0% found this document useful (0 votes)
388 views57 pages

Green Methanol Production Process From Indirect CO2

This document summarizes two scenarios for an indirect CO2 conversion process to produce methanol via reverse water gas shift (RWGS) reaction. In the first scenario, CO2 is converted to syngas in a RWGS reactor and then purified before entering the methanol synthesis reactor. In the second scenario, a water-selective membrane coats the RWGS reactor to remove water in-situ, eliminating the need for downstream purification. Both processes are numerically modeled and optimized to maximize methanol productivity. The second scenario achieves higher CO2 conversion, CO yield, and methanol production compared to the first scenario and a conventional process.

Uploaded by

Melinda Fischer
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
388 views57 pages

Green Methanol Production Process From Indirect CO2

This document summarizes two scenarios for an indirect CO2 conversion process to produce methanol via reverse water gas shift (RWGS) reaction. In the first scenario, CO2 is converted to syngas in a RWGS reactor and then purified before entering the methanol synthesis reactor. In the second scenario, a water-selective membrane coats the RWGS reactor to remove water in-situ, eliminating the need for downstream purification. Both processes are numerically modeled and optimized to maximize methanol productivity. The second scenario achieves higher CO2 conversion, CO yield, and methanol production compared to the first scenario and a conventional process.

Uploaded by

Melinda Fischer
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 57

Accepted Manuscript

Title: Green Methanol Production Process from Indirect CO2


Conversion: RWGS Reactor versus RWGS Membrane Reactor

Authors: Fereshteh Samimi, Nazanin Hamedi, Mohammad


Reza Rahimpour

PII: S2213-3437(18)30736-X
DOI: https://doi.org/10.1016/j.jece.2018.102813
Reference: JECE 102813

To appear in:

Received date: 4 October 2018


Revised date: 4 November 2018
Accepted date: 28 November 2018

Please cite this article as: Samimi F, Hamedi N, Rahimpour MR, Green Methanol
Production Process from Indirect CO2 Conversion: RWGS Reactor versus RWGS
Membrane Reactor, Journal of Environmental Chemical Engineering (2018),
https://doi.org/10.1016/j.jece.2018.102813

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.
Green Methanol Production Process from Indirect CO2

Conversion: RWGS Reactor versus RWGS Membrane

Reactor

PT
RI
Fereshteh Samimi, Nazanin Hamedi, Mohammad Reza Rahimpour1

SC
Department of Chemical Engineering, School of Chemical and Petroleum Engineering, Shiraz University, Shiraz

U
71345, Iran

1
N
Corresponding author. Tel.: +98 713 2303071; fax: +98 713 6287294.
A
E-mail address: [email protected] (Prof. M.R. Rahimpour).
M

Graphical abstract:
D
TE
EP
CC
A
Abstract:

Herein, two different scenarios for indirect CO2 conversion process to produce methanol via

reverse water gas shift (RWGS) reaction were investigated and compared. In the first scenario,

CO2 is converted to synthesis gas over Fe2O3/Cr2O3/CuO catalyst in a RWGS reactor and then

PT
after passing a condenser to eliminate water, as an inhibitor of methanol synthesis catalyst, the

product is sent to the methanol production reactor. While in the second, although the process is

RI
totally as the same as the previous one, a water permselective membrane was applied in the

SC
RWGS reactor. In this scenario, the produced water during the RWGS reaction is separated by

the membrane; therefore, the free of water product no longer needed a separator before the

U
methanol synthesis reactor. Both processes were numerically modeled and differential evolution
N
(DE) method was employed to optimize the processes in order to achieve high methanol
A
productivity. Also, the methanol production reactor from these two scenarios were compared
M

with the conventional route (CR), in which methanol is produced from coal and natural gas. The
D

results indicate that, in addition to water removal in the RWGS membrane reactor, a higher CO2
TE

conversion and CO yield and subsequently a more proper synthesis gas composition are

obtained. More importantly, methanol production rate increases 13 ton/day (% 4.15 increase) and
EP

109 ton/day (% 50.23 increase) in this scenario compared to the first scenario and CR. Also

lower water is produced (17% reduction) in the methanol synthesis reactor of the second scenario
CC

respect to the first one.


