Vous êtes sur la page 1sur 37

Accepted Manuscript

Title: Fixed-bed reactor modeling for methanol to dimethyl


ether (DME) reaction over ␥-Alumina using a new practical
reaction rate model

Author: Mohammad Ghavipour Reza Mosayebi Behbahani

PII: S1226-086X(13)00435-8
DOI: http://dx.doi.org/doi:10.1016/j.jiec.2013.09.015
Reference: JIEC 1550

To appear in:

Received date: 11-6-2013


Revised date: 10-9-2013
Accepted date: 12-9-2013

Please cite this article as: M. Ghavipour, R.M. Behbahani, Fixed-bed reactor modeling
for methanol to dimethyl ether (DME) reaction over ␥-Alumina using a new
practical reaction rate model, Journal of Industrial and Engineering Chemistry (2013),
http://dx.doi.org/10.1016/j.jiec.2013.09.015

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Graphical Abstract (for review)

i
cr
us
an
M
ed
pt
ce
Ac

Page 1 of 36
Manuscript

1 1
2
3
4
5 Fixed-bed reactor modeling for methanol to dimethyl ether (DME) reaction
6
7 over γ-Alumina using a new practical reaction rate model
8
9
Mohammad Ghavipoura,1, Reza Mosayebi Behbahania
10
11
a
12 Gas Engineering Department, Petroleum University of Technology, 63431, Ahwaz, Iran

t
13

ip
14 Abstract. Dimethyl ether (DME) synthesis reaction rate was studied over a commercial sample
15
16 of γ-Alumina to investigate the accuracy of the most applicable rate models. Due to the deviation

cr
17
18 of the former proposed models especially at temperatures below 593 K, a new simple empirical
19
rate model was proposed. Besides, previous proposed correlations for methanol dehydration

us
20
21
22 equilibrium constant were examined experimentally and a new equation was developed.
23 Subsequently, one dimensional unsteady state heterogeneous model was applied to simulate
24

an
25 adiabatic and non adiabatic fixed-bed reactors. Temperature profile and methanol conversion
26
27 along the reactors were predicted while varying feed rate and feed temperature.
28
M
29
30 Key words. Methanol dehydration; DME; γ-Alumina; Fixed-bed reactor modeling; Reaction
31
32 rate
33
d

34 1. Introduction
35
te

36
37 Considering environmental pollution, energy security and future oil supplies, the global
38
community is seeking new alternative fuels. A promising alternative fuel could be dimethyl ether
p

39
40
41 (DME) due to low NOx and SOx emission and because it does not have large issues with toxicity,
ce

42
43
production, infrastructure, and transportation [1-3]. It also can be used as hydrogen-rich feed of
44 fuel-cells [4, 5] or as a substitute fuel in domestic appliances [6]. DME synthesis can be
45
Ac

46 performed through two routes: direct synthesis (i.e. syngas to DME over hybrid catalysts) and
47
48 indirect synthesis (i.e. methanol dehydration to DME) [7]. γ-Alumina and HZSM-5 are the most
49
50 common catalysts for DME indirect synthesis reaction. γ-Alumina tends to adsorb water on its
51
52 surface and thereby loses its activity because of its hydrophilic nature. When pure methanol is
53
used as the process feed, the catalyst deactivation occurs very slowly and this hydrophilic nature
54
55
56
1
57 Address correspondence to Mohammad Ghavipour, Gas Engineering Department, Petroleum University
58 of Technology, 63431, Ahwaz, Iran. E-mail: Ghavipour@put.ac.ir , M.Ghavipour@gmail.com
59 Tel/Fax: +98-6115550868
60
61
62
63
64 Page 2 of 36
65
1 2
2
3
4 is not a considerable problem [8]. The γ-Alumina modified by silica or phosphorous has shown
5
6 better performance compared to the untreated one. The best operating temperatures are from 523
7
8 to 673 K and the amounts of coking and by-products are very low [9-11]. The activity of
9
10 HZSM-5 is higher than the γ-Alumina because of more strong acidic sites. Unlike the γ-Alumina,
11
12 which exhibits only Lewis acidity, HZSM-5 has both Lewis (due to the extra-framework

t
13

ip
14
aluminum) and Brønsted acid sites. The reaction takes place at the temperature range of 423 to
15 523 K. The HZSM-5 catalyst is deactivated sooner than the γ-Alumina because a layer of coke
16

cr
17 covers its strong acid sites and it has lower DME selectivity especially at higher temperatures.
18
19 Modification of the strong acidic sites by Na, Si or P inhibits hydrocarbon formation (such as

us
20
21 methane) and thereby enhances the catalyst stability [10-14]. By taking all of the above facts into
22
23 account, it seems that the γ-Alumina is a better choice and its commercial use in large extent for
24

an
25
this reaction proves this opinion.
26
27 The mechanism of methanol adsorption, dehydration and specially, formation of the first C-C
28
M
29 bond and the nature of the intermediates involved, are still not fully understood, but they have
30
31 been studied extensively and different rate equations have been developed till now (see Table 1)
32
33 [15]. Most of these rate equations have been derived from experiments conducted at conditions
d

34
35
far away from an industrial reactor. Almost all of these experiments were performed with diluted
te

36 methanol by nitrogen and/or water, as the reactor feed. In spite of that, to meet the highest
37
38 production in the industrial reactors, pure methanol is consumed as the feed. However, water in
p

39
40 the reactor feed will deactivate the γ-Alumina by covering the catalyst active sites [8] and also
41
ce

42 will retard the DME production reaction (i.e. forward reaction as indicated below) and speed up
43
44 the DME consumption reaction (i.e. backward reaction) according to Le Châtelier's principle in
45
equilibrium reactions.
Ac

46
47
48 2 CH3OH CH3OCH3 + H2O
49
50
51 From another point of view, it should be mentioned that in the previous works, to find the rate
52
53 parameters, differential reactors were used and because they did not use a partially converted
54
55 feed (i.e. a feed that contains DME and water as well as methanol) and the reactor outlet
56 conversion could not exceed 10% due to the restriction of differential reactor assumption; the
57
58 reaction rates were measured at average conversions from zero to less than 5%, causing low
59
60 accuracy in the prediction of the rate parameters especially at higher conversions where the
61
62
63
64 Page 3 of 36
65
1 3
2
3
4 industrial reactors operate. During the present study, integral reactor has been used and the
5
6 methanol dehydration rate model has been investigated based on the realistic experimental data
7
8 (i.e. more similar to the industrial conditions).
9
10
11 Among the proposed kinetic equations for DME direct synthesis from syngas over a bifunctional
12

t
13 catalyst of CuO-ZnO-Al2O3 and γ-Alumina in a fixed-bed reactor [16,17], one of the most

ip
14 common kinetic models [18-20]is the combination of the methanol synthesis model proposed by
15
16 Graaf [21] and the methanol dehydration model proposed by Bercic and Levec [15]. Therefore,

cr
17
18 the new proposed model at present work can also be used in DME direct synthesis investigations
19
as the methanol dehydration model.

