors. The column was TRB-5 (Teknokroma Co.(95%) Dimethyl-(5%) diphenylpolysiloxane bonded and cross
linked phase-length 30 m-inside diameter 0.53 mm) with helium
as the carrier gas. The oven temperature was kept constant at
313 K and run time was 3 min. The 6-port gas sampling valve with
a 0.05 mL sample loop was used to inject the gas samples into the
capillary column. A calibration curve was used to determine the
gas samples composition. The compositions of the reactor outlet
gases were monitored via the Gas Chromatograph until no change
observed by time and steady state conditions were obtained.
3. Equilibrium constant estimation
Equilibrium constant of the methanol to dimethyl ether
reaction has been studied previously and some correlations have
been derived based on the experimental works or thermodynamic
calculations. Hayashi and Moffat (1982) developed the equilibrium
constant using thermodynamic properties at 298 K and heat
capacity data at higher temperatures as following [27]:
ln K P 3440=T  1:67 ln T 2:39  104 T 0:055
 106 T 2 5:496

(1)

Diep and Walnwright (1987) determined thermodynamic


equilibrium constant of the methanoldimethyl etherwater
system experimentally at temperatures from 498 to 623 K over
a commercial g-Alumina catalyst in a ow reactor maintained at a
constant pressure of 200 kPa [27]
ln K P 2835:2=T 1:675 ln T  2:39  104 T  0:21
 106 T 2  13:360

(2)

They also calculated the enthalpy of formation (DHf 298) and the
free Gibbs energy of formation (DGf 298) of dimethyl ether by use of
published thermochemical data together with the experimental
data to be 180.22 and 109.66 kJ/mol, respectively.
Finally, the equilibrium constant which is coupled to Bercic and
Levec model and has been used extensively in xed-bed reactor
modeling papers [2226].
ln K P 3138=T 0:86 log T 1:33  103 T  1:23
 105 T 2 3:5  1010 T 3

(3)

A model based on the thermodynamic calculations has been


developed in the present work using the thermodynamic properties of the pure substances which are listed in Table 3 [28]. It
should be mentioned that the pressure is not an effective
parameter in this reaction due to the conversion of two moles
of methanol to one mole of water and one mole of DME (i.e. the
reaction occurs with no changes in gaseous moles).
If the variation of the reaction heat with temperature is taken
into account, temperature dependency of the equilibrium constant
(Keq) will be found by using the integrated form of Vant Hoff
equation:

ln

K
1

K0 R

ZT
T0

DHr
T2

dT

&

DHr DHr0

ZT

rC P dT

(4)

T0

rC P rC 1 rC 2  T rC 3  T 2 rC 4  T 3 rC 5
 T4
Cn

&

DME

rC n

Cn

H2 O

K 0 expDG0r =R T 0

 2C n

MeOH

(5)
(6)

M. Ghavipour, R.M. Behbahani / Journal of Industrial and Engineering Chemistry 20 (2014) 19421951

1945

Table 3
Thermodynamic properties of pure substances.
Substance

Heat capacity (J/kmol K)


C1

C2

C3

C4

C5

MeOH
DME
H2O

1.0580e5
1.1010e5
2.7637e5

3.6223e2
1.5747e2
2.0901e3

9.3790e1
5.1853e1
8.125

0
0
1.4116e2

0
0
9.3701e6

After integration we will have:



K
1
T
rC 2
rC 3 2
rC 1 ln
T  T 0
T  T02
ln

K0 R
T0
2
6
rC 4 3
rC 5 4
T  T03
T  T04

12
20


rC 2 2 rC 3 3 rC 4 4 rC 5 5
T0
T0
T0
T0
DH0r rC 1 T 0
2
3
4
5

 1=T  1=T 0
Finally, the following expression is obtained:


1
T
174870 ln
 761:555T  298
K eq 863:2 exp
8314
298
1:127955T 2  2982  0:00117633T 3  2983 4:68550

 107 T 4  2984 44789470:261=T  1=298
To nd the best correlation of equilibrium constant, the
correlations mentioned above have been used to calculate
methanol conversion at different temperatures as bellow:
h

K eq

i2

eq
0
2
Xeq
C DME  C H2 O
2 CMeOH

2
2
0
CMeOH
41  X eq 2
1  X eq CMeOH 

(9)

0
is the
Xeq is the equilibrium conversion of methanol and CMeOH
initial concentration of methanol.