A

Keywords: CO2 Utilization; RWGS Reactor; H-SOD Membrane; Methanol Production; Process

Modeling

1. Introduction
Global warming is currently one of the greatest threat to humanity and environment. CO2

accumulation in the atmosphere as a cause of fossil fuels combustion has a major contribution to

global warming. CO2 capturing from large stationary emission sources and utilization as a

feedstock for production of fuel products not only address the above problems, but also emerge

PT
as economically viable process [1-7]. Methanol (MeOH) is one of the candidate fuel products for

the conversion of CO2 through a chemical process. As a raw material has a wide industrial

RI
applications for the production of methyl tertbutyl ether, dimethyl ether, dimethyl terephthalate,

SC
formaldehyde, methyl methacrylate, and acetic acid. Methanol can also be regarded as a vehicle

fuel and as a fuel cell hydrogen carrier [8-10]. Hydrogen should be produced via a renewable

U
energy (i.e., water splitting, solar, wind, geothermal, and biomass), in order to reduce the life

N
cycle carbon dioxide emissions in the process. The eco-friendly sources of raw materials yields
A
to an attractive green methanol synthesis process [11].
M

There are two different processes for methanol synthesis through CO2 chemical conversion:
D

direct and indirect conversion of CO2. In the former, CO2 is directly converted into methanol,
TE

while in the later, CO2 is first converted into synthesis gas in a reverse water gas shift (RWGS)

reactor, and then the produced synthesis gas is fed as the feedstock of methanol synthesis reactor
EP

[12]. Newly, Samimi et al. [13] have investigated methanol synthesis process from first route in

three configurations with different cooling systems, namely water cooled, gas cooled and double
CC

cooled reactors. They presented and compared the capability of each configuration in converting
A

CO2 to methanol by focusing on thermal behavior, possibility of second phase formation and

effect of inlet pressure, inlet temperature and CO2/CO molar ratio.

The second route was first suggested by Joo et al. in 1999 [14]. After that some studies regarding

to RWGS reaction were provided in the literature, when the focus of the researches was on the
catalyst development and reaction mechanism. In 2001, Park et al. [15] synthesized and

characterized ZnO/Al2O3 catalyst for the RWGS reaction. Cao et al. [16] investigated on Rh–

Mo6S8 as the catalyst for the RWGS reaction. The mechanism and characteristics of the RWGS

reaction was studied by Kim et al. [17] using Pt/TiO2 and Pt/Al2O3 catalysts. Their results

PT
indicated a higher turnover frequencies and CO2 conversion over Pt/TiO2. Chen and coworkers

[18] synthesized silica-supported Cu nanoparticles for CO2 hydrogenation in RWGS reaction. It

RI
was observed that the reaction mechanism basically involves formate species formation. Anicic

SC
et al. [19] compared economical and energy-efficiency bases of direct and indirect conversion of

CO2 in methanol production process. The results indicated that economically, the price of

U
electricity had the most considerable impact as hydrogen is produced via the water electrolysis.

N
Recently, Samimi et al. [20] have studied CO2 hydrogenation to methanol via RWGS reaction. In
A
their work, the syngas was produced in a RWGS reactor over Ni/Al12O19 catalyst and then the
M

syngas was sent to a methanol synthesis membrane reactor. They have applied a wo-dimensional

model to evaluate both reactors performance.