us
20
21
22
23 Some studies have been performed on the modeling of fixed-bed reactors for the methanol
24

an
dehydration. Nasehi et al. [22] simulated an industrial adiabatic fixed-bed reactor for DME
25
26 production from methanol dehydration at steady state conditions and found that the difference
27
28 between one and two dimensional modeling for adiabatic fixed bed reactor is negligible. Farsi et
M
29
30 al. [23] simulated an industrial adiabatic reactor of DME synthesis with accompanying feed pre
31
32 heater and controlled it in dynamic conditions. Fazlollahnejad et al. [24] investigated methanol
33
d

34 dehydration in a bench scale adiabatic packed bed reactor. They assessed the effects of weight
35 hourly space velocity and temperature on the methanol conversion. Bercic and Levec [25]
te

36
37 employed one-dimensional heterogeneous and pseudo-homogeneous plug flow models to assess
38
p

39 an adiabatic fixed bed reactor for the catalytic dehydration of methanol to dimethyl ether.
40
41 They found that intraparticle mass transfer was the rate-controlling step while using 3mm γ-
ce

42
43 Alumina pellets as the catalyst. Farsi et al. [26] modeled a shell and tube fixed-bed reactor and
44
45
optimized it for maximum DME production via adjusting the optimal temperature distribution
Ac

46 along the reactor using genetic algorithm. None of these studies has performed an unsteady state
47
48 modeling to show the progress of the reaction along the reactor from the start up to the steady
49
50 state conditions and none of them has discussed in detail the temperature profile through the
51
52 reactor at different weight hourly space velocities (WHSV’s) and feed temperatures. Moreover, a
53
54 clear comparison between adiabatic and non adiabatic fixed bed reactors for this reaction has not
55
56
been accomplished yet.
57
58 In the present study, the methanol conversion to DME over a commercial sample of γ-Alumina
59
60 at different temperatures and WHSV’s was measured experimentally. Due to the considerable
61
62
63
64 Page 4 of 36
65
1 4
2
3
4 difference between the equilibrium constant models proposed recently, a new reliable correlation
5
6 was proposed that matched our experimental data satisfyingly. The Bercic and Levec rate model
7
8 which is the most applicable rate model in the literature, shows lower conversions than our
9
10 experimental data. Therefore, various reaction rate models were examined via nonlinear
11
12 regression to find a new reliable model and a new empirical model based on the experimental

t
13

ip
14
data was proposed. Finally, a one dimensional heterogeneous model was used to study the
15 behavior of fixed bed reactor for the DME synthesis via methanol dehydration. The experimental
16

cr
17 data that had been acquired of our biggest size fixed bed reactor was employed to validate the
18
19 simulation results of non-adiabatic fixed bed reactor. Two types of fixed bed reactors (adiabatic

us
20
21 and nonadiabatic) were simulated and the methanol conversion and the temperature profile of the
22
23 reactor were described in detail.
24

an
25 2. Experiments
26
27
28 In our experiments, a commercial sample of γ-Alumina from Engelhard Corporation in powder
M
29
30 form was used. Several catalyst characterization tests have been performed such as BET, NH3-
31
32 TPD, Bulk density and particle size distribution that are listed in Table 2.
33
d

34
A multipurpose micro reactor and catalyst test setup was employed for this work. Pure liquid
35
te

36 methanol was injected to a pre heater by means of a HPLC metering pump (working range of 0.1
37
38 to 99.9 mL/min), then evaporated in the pre heater at the constant temperature of 453 K. The
p

39
40 reactor consisted of two heating zones. First zone was to raise the feed temperature to a desirable
41
ce

42 level and the other one was to maintain the reactor surrounding at a proper temperature to
43
44 minimize the heat losses. The thermocouples at the second zone were arranged in a way that the
45 operator could fix the outer surface of the stainless steel reactor at a constant temperature (non-
Ac

46
47 adiabatic reactor). There were other thermocouples to report the temperatures of the pre heater,
48
49 the feed, the reactor center and other critical sections.
50
51
52 The experimental data were obtained from three different sizes of reactors (i.e. outside diameter
53
54 of 3/8, 1/2 and 3/4 inches with heights of 0.4, 1 and 1.5 inches respectively). Catalyst weights of
55
1.05, 6.32, 12.64 g with the constant liquid methanol flow rate of 2 mL/min were used (weight
56
57 hourly space velocities of 90, 15 and 7.5 respectively). The temperatures of the reactor zones
58
59 were adjusted in a way that the reactor center temperature varied from 523 to 623 K. All of these
60
61
62
63
64 Page 5 of 36
65
1 5
2
3
4 reactors were assumed to be isothermal at their center temperatures which were measured
5
6 through the submerged thermocouple. For the assessment of our fixed bed reactor modeling, the
7
8 biggest size reactor was used (i.e. outside diameter of 1/2 with height of 8 inches), 40 g catalyst
9
10 was loaded and with different liquid methanol flow rates, various WHSV’s were adjusted.
11
12 Reactor pressure was maintained constant at 3 bar via a back pressure regulator installed at the

t
13

ip
14
reactor exit. Activity of the commercial catalyst decreased during the first day (activity depletion
15 of 1 to 3% that was attributed to the strong unstable acidic sites) and the conversion remained
16

cr
17 almost constant during 120 hours. Therefore, the experimental data were recorded during 30 to
18
19 60 hours after reactor startup. For every specific temperature or WHSV, fresh catalyst was used

us
20
21 and data gathering started after 30 hours of reaction. The effect of coke formation on the catalyst
22
23 activity was neglected, because the experiments were not long enough that the deactivation of
24

an
25
the catalyst could be monitored. Consequently, the proposed model of the present work is
26 applicable to predicting the activity of the fresh catalyst.
27
28
M
29 Gas sample analysis was performed by means of Agilent 6890 Gas Chromatograph equipped
30
31 with flame ionization and thermal conductivity detectors. The column was TRB-5 (Teknokroma
32
33 Co.-(95%) Dimethyl-(5%) diphenylpolysiloxane bonded and cross linked phase-length 30m-
d

34
35
inside diameter 0.53mm) with helium as the carrier gas. The oven temperature was kept constant
te

36 at 313 K and run time was 3 minutes. The 6-port gas sampling valve with a 0.05 mL sample loop
37
38 was used to inject the gas samples into the capillary column. A calibration curve was used to
p

39
40 determine the gas samples composition. The compositions of the reactor outlet gases were
41
ce

42 monitored via the Gas Chromatograph until no change observed by time and steady state
43
44 conditions were obtained.
45
Ac