4. Fixed-bed reactor modeling for methanol to DME reaction

Enthalpy of formation (J/kmol)

Gibbs energy of formation (J/kmol)

20.0940e7
18.4100e7
24.1814e7

16.2320e7
11.2800e7
22.8590e7

rB C ps 1  e

@T s
hi rB DHr r i  h f av gTss  T
@t

(13)

For adiabatic reactor modeling, last term of the gas phase


energy balance should be omitted.
Boundary conditions: at t > 0 and z 0; C C 0 & T T 0
Initial conditions: at t 0 and all z; C C i & T T i
Previous experiments showed that within the particle size of
0.17 mm, the intra-particle resistances are negligible [15,29]. In
our experiments, powder catalyst has been used and when the
catalyst particle size decreases, effectiveness factor tends to 100%
and can be neglected [30]. Effectiveness factor is dened as bellow:
R
1
r M dV
Actual reaction rate
(14)
hV

Rate predicted from intrinsic kinetics


r M jS
Overall mass and heat transfer coefcients between solid and
gas phases were estimated from the proposed correlations by
Cussler and Smith respectively [31,32].
Overall mass transfer coefcient:
kgi 103 1:17Re0:42 Sci 0:67 ug

(15)

Overall heat transfer coefcient:





0:407
C P m 2=3 0:458 rudP

C P rm
K
eB
m
hf

(16)

The pressure drop in the bed is calculated by Ergun equation [33]:




dP
1  e3 mus
1  erus 2
150
1:75
l
d2p e3
d p e3

(17)

The overall heat capacity:


The mathematical model was developed based on the following
assumptions:
 Negligible concentration and temperature variations in radial
direction (plug ow).
 Heat losses to the surrounding are neglected (for adiabatic
reactors).
 There are interfacial gradients of temperature and concentration
between solid and gas phases (heterogeneous model).
 Pore diffusion resistance is negligible due to the catalyst powder
shape (h = 1).
The mass and energy balances for the gas and solid phases in
non-adiabatic reactors are expressed by the following equations.
Gas phase:

@C i
@C
us i  kgi av C i  Ciss
@t
@z

rg C pg e

@T
@T
U
us rg C p
h f av Tss  T  4 T  T r
dt
@t
@z

(10)

(11)

@C is
hi rB r i  kgi av C i  Ciss
@t

&

CP
(18)

The differential equations were converted to algebraic equations by means of Finite Difference method. The reactor length was
divided into equal discrete intervals, and the algebraic equations
were solved by 4th order RungeKutta method for every segment
at each time step [34]. Calculations were repeated at the next time
step with the initial values which were the results of the previous
time step and the iteration continued until no changes observed by
time (i.e. steady state condition). Because, the reactor length was
divided into separate segments that conversion through each one
was small enough, differential reactor assumption could be applied
for each segment. For each run in a differential reactor, the plug
ow performance equation is used as the following form:
W

F A0

X AZ out
X A in

dX A
1

r A r A ave

X AZ out
XA

dX A

X A

out  X A
r A ave

in

(19)

in

5. Results and discussion

Solid phase:
1  e

C Pi c1 c2  T c3  T 2 c4  T 3 c5  T 4
X

yi  C Pi

(12)

As shown in Fig. 1, the methanol conversion increases at higher


temperatures and lower WHSVs. Higher temperature speeds up

M. Ghavipour, R.M. Behbahani / Journal of Industrial and Engineering Chemistry 20 (2014) 19421951

1946
1

0.9

0.9

experiment

0.8

modeling

0.8

0.7
0.7
Exp. Data WHSV=90

0.5

Exp. Data WHSV=15

0.4

Exp. Data WHSV=7.5

0.4

0.3

Equi. Conv.(This work)