D

In this work, Fe2O3/Cr2O3/CuO was applied as the catalyst of the RWGS reactor in indirect CO2
TE

conversion process for methanol production. Two scenarios, where the difference between them
EP

is only in syngas production route, were modeled, optimized, and compared. In the first scenario,

CO2 is converted into CO in the RWGS reactor, then the produced syngas passes a separator to
CC

eliminate water, while in the second scenario a water permselective membrane was coated the

wall of RWGS reactor to eliminate water during the reaction. In the second process, there is no
A

need to any separator for water removal from the syngas. A hydroxy sodalite (H-SOD)

membrane was applied in this study for water separation. This membrane is a zeolite-like
material with excellent selectivity of water. The ultimate value of water permeation through H-

SOD was published 10-6 mol/(s m2 Pa) for an ideal case [21-23].

2. Process description

A schematic diagram of the indirect CO2 conversion process for methanol production by two

PT
different scenarios was shown in Fig. 1(a, b).

RI
2.1. The first scenario

SC
The process consists of a RWGS reactor and a methanol synthesis reactor (see Fig. 1(a)). The

feedstock which contains carbon dioxide and hydrogen are divided into two streams: the main is

U
sent to the RWGS reactor, while the reminded part is required to adjust the composition of
N
produced syngas. CO2 and H2 are partially converted to CO and H2O by the RWGS reaction in
A
an adiabatic reactor. The reaction products which include syngas (CO, CO2 and H2) and water
M

are conveyed to a condenser to eliminate water from the stream, as it is a poison for the

Cu/ZnO/Al2O3 catalyst in the next reactor. The produced syngas is then fed into the methanol
D

synthesis reactor, after transmission through a series of compressors and a heat exchanger to
TE

attain the desired temperature and pressure of the reactor. Boiling water in the shell side of the
EP

reactor is used to control the temperature of exothermic reactions. Afterwards, the reactor outlet

is transported to a condenser to separate methanol and water, as the condensable gases from the
CC

unreacted gas. Some part of the unreacted gas is recycled to the reactor to increase conversion.
A

2.2. The second scenario

In this scenario (see Fig. 1(b)), the process is totally as the same as the previous one, except that

an H-SOD membrane for water removal was employed in the RWGS reactor, which can

eliminate the intended separator after the reactor. As demonstrated in Fig. 2, the reactor walls
were coated by the membrane. H2O penetrates through the membrane layer during the reaction.

A sweep gas (N2) flows in the permeation side to carry the permeated H2O.

Figure 1

Figure 2

PT
The specifications and operating conditions of RWGS, RWGS membrane and methanol

RI
synthesis reactors were provided in Tables 1, 2 and 3.

SC
Table 1

U
Table 2

Table 3
N
A
M
3. Reaction scheme and kinetics

3.1. RWGS reaction


D
TE

For green methanol production process, RWGS reaction is needed for production of syngas.

Because of high thermodynamic stability of CO2, high amount of input energy (high
EP

temperature) is required for reaction with H2 to produce CO and H2O. However, temperature

should be low as much as possible to minimize the capital costs and energy losses. CO2
CC

hydrogenation into CO and H2O depends on the catalyst type, CO2/H2 ratio, and the reaction
A

temperature and pressure. The RWGS is described as below:

CO2  H 2  CO  H 2O H 298
0
 41.2 kJ mol CO2 (1)

Fe2O3/Cr2O3/CuO as a well-known commercial catalyst for the water gas shift (WGS) reaction

has been used in order to adjust the H2/CO ratio in the syngas and thus it can be a logical
candidate for the reverse reaction [12]. The kinetics of WGS reaction over this iron-based

catalyst were developed by Hla et al. [24]. They determined the effects of CO, CO2, H2O and H2

concentrations on the WGS reaction rate using four different inlet gas compositions and the

catalyst was found to be more applicable to the gas streams with higher CO2 levels. The reaction

PT
kinetic is described as below:

  88  2.18  0.90.041 0.310.056 0.1560.078 0.050.006


1   

RI
r  100.6590.0125 exp   PCO PH 2O PCO2 PH 2 (2)
 RT 

SC
1  PCO2 PH 2 
   (3)
K  PCO PH O 
 2 

U
 kJ 
where R  8.314  10 
3
 and K is the equilibrium constant of the reaction.
 mol K  N
A
The above kinetic equation is derived at low pressures. Therefore, a pressure scale up correlation
M

is required to apply a kinetic equation applicable at low pressures to higher pressures.