46 3. Equilibrium constant estimation


47
48
49 Equilibrium constant of the methanol to dimethyl ether reaction has been studied previously and
50
51 some correlations have been derived based on the experimental works or thermodynamic
52
53 calculations. Hayashi and Moffat (1982) developed the equilibrium constant using
54
55 thermodynamic properties at 298 K and heat capacity data at higher temperatures as following
56 [27]:
57
58
59 ln K P = 3440/T- 1.67 lnT+ 2.39 10-4 T+ 0.055 10-6 T2 + 5.496 (1)
60
61
62
63
64 Page 6 of 36
65
1 6
2
3
4 Diep and Walnwright (1987) determined thermodynamic equilibrium constant of the methanol-
5
6 dimethyl ether-water system experimentally at temperatures from 498 to 623 K over a
7
8 commercial -Alumina catalyst in a flow reactor maintained at a constant pressure of 200 kPa
9
10 [27].
11
12

t
13 ln K P = 2835.2/T+1.675 lnT- 2.39 10-4 T- 0.2110-6 T2 - 13.360 (2)

ip
14
15
16 They also calculated the enthalpy of formation (Hf 298) and the free Gibbs energy of formation

cr
17
18 (Gf 298) of dimethyl ether by use of published thermochemical data together with the
19
experimental data to be -180.22 and -109.66 kJ/mol, respectively.

us
20
21
22
23
Finally, the equilibrium constant which is coupled to Bercic and Levec model and has been used
24

an
extensively in fixed-bed reactor modeling papers [22-26].
25
26
27 ln K P = 3138/T+0.86log T+1.33 10-3 T-1.23 10-5 T2 +3.5 10-10 T3 (3)
28
M
29
30 A model based on the thermodynamic calculations has been developed in the present work using
31
32 the thermodynamic properties of the pure substances which are listed in Table 3 [28]. It should
33
d

34 be mentioned that the pressure is not an effective parameter in this reaction due to the conversion
35
te

36 of two moles of methanol to one mole of water and one mole of DME (i.e. the reaction occurs
37
38
with no changes in gaseous moles).
p

39
40 If the variation of the reaction heat with temperature is taken into account, temperature
41
ce

42 dependency of the equilibrium constant (K eq ) will be found by using the integrated form of
43
44 Van't Hoff equation:
45
Ac

46
H
47 T T
K 1
48
49
ln =
K0 R T T 2 r dT & H r =H r 0 +  C
T0
P dT (4)
0
50
51
52 CP =C1 +C2  T+C3  T 2 +C4  T3 +C5  T 4 & Cn = Cn DME +Cn H2O -2Cn MeOH (5)
53
54
55 K 0 =exp(-G 0r /R T0 ) (6)
56
57
58 After integration we will have:
59
60
61
62
63
64 Page 7 of 36
65
1 7
2
3
4
K 1 T C2 C3 C4 C5
5 ln = [C1 ln + (T-T0 )+ (T 2 -T02 )+ (T 3 -T03 )+ (T 4 -T04 )+
6 K0 R T0 2 6 12 20
7 (7)
C2 2 C3 3 C4 4 C5 5
8 (-H r +C1 T0 
0
T0 + T0 + T0 + T0 )  (1/T-1/T0 ) ]
9 2 3 4 5
10
11
12 Finally, the following expression is obtained:

t
13

ip
14 1 T
15 K eq =863.2 exp( (174870 ln -761.555(T-298)+1.127955(T 2 -2982 )-0.00117633(T 3 -2983 )+
16 8314 298 (8)

cr
17 4.68550 10 (T -298 )+44789470.26(1/T-1/298)))
-7 4 4

18
19
To find the best correlation of equilibrium constant, the correlations mentioned above have been

us
20
21
22 used to calculate methanol conversion at different temperatures as bellow:
23
24

an
X eq
25
CDME × CH2O [ (C0MeOH )]2 X 2eq
26
K eq = = 2 = (9)
27 C2MeOH [(1-X eq ) C0MeOH ]2 4(1-X eq ) 2
28
M
29
30
31
X eq is the equilibrium conversion of methanol and C0MeOH is the initial concentration of methanol.
32
33
d

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


35
te

36 The mathematical model was developed based on the following assumptions:


37
38

p

39 Negligible concentration and temperature variations in radial direction (plug flow).


40
41  Heat losses to the surrounding are neglected (for adiabatic reactors).
ce

42
43  There are interfacial gradients of temperature and concentration between solid and gas
44
45 phases (heterogeneous model).
Ac

46
47  Pore diffusion resistance is negligible due to the catalyst powder shape (η=1).
48
49 The mass and energy balances for the gas and solid phases in non adiabatic reactors are
50
51 expressed by the following equations.
52
53
54 Gas phase:
55
56
Ci C
57  = -u s i - k gi a v (Ci -Csis ) (10)
58 t z
59
60
61
62
63
64 Page 8 of 36
65
1 8
2
3
4
T T U
5 g Cpg  = -u s g Cp + h f a v (Tss -T) - 4 (T-Tr ) (11)
6 t z dt
7
8
9 Solid phase:
10
11
Cis
12 (1   ) =i B ri - k gi a v (Ci -Csis ) (12)

t
13 t

ip
14
15
Ts
16
B Cps (1   ) = -i B (H r )ri - h f a v g(Tss -T) (13)

cr
17 t
18
19
For adiabatic reactor modeling, last term of the gas phase energy balance should be omitted.

us
20
21
22
23 Boundary conditions: at t > 0 and z = 0 , C =C0 & T = T0
24

an
25
26 Initial conditions: at t = 0 and all z, C =Ci & T = Ti
27
28
M
29 Previous experiments showed that within the particle size of 0.17 mm, the intra-particle
30
31 resistances are negligible [15, 29]. In our experiments, powder catalyst has been used and when
32
33
the catalyst particle size decreases, effectiveness factor tends to 100% and can be neglected [30].
d

34 Effectiveness factor is defined as bellow:


35
te

36
37 1
38  (-rM ) dV (Actual reaction rate)
= V = (14)
p

39
40 (-rM )|S (Rate predicted from intrinsic kinetics)
41
ce

42
43 Overall mass and heat transfer coefficients between solid and gas phases were estimated from the
44
45
proposed correlations by Cussler and Smith respectively [31, 32].
Ac

46
47 Overall mass transfer coefficient:
48
49
50 k gi =103 (1.17 Re-0.42 Sci -0.67 u g ) (15)
51
52
53 Overall heat transfer coefficient:
54
55
56 hf CP  23 0.458  ud P -0.407
57 ( ) = ( ) (16)
58 CP   K B 
59
60
61
62
63
64 Page 9 of 36
65
1 9
2
3
4 The pressure drop in the bed is calculated by Ergun equation [33]:
5
6
7 dP (1- )3 us (1- ) us 2
8  = 150 + 1.75 (17)
9 l d p2 3 d p 3
10
11
12 The overall heat capacity:

t
13

ip
CP = yi  CPi
14
15 CPi = c1 +c2  T+c3  T2 +c4  T3 +c5  T4 & (18)
16

cr
17
18 The differential equations were converted to algebraic equations by means of Finite Difference
19
method. The reactor length was divided into equal discrete intervals, and the algebraic equations

us
20
21
22
were solved by 4th order Runge-Kutta method for every segment at each time step [34].
23 Calculations were repeated at the next time step with the initial values which were the results of
24

an
25 the previous time step and the iteration continued until no changes observed by time (i.e. steady
26
27 state condition). Because, the reactor length was divided into separate segments that conversion
28
through each one was small enough, differential reactor assumption could be applied for each
M
29
30
31 segment. For each run in a differential reactor, the plug flow performance equation is used as the
32
following form:
33
d