0.3

0.5

0.6
0.6

Equi. Conv. (Hayashi & Moat)

0.2

0.2
Equi. Conv. (Diep & Walnwright)

0.1

0.1

Equi. Conv. (Bercic & Levec)


0
498

523

548

573

598

623

648

498

T(K)
Fig. 1. Methanol conversion (experimental data) and methanol equilibrium
conversion versus temperature.

the reaction and lower WHSV means more retention time to let the
reaction to proceed. For exothermic reactions, temperature rising
has a negative effect too. It shifts the reaction to the backward
direction and methanol equilibrium conversion decreases. This
effect defeats the positive effect (i.e. speeding up the forward
reaction rate) near equilibrium points. Conversion drops at 598
623 K in lower WHSVs (i.e. WHSV = 15 or 7.5) conrm this fact.
DME selectivity was almost 100% at the temperature range of 523
623 K. The optimum reaction temperature is a function of WHSV
and as the WHSV decreases, the best operating temperature
decreases too. Roughly, 570620 K can be known as the optimum
operating temperature range.
The equilibrium conversions which are calculated from the
correlations presented earlier, and our experimental data have
been depicted together to nd the best equilibrium constant
model. From Fig. 1, it is clear that our correlation and the
correlation proposed via Diep and Walnwright match the
experimental data better than other correlations.
The reactor modeling program was conducted at the experimental conditions as the same as experiments (i.e. the same
temperature, WHSV and feed composition) while Bercic and Levec
rate model was used. In Fig. 2, both of the experimental and
modeling results have been depicted. From this Figure, it can be
concluded that Bercic and Levec model is not valid for temperatures below 593 K, because the parameters in this model were
evaluated from the experimental data at three temperatures of
593, 613 and 633 K with an impure methanol as the feed.
Therefore, another rate equation should be developed, especially
for lower temperatures. At rst, reaction rate as a function of
temperature and methanol conversion should be calculated.
Integral reactor method was used for this goal. Differential reactor
assumption is used when the rate is considered to be constant at all
points through the reactor. Since the rate is concentrationdependent, this assumption is reasonable only for small conversions or small reactors. But, for pure methanol as used in this study,
conversion was so large that differential reactor assumption was
not acceptable. Therefore, an integral reactor was assumed and the
differential analysis was used to nd the rate equation:
dX
r A  A
d FWA0

(20)

At rst, methanol conversion versus (W/FA0) was plotted. After


tting a smooth curve to the available points, slopes at each point

523

548

573

598

623

648

T(K)
Fig. 2. Methanol conversion versus reactor temperature. Comparison between
experimental data and modeling results based on Bercic rate model (WHSV = 15 g/
h gcat).

were calculated. Every slope equals reaction rate at that special


point (i.e. at that conversion). Then, the rates of the reaction at
different temperatures and conversions were measured. (See
Table 4).
Non-linear regression toolbox of PASW Statistics 18 was
employed to try the most common rate expressions suggested
earlier. Among all rate expressions that are listed at Table 1,
equations number IV, V and VII were better, but simple reversible
rst order equation (i.e. number VII) had the lowest deviation from
the experimental data. But, why a rst order rate equation was
examined? There are numerous mechanisms and reaction rate
models which are proposed for this reaction and every one claims
to t the experimental data well. If a number of alternative
mechanisms t the data equally well, it is recognized that the
selected equation can only be considered to be one of the good ts,
not one that represents reality. With this admitted, there is no
reason why we should not use the simplest and easiest-to-handle
equation of satisfactory t. Also it is mentioned in literature that
most catalytic conversion data can be tted adequately by
relatively simple rst- or nth-order rate expressions [35].
Therefore, it seems possible to replace the multi constant rate
equations by an equivalent rst-order expression. Although this
model does not follow any mechanism and is empirical, but it