D

r  Fpress~r (4)
TE

where ~
r represents the reaction rate at atmospheric pressure and Fpress indicates a pressure scale
EP

up factor. The following equation was proposed for high temperatures applications:
CC

Fpress  P 0.5P 250 (5)

Where P is the actual reactor pressure in bar, valid up to 30 bar [25].


A

3.2. Methanol synthesis reaction

Methanol synthesis reactions are as below:



 CH3OH
CO  2H 2 
 H 298
0
 90.7kJ / molCO (6)


CO  H 2O
CO2  H 2 
 H 298
0
 41.2kJ / molCO2 (7)


CH3OH  H 2O
CO2  3H 2 
 H 298
0
 49.5kJ / molCO2 (8)

PT
The reactions kinetics of Graaf et al. [26] over Cu/ZnO/Al2O3 catalyst (as the commercial

methanol synthesis catalyst) was selected. The reactions kinetics are expressed as below:

RI
SC
 f 
 CH OH 
kK  f f 3/ 2  3 
1 CO  CO H
2 f 1 / 2K 
H p1
r   2 
1  
(9)

U
K 
   1 / 2  H 2O  
1  KCO fCO  KCO fCO   f H   1 / 2  f 
 2 2   2  K  H 2O 

 
H
2  
N
A
 f f 
 H O CO 
k K f f  2
2 CO  CO H 
2 2 2 K 
M

r   p2 
2  K   (10)
   1 / 2  H 2O  
1  KCO fCO  KCO fCO   f H   1 / 2 f 
 2 2   2  K  H 2O 
 H  
D

 2 
TE

 f f 
 3 / 2 CH OH H O 
3 2

3 CO  CO H 
k K f f
2 2 2 f 3 / 2K 
 H p3 
r  2 
3  
(11)
K 
EP

   1 / 2  H 2O  
1  KCO fCO  KCO fCO   f H   1 / 2  f H O 
 2 2   2  K  2 
H 
  2  
CC

f denotes as the fugacity of the component. The kinetic parameters were presented in Table 4.
A

Table 4

For temperatures higher than 245 ˚C, which occurs in industrial methanol synthesis reactors,

intra-particle diffusion limitations for the commercial catalyst pellet sizes should be taken into
consideration. The dusty-gas model was employed by Lommerts et al. for considering internal

mass transport limitations in methanol synthesis using pseudo first-order reaction rates for the

formation of CH3OH and H2O as diffusion-limiting species [27]:

 C 
 3OH  CH 2  CHeq3OH
 3OH  kCH
rCH  (12)
  3OH
K CH 

PT
 

 CH O 
rH 2O  k H 2O  C H 2  eq2 

RI
 (13)
 K H 2O 

SC
where

U
eq
yCH
K ee
 3OH
CH3OH
y Heq2 N (14)
A
y Heq2O
K eq

M

H 2O (15)
y Heq2
D

The Thiele modulus is then described as:


TE

dp k j  p ( K jeq  1)
j  (16)
EP

6 D eff
j Kj
eq

1  1 1 
CC

j    (17)
j  tanh(3 ) 3 
 j j 

The effective diffusion coefficient, the Knudsen diffusion coefficient and the binary diffusion
A

coefficient are estimated as follow:

1   1 N
yi 
   (18)
eff
DCH 3OH
  DKnudsen,CH OH iCH OH DCH OH ,i 
3 3 3
1   1 N
yi 
   (19)
DH 2O   DKnudsen,H 2O i H 2O DH 2O ,i 
eff

d pore 8RT
DKnudsen,i  (20)
3 MWi

2
 1 1 

PT
3.16  10 T8 1.75  
 MW MW 
Di , j   i j  (21)
 v  13
  v j 
13
 2

RI
P i

τ is equal to 8.13, as the ratio between tortuosity and internal porosity of the catalyst pellet.