34
35 X A out X A out
W dX A 1 (X A out -X A in )
te

36
37 FA0
= 
X A in
-rA
=
-rA ave 
X A in
dX A =
-rA ave
(19)
38
p

39
40 5. Results and Discussion
41
ce

42
43 As shown in Figure 1, the methanol conversion increases at higher temperatures and lower
44
45 WHSV’s. Higher temperature speeds up the reaction and lower WHSV means more retention
Ac

46
47
time to let the reaction to proceed. For exothermic reactions, temperature rising has a negative
48 effect too. It shifts the reaction to the backward direction and methanol equilibrium conversion
49
50 decreases. This effect defeats the positive effect (i.e. speeding up the forward reaction rate) near
51
52 equilibrium points. Conversion drops at 598 to 623 K in lower WHSV’s (i.e. WHSV=15 or 7.5)
53
54 confirm this fact. DME selectivity was almost 100% at the temperature range of 523 to 623 K.
55
56 The optimum reaction temperature is a function of WHSV and as the WHSV decreases, the best
57 operating temperature decreases too. Roughly, 570 to 620 K can be known as the optimum
58
59 operating temperature range.
60
61
62
63
64 Page 10 of 36
65
1 10
2
3
4 The equilibrium conversions which are calculated from the correlations presented earlier, and
5
6 our experimental data have been depicted together to find the best equilibrium constant model.
7
8 From Figure 1, it is clear that our correlation and the correlation proposed via Diep and
9
10 Walnwright match the experimental data better than other correlations.
11
12

t
13 The reactor modeling program was conducted at the experimental conditions as the same as

ip
14 experiments (i.e. the same temperature, WHSV and feed composition) while Bercic and Levec
15
16 rate model was used. In Figure 2, both of the experimental and modeling results have been

cr
17
18 depicted. From this Figure, it can be concluded that Bercic and Levec model is not valid for
19
temperatures below 593 K, because the parameters in this model were evaluated from the

us
20
21
22 experimental data at three temperatures of 593, 613 and 633 K with an impure methanol as the
23
24
feed. Therefore, another rate equation should be developed, especially for lower temperatures. At

an
25 first, reaction rate as a function of temperature and methanol conversion should be calculated.
26
27 Integral reactor method was used for this goal. Differential reactor assumption is used when the
28
M
29 rate is considered to be constant at all points through the reactor. Since the rate is concentration-
30
31 dependent, this assumption is reasonable only for small conversions or small reactors. But, for
32
33 pure methanol as used in this study, conversion was so large that differential reactor assumption
d

34
35
was not acceptable. Therefore, an integral reactor was assumed and the differential analysis was
te

36 used to find the rate equation:


37
38
p

39 dX A
40 -rA = (20)
W
41 d( )
ce

42 FA0
43
44
45 At first, methanol conversion versus (W/FA0) was plotted. After fitting a smooth curve to the
Ac

46
47 available points, slopes at each point were calculated. Every slope equals reaction rate at that
48
49
special point (i.e. at that conversion). Then, the rates of the reaction at different temperatures and
50 conversions were measured. (See Table 4)
51
52
53 Nonlinear regression toolbox of PASW Statistics 18 was employed to try the most common rate
54
55 expressions suggested earlier. Among all rate expressions that are listed at Table 1, equations
56
57 number v , v and v  were better, but simple reversible first order equation (i.e. number v 
58
59 ) had the lowest deviation from the experimental data. But, why a first order rate equation was
60
61
62
63
64 Page 11 of 36
65
1 11
2
3
4 examined? There are numerous mechanisms and reaction rate models which are proposed for
5
6 this reaction and every one claims to fit the experimental data well. If a number of alternative
7
8 mechanisms fit the data equally well, it is recognized that the selected equation can only be
9
10 considered to be one of the good fits, not one that represents reality. With this admitted, there is
11
12 no reason why we should not use the simplest and easiest-to-handle equation of satisfactory fit.

t
13

ip
14
Also it is mentioned in literature that most catalytic conversion data can be fitted adequately by
15 relatively simple first- or nth-order rate expressions [35]. Therefore, it seems possible to replace
16

cr
17 the multi constant rate equations by an equivalent first-order expression. Although this model
18
19 does not follow any mechanism and is empirical, but it simplifies reactor modeling and increases

us
20
21 the accuracy of the results. Figure 3 indicates that the rate constants obtained by nonlinear
22
23 regression for equation number v  at different temperatures, satisfy reasonably the Arrhenius
24

an
25
theory and its temperature dependency, with pre exponential coefficient of 8537681 L/gcat.h and
26 activation energy of 66828 J/mol and therefore, this equation can be considered as the most
27
28 favorable rate expression.
M
29
30
31 Two types of reactors were simulated (i.e. adiabatic and non adiabatic reactors). In adiabatic
32
33 reactors, it was assumed that there was no heat loss in the radial direction and the reactor wall
d

34
35
was insulated completely. In non-adiabatic reactors, the outside surface of the reactor wall was
te

36 maintained at the particular temperature equaled the feed temperature (i.e. first and second zones
37
38 were at the same temperature). Validity of the simulated non-adiabatic reactor by mathematical
p

39
40 modeling using the new proposed rate model, was tested by varying the methanol flow rate and
41
ce

42 reactor temperatures (i.e. feed temperature and temperature of the outside wall of the reactor) and
43
44 the experimental data of the biggest size reactor (i.e. outside diameter of 3/8 with height of 8
45
Ac

46
inches) were compared to the simulation results. Table 5 and Figure 4 show the results of both
47 modeling and experiments. The average relative error of the modeling predictions was 4 percent.
48
49 Figure 4 and Table 5 confirm the accuracy of the proposed model. Also, the temperature profile
50
51 along the non-adiabatic reactor was investigated. Figure 5 indicates the modeling results and it
52
53 shows that as the feed flow rate increases, the hot spot (i.e. maximum temperature location)
54
55 moves to the end of the reactor and the conversion decreases. When methanol concentration is
56
57
high, at low feed flow rates, the reaction prceeds in forward direction at the reactor beginning.
58 As methanol converts to DME, at the end of the reactor, methanol concentration decreases and
59
60 the forward reaction is retarded, therefore, much more heat is produced at the reactor inlet than
61
62
63
64 Page 12 of 36
65
1 12
2
3
4 the reactor outlet. When the feed flow rate increases, there is not enough retention time for
5
6 methanol molecules to react and to be consumed at the reactor inlet. So, there are still enough
7
8 methanol molecules to maintain the reaction rate high along the reactor and the hot spot moves to
9
10 the end of the reactor. In Figure 6 the temperature profile of the adiabatic reactor is presented. As
11
12 shown in Figures 5 and 6, in non adiabatic reactors, due to high rate of heat transfer from the

t
13

ip
14
reactor wall, after that the reaction reaches to its maximum rate and the hot spot forms, the
15 reaction rate decreases and less heat is produced and the temperature drops down. But, in
16

cr
17 adiabatic reactors, there is no heat loss from the wall and temperature rises along the reactor,
18
19 then the conversion is more than that of non adiabatic reactors at the same initial temperature.