Table 4
Calculated rate values at different conditions based on experimental data.
Liquid
Methanol
Catalyst
Rate
Temperature Methanol
(K)
concentration conversion methanol weight (g) (mol/gcat h)
ow rate
(%)
(mol/L)
(mL/min)
523
523
523
523
548
548
548
548
573
573
573
573

0.0709
0.0653
0.0549
0.0486
0.0677
0.0522
0.0361
0.0250
0.0647
0.0340
0.0223
0.0117

0.0000
0.0793
0.2264
0.3160
0.0000
0.2295
0.4668
0.6309
0.0000
0.4742
0.6555
0.8203

2.00
2.00
2.00
2.00
2.00
2.00
2.00
2.00
2.00
2.00
2.00
2.00

0.00
1.05
6.32
12.6
0.00
1.05
6.32
12.6
0.00
1.05
6.32
12.6

0.1460
0.1361
0.0862
0.0268
0.2710
0.2483
0.1343
0.0016
0.3750
0.3381
0.1528
0.0679

M. Ghavipour, R.M. Behbahani / Journal of Industrial and Engineering Chemistry 20 (2014) 19421951

1947

0.9

R-squared = 0.997
y = -8038x + 15.966
R = 0.9868

0.7

0.5

0
0.0017

0.00175

0.00185

0.0018

0.0019

0.00195

1/T
Fig. 3. Arrhenius plot for constants obtained by non-linear regression for reversible
rst order equation rate.

Experimental Conversion

Ln(K)

1.5

0.8

0.6
0.5
0.4
0.3
0.2
0.1

simplies reactor modeling and increases the accuracy of the


results. Fig. 3 indicates that the rate constants obtained by nonlinear regression for equation number VII at different temperatures, satisfy reasonably the Arrhenius theory and its temperature
dependency, with pre exponential coefcient of 8,537,681 L/gcat h
and activation energy of 66,828 J/mol and therefore, this equation
can be considered as the most favorable rate expression.
Two types of reactors were simulated (i.e. adiabatic and nonadiabatic reactors). In adiabatic reactors, it was assumed that there
was no heat loss in the radial direction and the reactor wall was
insulated completely. In non-adiabatic reactors, the outside
surface of the reactor wall was maintained at the particular
temperature equaled the feed temperature (i.e. rst and second

0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Modeling Conversion
Fig. 4. Accuracy assessment of the modeling results with the experimental data.

zones were at the same temperature). Validity of the simulated


non-adiabatic reactor by mathematical modeling using the new
proposed rate model, was tested by varying the methanol ow rate
and reactor temperatures (i.e. feed temperature and temperature
of the outside wall of the reactor) and the experimental data of the
biggest size reactor (i.e. outside diameter of 3/8 with height of 8

Fig. 5. Reactor modeling results. Temperature prole along the non-adiabatic reactor at different feed ow rates (mcat = 40 g and Tr = Tin = 523 K). (a) Q = 19 mL/min;
WHSV = 22.5. (b) Q = 38 mL/min; WHSV = 45. (c) Q = 76 mL/min; WHSV = 90.

M. Ghavipour, R.M. Behbahani / Journal of Industrial and Engineering Chemistry 20 (2014) 19421951

1948

Table 5
Comparing the modeling and experimental results of non-adiabatic reactor at different conditions.
Tin = Tr (K)

Liquid methanol
ow rate (mL/min)

Catalyst
weight (g)

WHSV (gMeOH/gcat h)

Modeling
conversion (%)

Experimental
conversion (%)

Relative error (%)

523
523
523
498
523
548
573
598
623
498
523
548
573
598

19.0
38.0
76.0
25.4
25.4
25.4
25.5
25.4
25.4
12.7
12.7
12.7
12.7
12.7

40.00
40.00
40.00
40.00
40.00
40.00
40.00
40.00
40.00
40.00
40.00
40.00
40.00
40.00