SC
4. Mathematical model

U
4.1. RWGS membrane reactor

N
At first, a differential element with the length of dz in the axial direction was taken into
A
consideration to obtain the governing equations. For a steady state, one dimensional
M

homogeneous model, the material balance leads to ordinary differential equations:


D

Mass balances for H2O in the reaction side:


TE

1 d A
 ( F1i )   B ri  ( P ) J H 2O  0 (22)
AC1 dz AC1
EP

i demonstrates the component numerator, F1i is the component molar flow rate, AC1 is the cross
CC

sectional area, AP is perimeter area and ρB is the catalyst bulk density.

Mass balances for H2 in the reaction side:


A

1 d A
 ( F1i )   B ri  0.85 ( P ) J H 2O  0 (23)
AC1 dz AC1

Mass balances for other components in the reaction side:


1 d
 ( F1i )   B ri  0 (24)
AC1 dz

where JH2O is the water permeation in H-SOD membrane and is defined as below:

J H2O  QH2O ( PH2O  PHperm


2O
) (25)

PT
It is worth mentioning that H-SOD membranes are not 100% selective to H2O. Some amounts of

RI
hydrogen also permeate through the membrane. A H2/H2O selectivity equals to 0.85 at 389°C

was considered according to Rohde et al. [23]. QH2O is the H2O permeance reported 10-7–10-6

SC
mol/(s m2 Pa). In the present study, the value of 10-7 mol/(s m2 Pa) was selected for the H2O

permeance.

U
Energy balances for the reaction side: N
A
The RWGS reactor wall is supposed to be adiabatic; hence there is no heat transfer through the
M

reactor wall.
D

C pg1 d N
A A
 ( F1T )   B ri (H f ,i )  U ( P )(T  T perm )  J H 2O C p ,H 2O ( P )(T  Tref )
TE

AC1 dz i 1 AC1 AC1


(26)
A
 (0.85) J H 2O C p , H 2 ( P )(T  Tref )  0
AC1
EP

where T represents the temperature, Cpg1 denotes the gas phase heat capacity, F1 is the total
CC

molar flow rate and ΔHfi demonstrates the heat change of the reaction.

Mass balance for H2O in the permeation side:


A

1 d A
 ( F2i )  P J H 2O  0 (27)
AC1 dz AC1

Mass balance for H2 in the permeation side:


1 d A
 ( F2i )  (0.85) P J H 2O  0 (28)
AC1 dz AC1

Mass balance for N2 in the permeation side:

d
 ( F2i )  0 (29)
dz

PT
Energy balances for the permeation side:

RI
C pg2 d A
 ( F2T perm )  U ( P )(T  T perm )  J H 2O C p ,H 2O (T  Tref )
AC1 dz AC1

SC
(30)
 (0.85) J H 2O C p ,H 2 (T  Tref )  0

U
4.2. Methanol synthesis reactor

N
For a steady state, one dimensional homogeneous model, the following governing equations
A
were derived.
M

Mass balances:
D

1 d
 ( Fi )   B ri  0
TE

(31)
AC dz

Energy balances:
EP

C pg d N Ap
 ( FT )   B ri (H f ,i )  U shell ( )(T  TBW )  0
CC

(32)
AC dz i 1 AC

Since the methanol synthesis reactor is surrounded by the boiling water to adsorb the generated
A

heat, there is a heat transfer between the reaction and boiling water sides. TBW is the boiling water

temperature and Ushell is the overall heat transfer coefficient.