us
20
21 The problem will rise at industrial reactors where it is attempted to maximize the feed flow and
22
23 conversion. In this situation, adiabatic reactors are more sensitive to control [36] and the reactor
24

an
25
may run away due to accumulation of the reaction heat. Another point is that it takes more time
26 for adiabatic reactors to reach to the steady sate conditions in comparison with non adiabatic
27
28 reactors, because in non adiabatic reactors, temperature of the catalyst bed rises until the
M
29
30 temperature difference can cause enough heat transfer from the reactor wall and also the
31
32 convective heat transfer between catalyst bed and gas bulk flow that compensate the heat which
33
d

34 is produced from the reaction and no more temperature rising occurs by time. In spite of that, in
35
adiabatic reactors, the only way to release the heat of the reaction is the convective heat transfer
te

36
37 from catalyst bed to the gas bulk flow; hence temperature of the catalyst bed rises much more
38
p

39 until the thermal equilibrium takes place, so it takes more time to reach the steady state
40
41 condition. Figure 7 and 8 indicate the methanol conversion and the temperature profile for both
ce

42
43 kinds of reactors. These figures are the steady state equivalents of Figures 5 and 6.
44
45
Finally, conversion and temperature profiles of the non adiabatic reactor with different reactor
Ac

46
47 temperatures, in two WHSV’s are depicted in Figures 9 and 10. The results of these modelings
48
49 have been compared to the experimental data at Table 6. It is concluded from Figure 9 and 10
50
51 that in our experimental reactor with 40 grams of catalyst, at WHSV of 30, the minimum
52
53 required temperature of reactor outside wall and feed to reach the equilibrium conversion is
54
55 623K and at WHSV of 15 this value is 573K. Another point is that in non adiabatic reactors at a
56
constant WHSV, as the temperature of the reactor surrounding increases, the hot spot location
57
58 shifts to the reactor inlet and this phenomenon is reasonable because as the medium temperature
59
60
61
62
63
64 Page 13 of 36
65
1 13
2
3
4 increases, the reaction speeds up and most of methanol molecules will react at the reactor inlet
5
6 and a bulk of heat will be released as the reaction heat.
7
8
9 6. Conclusion
10
11
12

t
13 The aim of this work was to examine one of the most applicable rate equations (i.e. Bercic and

ip
14
15 Levec rate model) by comparing experimental data with the simulation results. Needless to say
16
that it was attempted to make the experimental conditions and the commercial ones close

cr
17
18 together. Due to the deviation of the relevant model from the experimental data especially at
19

us
20 temperatures below 593 K, nonlinear regression was employed to try the most common rate
21
22 expressions suggested earlier. Simple reversible first order equation had the lowest deviation
23
24 from the experimental data. The constants obtained by nonlinear regression for this simple

an
25
26 reversible first order equation agreed reasonably well with Arrhenius equation. Also, one
27
28
dimensional unsteady state heterogeneous model was applied to predict the temperature profile
M
29 and methanol conversion along the fixed-bed reactor. Adiabatic and non adiabatic fixed-bed
30
31 reactors were simulated and methanol conversion and thermal behavior of these reactors were
32
33 discussed and the following results were concluded:
d

34
35  The correlations for equilibrium constant of the present work and that of Diep and
te

36
37 Walnwright match the experimental data better than other correlations.
38
 In non adiabatic reactors, as the feed flow rate increases, the hot spot (i.e. maximum
p

39
40
temperature) moves to the end of the reactor and conversion decreases.
41
ce

42  In non adiabatic reactors, due to high rate of heat transfer from the reactor wall, after that
43
44 the reaction reaches to its maximum rate and the hot spot forms, when the reaction rate
45
Ac

46 decreases and less heat is produced, the temperature will drop. But in adiabatic reactors,
47
48 there is no heat loss from the wall and temperature will rise continuously along the
49
50 reactor and the conversion is more than the conversion of non adiabatic reactors at the
51
52
same initial temperature.
53  It takes more time for adiabatic reactors to reach to the steady sate conditions in
54
55 comparison with non adiabatic reactors.
56
57  Conversion and temperature profile of the non adiabatic reactor with different
58
59 temperatures of feed and heating medium at the range of 498 to 623K were calculated
60
61
62
63
64 Page 14 of 36
65
1 14
2
3
4 and the minimum required temperature of heating medium and feed temperature to reach
5
6 the equilibrium conversion at WHSV of 30 and 15, were estimated.
7
8  Another point is that in non adiabatic reactors at a constant feed flow, as the heating
9
10 medium temperature increases, the hot spot location shifts to the reactor inlet.
11
12

t
13

ip
14 Acknowledgments
15
16 The authors acknowledge the Iranian Petrochemical Research and Technology Company for

cr
17
18 their technical and financial supports (Project ID: 0860249001).
19

us
20
21
22 Nomenclature
23
24

an
25 a v = specific catalyst surface area (m2/m3)
26
27 Ci = molar concentration of component i (mol/m3)
28
M
29
30 C p = Heat capacity of gas mixture at constant pressure (J/mol.K)
31
32
33 d p = catalyst particle diameter (m)
d

34
35 d t = reactor diameter (m)
te

36
37
38 h f = overall heat transfer coefficient between gas and solid (J/m2.K.s)
p

39
40
41 k gi = mass transfer coefficient for component i (m/s)
ce

42
43 K = thermal conductivity (J/m.K)
44
45
Ac

46 K eq = equilibrium constant for methanol dehydration reaction


47
48 L = reactor length (m)
49
50
51 Q = liquid methanol flow rate (mL/min)
52
53 r = rate of methanol dehydration reaction (mol/kgcat.s)
54
55
56
R = reactor diameter (m)
57
58 Re = Reynolds number
59
60
61
62
63
64 Page 15 of 36
65
1 15
2
3
4
5
Sci = Schmidt number of component i
6
7 T = temperature (K)
8
9 Tin = Temperature of the reactor inlet feed (K)
10
11
12 Tr = Temperature of the reactor outside wall (K)

t
13

ip
14
15 t = time (s)
16

cr
17 u s = superficial velocity of gas phase (m/s)
18
19 2
U = Overall Heat transfer coefficient (J/m .K)

us
20
21
22 W = catalyst weight (g)
23
24 WHSV = weight hourly space velocity (gMeOH/gcat.hr)

an
25
26
27
X = methanol conversion (dimensionless)
28
M
29 z = axial reactor coordinates (m)
30
31
32
33
d
Greek letters
34
35
 = viscosity of fluid phase (kg/m.s)
te

36
37
38  = density (kg/m3)
p

39
40
41  = catalyst bed void fraction
ce

42
43  = Effectiveness factor (dimensionless)
44
45
H r = heat of reaction (J/molMeOH)
Ac