22.50
45.00
90.00
30.00
30.00
30.00
30.00
30.00
30.00
15.00
15.00
15.00
15.00
15.00

0.2135
0.1341
0.0701
0.0745
0.1771
0.3212
0.5073
0.7318
0.8144
0.1003
0.3038
0.5439
0.8210
0.8336

0.2211
0.1299
0.0734
0.0769
0.1655
0.3109
0.5298
0.7474
0.7903
0.1103
0.2853
0.5603
0.8067
0.8198

3.44
3.23
4.49
3.12
7.01
3.31
4.25
2.09
3.05
9.07
6.48
2.93
1.77
1.68

inches) were compared to the simulation results. Table 5 and Fig. 4


show the results of both modeling and experiments. The average
relative error of the modeling predictions was 4 percent. Fig. 4 and
Table 5 conrm the accuracy of the proposed model. Also, the
temperature prole along the non-adiabatic reactor was investigated. Fig. 5 indicates the modeling results and it shows that as the
feed ow rate increases, the hot spot (i.e. maximum temperature
location) moves to the end of the reactor and the conversion
decreases. When methanol concentration is high, at low feed ow

rates, the reaction proceeds in forward direction at the reactor


beginning. As methanol converts to DME, at the end of the reactor,
methanol concentration decreases and the forward reaction is
retarded, therefore, much more heat is produced at the reactor
inlet than the reactor outlet. When the feed ow rate increases,
there is not enough retention time for methanol molecules to react
and to be consumed at the reactor inlet. So, there are still enough
methanol molecules to maintain the reaction rate high along the
reactor and the hot spot moves to the end of the reactor. In Fig. 6

Fig. 6. Reactor modeling results. Temperature prole along the adiabatic reactor at different feed ow rates (mcat = 40 g and Tin = 523 K). (a) Q = 19 mL/min; WHSV = 22.5. (b)
Q = 38 mL/min; WHSV = 45. (c) Q = 76 mL/min; WHSV = 90.

M. Ghavipour, R.M. Behbahani / Journal of Industrial and Engineering Chemistry 20 (2014) 19421951

527
526.5

Temperature(K)

526
525.5
525
524.5
Q=19 ml/min

524

Q=38 ml/min
523.5

Q=76 ml/min

523
0

10

15

20

Reactor length(number of segments)


0.25

Q=19 ml/min
Q=38 ml/min

Methanol Conversion

0.2

Q=76 ml/min

0.15

0.1

0.05

0
0

10

15

20

1949

the temperature prole of the adiabatic reactor is presented. As


shown in Figs. 5 and 6, in non-adiabatic reactors, due to high rate of
heat transfer from the reactor wall, after that the reaction reaches
to its maximum rate and the hot spot forms, the reaction rate
decreases and less heat is produced and the temperature drops
down. But, in adiabatic reactors, there is no heat loss from the wall
and temperature rises along the reactor, then the conversion is
more than that of non-adiabatic reactors at the same initial
temperature. The problem will rise at industrial reactors where it is
attempted to maximize the feed ow and conversion. In this
situation, adiabatic reactors are more sensitive to control [36] and
the reactor may run away due to accumulation of the reaction heat.
Another point is that it takes more time for adiabatic reactors to
reach to the steady sate conditions in comparison with nonadiabatic reactors, because in non-adiabatic reactors, temperature
of the catalyst bed rises until the temperature difference can cause
enough heat transfer from the reactor wall and also the convective
heat transfer between catalyst bed and gas bulk ow that
compensate the heat which is produced from the reaction and
no more temperature rising occurs by time. In spite of that, in
adiabatic reactors, the only way to release the heat of the reaction
is the convective heat transfer from catalyst bed to the gas bulk
ow; hence temperature of the catalyst bed rises much more until
the thermal equilibrium takes place, so it takes more time to reach
the steady state condition. Figs. 7 and 8 indicate the methanol
conversion and the temperature prole for both kinds of reactors.
These gures are the steady state equivalents of Figs. 5 and 6.