Ergun equation (pressure drop):


For estimation of pressure drop in both RWGS and methanol synthesis reactor, the following

correlation is used.

dP  (1   ) 2 Q  (1   ) Q 2
 150 2 2  1.75 (33)
dz s d p  3 Ac1  s d p  3 Ac21

PT
where Q and P refer to the volumetric flow rate and pressure, respectively; μ, ε, φs, dp, Ac are the

gas phase viscosity, bed porosity, catalyst sphericity, catalyst diameter and the reactor cross

RI
sectional area, respectively.

SC
Boundary conditions:

U
z  0 : yi  yi ,0 , T  T0 , P  P0 (34)

N
The auxiliary correlations applied in the mathematical modeling were summarized in Table 5.
A
Table 5
M

There are a set of ordinary differential equations which are coupled with algebraic equations
D

consisting of ideal gas assumptions, reaction rates, heat and mass transfer coefficients
TE

correlations and physical properties of fluids. These equations form a set of non-linear

algebraic equations using backward finite difference and solved by Gauss-Newton method in
EP

MATLAB programming environment.

4.3. Flash calculation


CC

As noted, there are two separators in the indirect CO2 conversion process which were also

modeled. Flash calculation was used for separator model, in which the inlet temperature and
A

pressure of the feed were known. The detail of the model and equations were provided in our

previous work [28].

5. Optimization of the process


DE (Differential evolution) is a powerful, simple and fast method at numerical optimization and

also is more likely to find a function’s true global optimum [29, 30]. There a lot of works in the

literature in which DE strategy was employed to optimize a process through decision variables

which may be the size of the equipment, the operating conditions or etc [31-38]. Babu and

PT
Angira provided the details of DE algorithm and pseudo code in their work [29]. The procedure

applied in this work is “DE/best/1/bin” with NP (the number of population) equals to 30.

RI
In this study, maximization of MeOH production rate in the both scenarios was considered as the

SC
objective function:

U
f  y MeOH (35)

N
Three decision variables namely, the inlet temperature (T0) and pressure (P0) of the RWGS
A
reactor as well as the fraction of fresh feedstock (n) which is required to adjust the syngas
M

composition were taken into consideration in both scenarios. The ranges of decision variables
D

are:
TE

100 C  T0  450 C (36)


EP

1 bar  P0  30 bar (37)


CC

0  n 1 (38)

MeOH synthesis typically runs at 70-80 bar, and either the feed gas of RWGS reactor or the
A

produced syngas needs to be compressed. RWGS reactor can operate up to 30 bar, thus the

higher bound of 30 bar was considered for the pressure.


One constraint for the temperature bed of MeOH synthesis reactor was also considered in the

optimization:

T  550 K (39)

The maximum temperature during the methanol synthesis reactions was selected to be lower than

PT
550 K to be far away from catalyst deactivation (Cu/ZnO/Al2O3 catalyst cannot be used at

RI
temperatures above 573 K).

SC
Penalty function procedure was then applied to add constraints into the optimization. In this

work, 107 was chosen as the penalty parameter. The objective function finally was described as

U
below:

Minimize:
N
A
M

F   f  107 G (40)
D

where
TE

G  max0, (T  550) (41)


EP

7. Model validation
CC

There are two catalytic reactors in the indirect CO2 conversion process: RWGS reactor and

MeOH synthesis reactor. To prove the accuracy of the RWGS reactor model, the experimental
A

work of Hla et al. [24] was simulated and the results were shown in Fig. 3. CO conversion during

WGS reaction for four different inlet gas compositions (reacting temperature = 450◦C, 0.2 g of

catalyst, wet gas velocity = 79.7 cm s−1 and steam: carbon ratio of 3) were compared. As shown

the model predictions are in a reasonable agreement with the Kaiser et al. data. To verify the
validity of the considered model in MeOH synthesis reactor, the simulation results of a

conventional reactor were compared with the plant data of Shiraz Petrochemical Company under

the same process conditions. As listed in Table 6, there are a good agreement between the model

and the plant data.