46
47
48 Subscripts
49
50
51 B = bulk of solid phase (catalyst bed)
52
53 g = in bulk of gas phase
54
55
i = gaseous chemical species involved in the reaction
56
57
58 M = methanol
59
60 s = solid phase
61
62
63
64 Page 16 of 36
65
1 16
2
3
4 W =water
5
6
7 Superscripts
8
9 s = at the catalyst surface
10
11
12 References

t
13

ip
14 [1] M.Ch. Lee, S.B. Seo, J.H. Chung, Y.J. Joo, D.H. Ahn, Fuel. 87 (2008) 2162-2167.
15
16 [2] R.J. Crookes, K.D.H. Bob-Manuel, Energ. Convers. Manage. 48 (2007) 2971-2977.

cr
17
18 [3] T.A. Semelsberger, R.L. Borup, H.L. Greene, J. Power. Sources. 156 (2005) 497-511.
19

us
20
21 [4] T.A. Semelsberger, K.C. Ott, R.L. Borup, H.L. Greene, Appl. Catal. B. 65 (2006) 291-
22 300.
23
24

an
[5] J.H. Yoo, H.G. Choi, Ch.H. Chung, S.M. Cho, J. Power. Sources. 163 (2006) 103-106.
25
26
27 [6] M. Marchionna, R. Patrini, D. Sanfilippo, G. Migliavacca, Fuel. Proc. Tech. 89 (2008)
28 1255-1261.
M
29
30 [7] J.H. Kima, M.J. Park, S.J. Kim, O.Sh. Joo, K.D. Jung, Appl. Catal. A. 264 (2004) 37-41.
31
32
33 [8] F. Raoof, M. Taghizadeh, A. Eliassi, F. Yaripour, Fuel. 87 (2008) 2967-2971.
d

34
35 [9] D. Liu, Ch. Yao, J. Zhang, D. Fang, D. Chen, Fuel. 90 (2011) 1738-1742.
te

36
37 [10] M. Xu, J.H. Lunsford, D.W. Goodman, A. Bhattacharyya, Appl. Catal. A. 149 (1997)
38
289-301.
p

39
40
41 [11] F. Yaripour, F. Baghaei, I. Schmidt, J. Perregaard, Catal. Commun. 6 (2005) 147-152.
ce

42
43 [12] Y.J. Lee, J.M. Kima, J.W. Bae , Ch.H. Shin , K.W. Jun, Fuel. 88 (2009) 1915-1921.
44
45
[13] V. Vishwanathan, K.W. Jun, J.W. Kim, H.S. Roh, Appl. Catal. A. 276 (2004) 251-255.
Ac

46
47
48 [14] S. Hassanpour, F. Yaripour, M. Taghizadeh, Fuel. Process. Technol. 91 (2010) 1212-
49 1221.
50
51 [15] G. Bercic, J. Levec, Ind. Eng. Chem. Res. 31 (1992) 1035-1040.
52
53
54 [16] Z.G. Nie, H.W. Liu, D.H. Liu, W.Y. Ying, D.Y. Fang, Chem. Reac. Eng. Tech. 20
55 (2004) 1-7.
56
57 [17] A.T. Aguayo, J. Eren, D. Mier, J.M. Arandes, M. Olazar, J. Bilbao, Ind. Eng. Chem.
58
59 Res. 46 (2007) 5522-5530.
60
61
62
63
64 Page 17 of 36
65
1 17
2
3
4 [18] K.L. Ng, D. Chadwick, B.A. Toseland, Chem. Eng. Sci. 54 (1999) 3587- 3592.
5
6
7 [19] X.D. Peng, B.A. Toseland, P.J.A. Tijim, Chem. Eng. Sci. 54 (1999) 2787-2792.
8
9 [20] G.R. Moradi, J. Ahmadpour, F. Yaripour, Chem. Eng. J. 144 (2008) 88-95.
10
11 [21] G.H. Graaf, E.J. Stamhuis, A.A.C.M. beenackers, Chem. Eng. Sci. 43 (1988) 3185-
12

t
13 3195.

ip
14
15 [22] S.M. Nasehi, R. Eslamlueyan, A. Jahanmiri, Proc. 11th Chem. Eng. Conf. Iran, Kish
16 Island. (2006).

cr
17
18
19 [23] M. Farsi, R. Eslamloueyan, A. Jahanmiri, Chem. Eng. Process. 50 (2011) 85-94.

us
20
21 [24] M. Fazlollahnejad, M. Taghizadeh, A. Eliassi, G. Bakeri, Chinese. J. Chem. Eng. 17
22 (2009) 630-634.
23
24

an
25 [25] G. Bercic, J. Levec, Ind. Eng. Chem. Res. 32 (1993) 2478-2484.
26
27 [26] M. Farsi, A. Jahanmiri, R. Eslamloueyan, Int. J. Chem. React. Eng. 8 (2010) Article
28 A79.
M
29
30
31
[27] B.T. Diep, M.S. Walnwright, J. Chem. Eng. Data. 32 (1987) 330-333.
32
33 [28] R.H. Perry, D.W. Green, J.O. Maloney, Perry’s Chemical Engineers Handbook, Seventh
d

34 ed., McGraw-Hill, New York, 1997.


35
te

36
37
[29] S.J. Royaee, C. Falamaki, M. Sohrabi, S.S. Ashraf Talesh, Appl. Cat. A. 338 (2008)
38 114-120.
p

39
40 [30] E.B. Nauman, Chemical reactor design, optimization and scale up, McGraw-Hill, New
41
ce

42 York, 2002.
43
44 [31] E.L. Cussler, Diffusion Mass Transfer in Fluid Systems, Cambridge University Press,
45 United Kingdom, 1984.
Ac

46
47
48 [32] J.M. Smith, Chemical Engineering Kinetics, McGraw Hill, New York, 1980.
49
50 [33] J.F. Richardson, J.H. Harker, J.R. Backhurst, Chemical Engineering, fifth ed.,
51 Butterworth–Heinemann, 1999.
52
53
54
[34] J.D. Hoffman, Numerical methods for Engineers and Scientists, McGraw-Hill, New
55 York, 2001.
56
57 [35] C.D. Prater, R.M. Lago, Advances in Catalysis, Academic Press, New York, 1956.
58
59
60
61
62
63
64 Page 18 of 36
65
1 18
2
3
4 [36] F.G. Froment, K.B. Bischoff, Chemical Reactor Analysis and Design, John Wiley &
5
6 Sons, New York, 1979.
7
8
9
10
11
12

t
13

ip
14
15
16

cr
17
18
19

us
20
21
22
23
24

an
25
26
27
28
M
29
30
31
32
33
d

34
35
te

36
37
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 19 of 36
65
1 19
2
3
4 Table 1. Methanol dehydration Reaction rate models
5
6
7
8
9 CD CW
10 k s k M 2 (CM 2  ) 
11 Bercic and Levec,1992 -rM = K
12 (1+2(k M CM )0.5 + k W C W ) 4

t
13

ip
14 k s k M 2 CM 2
15 -rM =
(1+k M CM + k W CW )2 
16 Gates and Johanson,1971

cr
17
18
19
k s k M CM 0.5

us
20 -rM = k
1+k M CM 0.5 + k W CW 
21 Figueras et al.,1971
22
23
24

an
CM 0.5
25 -rM = k v
26 Kallo and Knozinger 1967 CM 0.5 + k 2CW
27
28
M
29
30 Rubio et al.,1980 -rM = k1  CM 0.5 -k 2  CW 0.5 v
31
32
33
d