Reactor length(number of segments)

Fig. 7. Reactor modeling results. (a) Temperature prole along the non-adiabatic
reactor. (b) Conversion changes along the non-adiabatic reactor (Q = 19, 38, 76 mL/
min (mcat = 40 g and Tr = Tin = 523 K)).

553
Q=19 ml/min

548

Q=38 ml/min
Temperature(K)

700

Temperature(K)

543

Q=76 ml/min

650
Tin=498K
Tin=523K
600

Tin=548K
Tin=573K

538

Tin=598K

550

Tin=623K

533
500

528

10

15

20

Reactor length(number of segments)

523
0

10

15

20

Reactor length(number of segments)

b
b

0.8

0.35
Q=19 ml/min
0.3

0.7

Q=38 ml/min

Methanol conversion

Methanol conversion

0.9

Q=76 ml/min

0.25
0.2
0.15
0.1

0.6

Tin=498K

0.5

Tin=523K

0.4

Tin=548K
Tin=573K

0.3

Tin=598K

0.2

0.05

Tin=623K

0.1

10

15

20

Reactor length(number of segments)


Fig. 8. Reactor modeling results. (a) Temperature prole along the adiabatic reactor.
(b) Conversion changes along the adiabatic reactor (Q = 19, 38, 76 mL/min
(mcat = 40 g and Tin = 523 K)).

10

15

20

Reactor length(number of segments)


Fig. 9. Reactor modeling results. (a) Temperature prole along the non-adiabatic
reactor. (b) Conversion changes along the non-adiabatic reactor (Q = 25.4 mL/min,
mcat = 40 g, WHSV = 30, Tr = Tin).

M. Ghavipour, R.M. Behbahani / Journal of Industrial and Engineering Chemistry 20 (2014) 19421951

1950

700
680
660

Temperature(K)

640
620

Tin=498K

600

Tin=523K

580

Tin=548K
Tin=573K

560

Tin=598K

540
520
500
0

10

15

20

Reactor length(number of segments)

0.9
0.8

Methanol conversion

0.7
0.6
Tin=498K

0.5

Tin=523K
0.4

Tin=548K

0.3

Tin=573K
Tin=598K

0.2
0.1
0
0

10

15

20

Reactor length(number of segments)


Fig. 10. Reactor modeling results. (a) Temperature prole along the non-adiabatic
reactor. (b) Conversion changes along the non-adiabatic reactor (Q = 12.7 mL/min,
mcat = 40 g, WHSV = 15, Tr = Tin).

Finally, conversion and temperature proles of the nonadiabatic reactor with different reactor temperatures, in two
WHSVs are depicted in Figs. 9 and 10. The results of these
modelings have been compared to the experimental data at Table
6. It is concluded from Figs. 9 and 10 that in our experimental
reactor with 40 grams of catalyst, at WHSV of 30, the minimum
required temperature of reactor outside wall and feed to reach the
equilibrium conversion is 623 K and at WHSV of 15 this value is
573 K. Another point is that in non-adiabatic reactors at a constant
WHSV, as the temperature of the reactor surrounding increases,
the hot spot location shifts to the reactor inlet and this
phenomenon is reasonable because as the medium temperature
increases, the reaction speeds up and most of methanol molecules
will react at the reactor inlet and a bulk of heat will be released as
the reaction heat.
6. Conclusion
The aim of this work was to examine one of the most applicable
rate equations (i.e. Bercic and Levec rate model) by comparing
experimental data with the simulation results. Needless to say that
it was attempted to make the experimental conditions and the
commercial ones close together. Due to the deviation of the
relevant model from the experimental data especially at temperatures below 593 K, non-linear regression was employed to try
the most common rate expressions suggested earlier. Simple
reversible rst order equation had the lowest deviation from the