PT
Figure 3

RI
Table 6

SC
8. Results and discussion

8.1. The effect of inlet temperature and pressure on the performance of RWGS reactor and

U
RWGS membrane reactor
N
A
It is noteworthy to mention that H2O concentration and stoichiometric number (SN) of the
M

produced syngas as a raw material for methanol synthesis reactor are the two important

parameters need to be considered. The SN is defined as below:


D

y H 2  yCO2
TE

SN  (42)
yCO  yCO2
EP

The SN equal to 2 in a process including H2, CO and CO2 in the feedstock is desired. In the

RWGS reactor, where CO2 and H2 are present in the feed gas, H2/CO2=3, results in SN=2. Also,
CC

H2O concentration in the feedstock of methanol synthesis reactor must not be higher than 1%

vol, as its poisoning effect on the traditional methanol catalyst. For this purpose, in the first
A

scenario, the produced water through the RWGS reactor is eliminated by a separator and then the

syngas is sent to the methanol synthesis reactor, while in the second, this is done by applying a

water permselective membrane in the reactor.


Figure 4 shows the CO2 conversion as a function of RWGS reactor (in the absence of membrane)

inlet temperature and pressure in H2/CO2 inlet ratio equals to 3. A close look reveals that at

temperatures lower than 500 K in all pressure ranges, CO2 conversion is equal to zero, which

means RWGS reaction requires a temperature higher than 500 K to begin. After that, CO2

PT
conversion increases with increasing reactor inlet temperature. This is because in high

temperatures, the thermodynamic equilibrium shifts the RWGS reaction towards CO production.

RI
The effect of pressure is much lower than temperature, however the effect of pressure in the

SC
range of 1-10 bar is more pronounced for CO2 conversion. The inlet pressure higher than 10 bar

has no remarkable influence on the reactor performance.

U
Figure 4
N
A
The effect of inlet temperature and pressure of RWGS membrane reactor was also investigated in
M

Fig. 5. In part (a), CO2 conversion was demonstrated. In contrast to the previous case, the

conversion of CO2 is strongly dependent on the inlet pressure and it increases as the pressure
D

increases. This is because in the reactor, a water permselective membrane was employed to
TE

separate water from the reaction side. By increasing the pressure differences between the

reaction and permeation sides, the driving force for water separation also increases (the
EP

permeation side is operated at 1 bar). Therefore, by removing water from the reaction side (see
CC

part (b)), the RWGS reaction (reaction (1)) is shifted toward more CO production (see part (c)),

and subsequently more CO2 is consumed (part (a)).


A

Figure 5

As demonstrated, for a certain feed flow rate and H2/CO2 ratio, the CO2 conversion and

consequently the yield of CO is proportional to temperature and pressure of the feed stream for
both scenarios. It means that it is possible to gain a composition of the produced syngas

compatible with the methanol production process requirements by changing only the operating

conditions of the RWGS reactor.

8.2. The effect of RWGS reactor inlet temperature on the performance of methanol

PT
synthesis reactor

RI
Methanol production is a strongly function of CO content in the feedstock (just taking a close

look at the reaction network). Thus, it is expected that more CO content in the feed gas results in

SC
a higher methanol yield. On the other hand, the amount of CO content in the feedstock (syngas)

U
directly depends on the RWGS reactor conditions; mainly on the temperature. Therefore, it is

N
clear that RWGS reactor temperature influences directly on the amount of methanol production
A
rate. So in this section the effect of inlet temperature of the RWGS reactor (in the absence of
M

membrane) was verified on the performance of methanol synthesis reactor.

Figures 6(a-e) depict the changes in the mole fractions of different components versus
D

temperature of the RWGS reactor and the length of methanol synthesis reactor. As seen in part
TE

(a), methanol as the main product is produced during the reactions and its concentration
EP

increases with augmentation of the RWGS reactor inlet temperature. Methanol increasing trend

is more pronounced in higher temperatures (>500 K), because of significantly high amount of
CC

CO in the feedstock at these temperatures. An inverse trend is observed for H2O concentration

(see part (b)). This is due to the fact that in the presence of high amount of CO in the feedstock
A

(proportional to high RWGS reactor inlet temperature), the RWGS reaction (Eq. (7)) in the

methanol synthesis reactor is shifted toward the left side (known as WGS reaction), and therefore

the formation of water is reduced.