34 Schmitz,1978 -rM = k1 +k 2  CM v
35
te

36
37
38 This work  C 
-rM = k 0 exp(-E a /RT)  CMeOH  W  v 
p

39  K eq
40  
41
ce

42
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 20 of 36
65
1 20
2
3
4
5
6
7
8
9
10
11
12

t
Table 2. Catalyst characterizations
13

ip
14
15
16

cr
17 Particle size Bulk BET surface NH3-TPD
18 distribution density(g/cm3) area(m2/gcat) Temp.( K) (mmolNH3/gcat)
19
Dp > 0.060mm 98% 402 0.295

us
20
21 Dp > 0.20mm 4% 0.869 183.2 666 0.452
22
23
24

an
25
26
27
28
M
29
30
31
32
33
d

34
35
te

36
37
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 21 of 36
65
1 21
2
3
4 Table 3. Thermodynamic properties of pure substances
5
6
7
Heat Capacity(J/kmol.K) Enthalpy of Gibbs energy
8 substance formation of formation
9 C1 C2 C3 C4 C5
10
(J/kmol) (J/kmol)
11
12
MeOH 1.0580e5 -3.6223e2 9.3790e-1 0 0 -20.0940e7 -16.2320e7

t
13

ip
14 DME 1.1010e5 -1.5747e2 5.1853e-1 0 0 -18.4100e7 -11.2800e7
15
16
H2 O

cr
17 2.7637e5 -2.0901e3 8.125 -1.4116e-2 9.3701e-6 -24.1814e7 -22.8590e7
18
19

us
20
21
22
23
24

an
25
26
27
28
M
29
30
31
32
33
d

34
35
te

36
37
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 22 of 36
65
1 22
2
3
4
5
6 Table 4. Calculated rate values at different conditions based on experimental data
7
8 Liquid
9 Temperature Methanol Methanol Catalyst Rate
methanol flow
10 (K) concentration(mol/L) conversion (%) weight(g) (mol/gcat.h)
11 rate(mL/min)
12 523 0.0709 0.0000 2.00 0.00 0.1460

t
13 523 0.0653 0.0793 2.00 1.05 0.1361

ip
14 523 0.0549 0.2264 2.00 6.32 0.0862
15 523 0.0486 0.3160 2.00 12.6 0.0268
16
0.2710

cr
17 548 0.0677 0.0000 2.00 0.00
18 548 0.0522 0.2295 2.00 1.05 0.2483
19 548 0.0361 0.4668 2.00 6.32 0.1343

us
20 548 0.0250 0.6309 2.00 12.6 -0.0016
21 573 0.0647 0.0000 2.00 0.00 0.3750
22
23 573 0.0340 0.4742 2.00 1.05 0.3381
24 573 0.0223 0.6555 2.00 6.32 0.1528

an
25 573 0.0117 0.8203 2.00 12.6 -0.0679
26
27
28
M
29
30
31
32
33
d

34
35
te

36
37
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 23 of 36
65
1 23
2
3
4
5
6
7
8
9 Table 5. Comparing the modeling and experimental results of non adiabatic reactor at different conditions
10
11 Tin=Tr Liquid methanol Catalyst WHSV Modeling Experimental Relative
12 (K) flow rate(mL/min) weight(g) (grMeOH/grcat.hr) conversion (%) conversion (%) Error(%)

t
13

ip
523 19.0 40.00 22.50 0.2135 0.2211 3.44
14
15 523 38.0 40.00 45.00 0.1341 0.1299 3.23
16 523 76.0 40.00 90.00 0.0701 0.0734 4.49

cr
17 498 25.4 40.00 30.00 0.0745 0.0769 3.12
18 523 25.4 40.00 30.00 0.1771 0.1655 7.01
19 548 25.4 40.00 30.00 0.3212 0.3109 3.31

us
20 573 25.5 40.00 30.00 0.5073 0.5298 4.25
21 598 25.4 40.00 30.00 0.7318 0.7474 2.09
22 623 25.4 40.00 30.00 0.8144 0.7903 3.05
23
498 12.7 40.00 15.00 0.1003 0.1103 9.07
24

an
25 523 12.7 40.00 15.00 0.3038 0.2853 6.48
26 548 12.7 40.00 15.00 0.5439 0.5603 2.93
27 573 12.7 40.00 15.00 0.8210 0.8067 1.77
28 598 12.7 40.00 15.00 0.8336 0.8198 1.68
M
29
30
31
32
33
d

34
35
te

36
37
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 24 of 36
65
1 24
2
3
4
5
6 1
7
8
9 0.9
10
11 0.8
12

t
13

ip
14 0.7
15
16 0.6 Exp. Data WHSV=90

cr
17
18
X

19 0.5 Exp. Data WHSV=15

us
20
21 0.4 Exp. Data WHSV=7.5
22
23 Equi. Conv.(This work)
24 0.3

an
25
Equi. Conv. (Hayashi & Moffat)
26 0.2
27
28 Equi. Conv. (Diep & Walnwright)
0.1
M
29
30 Equi. Conv. (Bercic & Levec)
31 0
32 498 523 548 573 598 623 648
33
d

34 T(K)
35
te

36 Figure 1.Methanol conversion (experimental data) and methanol equilibrium conversion versus
37
temperature
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 25 of 36
65
1 25
2
3
4 0.9
5
6 experiment modeling
7 0.8
8
9 0.7
10
11
12 0.6

t
13

ip
14 0.5
15
X

16

cr
17 0.4
18
19 0.3

us
20
21
22 0.2
23
24 0.1

an
25
26
0
27
28 498 523 548 573 598 623 648
M
29 T(K)
30
31
32 Figure 2. Methanol conversion versus reactor temperature. Comparison between experimental data and
33 modeling results based on Bercic rate model (WHSV=15 g/h.gcat ).
d

34
35
te

36
37
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 26 of 36
65
1 26
2
3
4
5 2
6
7 y = -8038x + 15.966
8 1.5 R² = 0.9868
9
10 Ln(K)
11 1
12

t
13

ip
14
15 0.5
16

cr
17
18 0
19 0.0017 0.00175 0.0018 0.00185 0.0019 0.00195

us
20
21 1/T
22
23
24

an
Figure 3. Arrhenius plot for constants obtained by nonlinear regression for reversible first order rate
25
26
27 Equation
28
M
29
30
31
32
33
d

34
35
te

36
37
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 27 of 36
65
1 27
2
3
4
5 0.9
6 R-squared = 0.997
7 0.8
8
9
10 0.7
11
12

t
Experimental Conversion

13 0.6

ip
14
15 0.5
16

cr
17
18 0.4
19

us
20
21 0.3
22
23 0.2
24

an
25
26 0.1
27
28
M
29 0
30 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
31 Modeling Conversion
32
33
d