experimental data. The constants obtained by non-linear regression for this simple reversible rst order equation agreed
reasonably well with Arrhenius equation. Also, one dimensional
unsteady state heterogeneous model was applied to predict the
temperature prole and methanol conversion along the xed-bed
reactor. Adiabatic and non-adiabatic xed-bed reactors were
simulated and methanol conversion and thermal behavior of these
reactors were discussed and the following results were concluded:
 The correlations for equilibrium constant of the present work
and that of Diep and Walnwright match the experimental data
better than other correlations.
 In non-adiabatic reactors, as the feed ow rate increases, the hot
spot (i.e. maximum temperature) moves to the end of the reactor
and conversion decreases.
 In non-adiabatic reactors, due to high rate of heat transfer from
the reactor wall, after that the reaction reaches to its maximum
rate and the hot spot forms, when the reaction rate decreases and
less heat is produced, the temperature will drop. But in adiabatic
reactors, there is no heat loss from the wall and temperature will
rise continuously along the reactor and the conversion is more
than the conversion of non-adiabatic reactors at the same initial
temperature.
 It takes more time for adiabatic reactors to reach to the steady
sate conditions in comparison with non-adiabatic reactors.
 Conversion and temperature prole of the non-adiabatic reactor
with different temperatures of feed and heating medium at the
range of 498623 K were calculated and the minimum required
temperature of heating medium and feed temperature to reach
the equilibrium conversion at WHSV of 30 and 15, were
estimated.
 Another point is that in non-adiabatic reactors at a constant feed
ow, as the heating medium temperature increases, the hot spot
location shifts to the reactor inlet.
Acknowledgements
The authors acknowledge the Iranian Petrochemical Research
and Technology Company for their technical and nancial supports
(Project ID: 0860249001).
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]

M.Ch. Lee, S.B. Seo, J.H. Chung, Y.J. Joo, D.H. Ahn, Fuel 87 (2008) 21622167.
R.J. Crookes, K.D.H. Bob-Manuel, Energy Convers. Manage. 48 (2007) 29712977.
T.A. Semelsberger, R.L. Borup, H.L. Greene, J. Power Sources 156 (2005) 497511.
T.A. Semelsberger, K.C. Ott, R.L. Borup, H.L. Greene, Appl. Catal., B 65 (2006) 291
300.
J.H. Yoo, H.G. Choi, Ch.H. Chung, S.M. Cho, J. Power Sources 163 (2006) 103106.
M. Marchionna, R. Patrini, D. Sanlippo, G. Migliavacca, Fuel Process. Technol. 89
(2008) 12551261.
J.H. Kima, M.J. Park, S.J. Kim, O.Sh. Joo, K.D. Jung, Appl. Catal., A 264 (2004) 3741.
F. Raoof, M. Taghizadeh, A. Eliassi, F. Yaripour, Fuel 87 (2008) 29672971.
D. Liu, Ch. Yao, J. Zhang, D. Fang, D. Chen, Fuel 90 (2011) 17381742.
M. Xu, J.H. Lunsford, D.W. Goodman, A. Bhattacharyya, Appl. Catal., A 149 (1997)
289301.
F. Yaripour, F. Baghaei, I. Schmidt, J. Perregaard, Catal. Commun. 6 (2005)
147152.
Y.J. Lee, J.M. Kima, J.W. Bae, Ch.H. Shin, K.W. Jun, Fuel 88 (2009) 19151921.
V. Vishwanathan, K.W. Jun, J.W. Kim, H.S. Roh, Appl. Catal., A 276 (2004) 251255.
S. Hassanpour, F. Yaripour, M. Taghizadeh, Fuel Process. Technol. 91 (2010) 1212
1221.
G. Bercic, J. Levec, Ind. Eng. Chem. Res. 31 (1992) 10351040.
Z.G. Nie, H.W. Liu, D.H. Liu, W.Y. Ying, D.Y. Fang, Chem. Reac. Eng. Tech. 20 (2004)
17.
A.T. Aguayo, J. Eren, D. Mier, J.M. Arandes, M. Olazar, J. Bilbao, Ind. Eng. Chem. Res.
46 (2007) 55225530.
K.L. Ng, D. Chadwick, B.A. Toseland, Chem. Eng. Sci. 54 (1999) 35873592.
X.D. Peng, B.A. Toseland, P.J.A. Tijim, Chem. Eng. Sci. 54 (1999) 27872792.
G.R. Moradi, J. Ahmadpour, F. Yaripour, Chem. Eng. J. 144 (2008) 8895.
G.H. Graaf, E.J. Stamhuis, A.A.C.M. Beenackers, Chem. Eng. Sci. 43 (1988) 3185
3195.

M. Ghavipour, R.M. Behbahani / Journal of Industrial and Engineering Chemistry 20 (2014) 19421951
[22] S.M. Nasehi, R. Eslamlueyan, A. Jahanmiri, Proc. 11th Chem. Eng. Conf. Iran, Kish
Island., 2006.
[23] M. Farsi, R. Eslamloueyan, A. Jahanmiri, Chem. Eng. Process. 50 (2011) 8594.
[24] M. Fazlollahnejad, M. Taghizadeh, A. Eliassi, G. Bakeri, Chin. J. Chem. Eng. 17
(2009) 630634.
[25] G. Bercic, J. Levec, Ind. Eng. Chem. Res. 32 (1993) 24782484.
[26] M. Farsi, A. Jahanmiri, R. Eslamloueyan, Int. J. Chem. React. Eng 8 (2010), Article A79.
[27] B.T. Diep, M.S. Walnwright, J. Chem. Eng. Data 32 (1987) 330333.
[28] R.H. Perry, D.W. Green, J.O. Maloney, Perrys Chemical Engineers Handbook, 70th
ed., McGraw-Hill, New York, 1997.
[29] S.J. Royaee, C. Falamaki, M. Sohrabi, S.S. Ashraf Talesh, Appl. Catal., A 338 (2008)
114120.
[30] E.B. Nauman, Chemical Reactor Design, Optimization and Scale Up, McGraw-Hill,
New York, 2002.
[31] E.L. Cussler, Diffusion Mass Transfer in Fluid Systems, Cambridge University
Press, United Kingdom, 1984.
[32] J.M. Smith, Chemical Engineering Kinetics, McGraw Hill, New York, 1980.

1951

[33] J.F. Richardson, J.H. Harker, J.R. Backhurst, Chemical Engineering, fth ed.,
ButterworthHeinemann, 1999.
[34] J.D. Hoffman, Numerical Methods for Engineers and Scientists, McGraw-Hill,
New York, 2001 .
[35] C.D. Prater, R.M. Lago, Advances in Catalysis, Academic Press, New York, 1956.
[36] F.G. Froment, K.B. Bischoff, Chemical Reactor Analysis and Design, John Wiley &
Sons, New York, 1979.
[37] B.C. Gates, L.N. Johanson, LungmuirHinshelwood kinetics of the dehydration of
methanol catalyzed by cation exchange resin, AICHE J. 17 (1971) 981983.
[38] F. Figueras, A. Nohl, L. Mourgues, Y. Trambouze, Dehydration of methanol and
tert-butyl alcohol on silicaalumina, Trans. Faraday SOC. 67 (1971) 11551163.
[39] D. Kallo, H. Knozinger, Zur Dehydratisierung von Alkoholen an Aliminiumoksid,
Chem. Ing. Tech. 39 (1967) 676680.
[40] F.C. Rubio, S.D. Diaz, D.D. Castillo, J.D. Trujillo, R.A. Alvarez, Deshidratacion
Catalitica de Metanol en Fase vapor, Ing. Quim. (Madrid) 12 (1980) 113119.
[41] G. Schmitz, Deshydration du Methanol Sur Silice-Alumine, Chim. Phys. (1978)
746504355.

You might also like