34
35 Figure 4. Accuracy assessment of the modeling results with the experimental data
te

36
37
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 28 of 36
65
1 28
2
3
4
5
6
7
8
9
10
11
12

t
13

ip
14
15
16

cr
17
18
19

us
20
21
22 a) b)
23
24

an
25
26
27
28
M
29
30
31
32
33
d

34
35
te

36
37
38
p

39
40 c)
41
ce

42
43 Figure 5.Reactor modeling results. Temperature profile along the non adiabatic reactor at different feed
44 flow rates. (mcat = 40 g and Tr = Tin= 523K ) (a) Q=19ml/min . WHSV=22.5 (b) Q=38ml/min . WHSV=45
45 (c) Q=76ml/min . WHSV=90
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 29 of 36
65
1 29
2
3
4
5
6
7
8
9
10
11
12

t
13

ip
14
15
16

cr
17
18
19

us
20
21
22
23
24

an
25
26
27
28
a) b)
M
29
30
31
32
33
d

34
35
te

36
37
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
c)
48
49 Figure 6. Reactor modeling results .Temperature profile along the adiabatic reactor at different feed flow
50 rates. (mcat = 40 g and Tin= 523K ) (a) Q=19ml/min . WHSV=22.5 (b) Q=38ml/min . WHSV=45 (c)
51 Q=76ml/min . WHSV=90
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 30 of 36
65
1 30
2
3
4
5
6
7
8
9
10
11
12

t
13

ip
14
15
16

cr
17
18
19

us
20
21
22
23
24

an
25
26
27
28
M
29
30
31
32
33
d

34
35
te

36
37
38
p

39
40 Figure 5.Reactor modeling results. Temperature profile along the non adiabatic reactor at different feed
41
ce

flow rates. (mcat = 40 g and Tr = Tin= 523K ) (a) Q=19ml/min . WHSV=22.5 (b) Q=38ml/min . WHSV=45
42
(c) Q=76ml/min . WHSV=90
43
44
45
Ac

46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 31 of 36
65
1 31
2
3
4
5
6
7
8
9
10
11
12

t
13

ip
14
15
16

cr
17
18
19

us
20
21
22
23
24

an
25
26
27
28
M
29
30
31
32
33
d

34
35
te

36
37
38
p

39
40
41
ce

42
43
44
45
Ac

46
47
Figure 6. Reactor modeling results .Temperature profile along the adiabatic reactor at different feed flow
48 rates. (mcat = 40 g and Tin= 523K ) (a) Q=19ml/min . WHSV=22.5 (b) Q=38ml/min . WHSV=45 (c)
49 Q=76ml/min . WHSV=90
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 32 of 36
65
1 32
2
3
4
5 527
6
7 526.5
8
9 Temperature(K) 526
10
11 525.5
12

t
13 525

ip
14
15 524.5
16 Q=19 ml/min

cr
17 524
18 Q=38 ml/min
19 523.5
Q=76 ml/min

us
20
21 523
22
a) 0 5 10 15 20
23 Reactor length(number of segments)
24

an
25
26 0.25
Q=19 ml/min
27
28 Q=38 ml/min
M
29 0.2
Q=76 ml/min
Methanol Conversion

30
31
32 0.15
33
d

34
35 0.1
te

36
37
38 0.05
p

39
40
41
ce

0
42 b) 0 5 10 15 20
43
44 Reactor length(number of segments)
45
Ac

46
47 Figure 7. Reactor modeling results (a)Temperature profile along the non adiabatic reactor (b)Conversion
48 changes along the non adiabatic reactor (Q=19, 38, 76ml/min. (mcat = 40 g and Tr = Tin= 523K ))
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 33 of 36
65
1 33
2
3
4
5 553
6
7
548
8 Q=19 ml/min
9
10 Q=38 ml/min
543
11 Temperature(K)
12 Q=76 ml/min

t
13 538

ip
14
15
16 533

cr
17
18
528
19

us
20
21 523
22
0 5 10 15 20
23 a) Reactor length(number of segments)
24

an
25
26
27 0.35
28 Q=19 ml/min
M
29
0.3
30 Q=38 ml/min
31
0.25 Q=76 ml/min
Methanol conversion

32
33
d

34 0.2
35
te

36
37 0.15
38
p

39 0.1
40
41
ce

0.05
42
43
b)
44 0
45 0 5 10 15 20
Ac

46 Reactor length(number of segments)


47
48
49 Figure 8. Reactor modeling results (a)Temperature profile along the adiabatic reactor (b)Conversion
50 changes along the adiabatic reactor (Q=19, 38, 76 ml/min. (mcat = 40 g and Tin= 523K ))
51
52
53
54
55
56
57
58
59
60
61
62
63
64 Page 34 of 36
65
1 34
2
3
4
5
6 700
7
8
9
10 650
Temperature(K)

11 Tin=498K
12

t
13 Tin=523K

ip
14 600
15 Tin=548K
16 Tin=573K

cr
17
18 Tin=598K
550
19 Tin=623K

us
20
21
22 500
23
0 5 10 15 20
24

an
25 a) Reactor length(number of segments)
26
27
28
M
29 0.9
30
31 0.8
32
33
d
0.7
34
35
Methanol conversion

0.6
te

36 Tin=498K
37
38 0.5 Tin=523K
p

39 Tin=548K
40 0.4
41 Tin=573K
ce

42 0.3
Tin=598K
43
44 0.2 Tin=623K
45
Ac

46 0.1
47
48 0
49 0 5 10 15 20
50
51 b) Reactor length(number of segments)
52
53
54
55
56 Figure 9. Reactor modeling results (a)Temperature profile along the non adiabatic reactor (b)Conversion
57 changes along the non adiabatic reactor (Q=25.4ml/min, mcat=40g, WHSV=30, Tr = Tin)
58
59
60
61
62
63
64 Page 35 of 36
65
1 35
2
3
4
5
6
7 700
8
9 680
10
660
11
12 640

t
13
Temperature(K)

ip
14 620 Tin=498K
15
16 600 Tin=523K

cr
17 Tin=548K
18 580
19 Tin=573K
560

us
20
21 Tin=598K
540
22
23 520
24

an
25 500
26 a) 0 5 10 15 20
27 Reactor length(number of segments)
28
M
29
30
31 0.9
32
33 0.8
d

34
35 0.7
te

36
37
Methanol conversion

0.6
38
Tin=498K
p

39
40 0.5
Tin=523K
41
ce

42 0.4 Tin=548K
43
44 0.3 Tin=573K
45 Tin=598K
Ac

46 0.2
47
48 0.1
49
50
0
51
52 0 5 10 15 20
53 b) Reactor length(number of segments)
54
55
56 Figure 10. Reactor modeling results (a)Temperature profile along the non adiabatic reactor
57 (b)Conversion changes along the non adiabatic reactor (Q=12.7ml/min, mcat=40 g, WHSV=15, Tr = Tin)
58
59
60
61
62
63
64 Page 36 of 36
65

Vous aimerez peut-être aussi