CO mole fraction distribution along the reactor was illustrated in part (c). At low temperatures of

the RWGS reactor, CO is produced along the methanol reactor. When CO concentration is low

in the feedstock, it is produced by the RWGS reaction (Eq. (7)) and the methanol is formed by

CO2 hydrogenation through reaction (8). As the amount of CO increases in the feedstock (with

PT
increasing the inlet temperature of the RWGS reactor), its generation rate is decreased and

consumed by the WGS reaction (the reverse of Eq. (7)) and its hydrogenation to methanol (Eq.

RI
(6)). Almost an inverse trend occurs for CO2 mole fraction (part (d)). At low temperatures of the

SC
RWGS reactor, the CO content in the produced syngas is very low, consequently CO2 plays the

main source of methanol production (Eq. (8)) and the RWGS reaction (Eq. (7)). Thus at low

U
temperatures, CO2 concentration in the feed stream is remarkably high and it is consumed during

N
the reactor. With increasing the temperature, CO2 concentration in the feedstock is reduced and
A
consequently CO2 is generated during the reactor as the results of the WGS reaction.
M

The changes of H2 mole fraction were also investigated. Fig. 6(e) shows reduction of hydrogen
D

along the reactor, due to CO and CO2 hydrogenation to methanol. However, H2 conversion is
TE

much more in high temperatures due to high CO concentration in the feedstock favors CO

hydrogenation reaction and subsequently consuming more H2. Also the initial concentration of
EP

H2 in lower temperatures is much higher, because of lower H2 conversion in the RWGS reactor

in the lower temperatures.


CC

Figure 6
A

Figure 7 shows temperature profile along the methanol synthesis reactor as a function of the

RWGS reactor temperature. As shown, there is a peak in the temperature profile at the first part

of the methanol synthesis reactor due to exothermic nature of the reactions. However, by passing
the fluid through the reactor, the effect of boiling water becomes more pronounced, and finally it

overcomes the generated heat; and therefore the temperature is reduced. This figure also reveals

that that a rise in the RWGS reactor temperature (proportional to CO content) results to a higher

hot spot in the reactor. This is because, increasing the CO content in the feedstock favors the

PT
exothermic WGS reaction, leads to more heat generation. Also, augmentation of CO

concentration enhances the reaction rate; and subsequently, more heat is generated. Thus higher

RI
concentration of CO in the syngas shifts the equilibrium towards the exothermic reactions and

SC
makes controlling of the reactor thermal behavior more difficult.

Figure 7

U
N
Although methanol production favors in high RWGS reactor temperature (because of high CO
A
content in the syngas feedstock), the temperature must be kept in the least possible way to
M

minimize the capital costs and energy losses. Also higher temperatures of the RWGS reactor lead

to a higher temperature in the methanol reactor bed, contributing to catalyst sintering. Therefore,
D

it is essential to optimize the temperature and pressure of the RWGS reactor in order to keep it as
TE

low as possible and to gain a high methanol yield with a reasonable hot spot along the reactor. In

the next section, the optimization results will be presented.


EP

8.3. Optimization results


CC

As mentioned, to gain high methanol productivity and a reasonable temperature profile along the
A

methanol synthesis reactor, an optimization was performed to find the best values for the fraction

of fresh feed as well as the RWGS reactor inlet temperature and pressure. Table 7 presents the

results of optimization for both scenarios. In the next section, the numerical results of the process

at the optimal conditions were presented.


Table 7

8.3.1. RWGS reactor (RWGS-R) versus RWGS membrane reactor (RWGS-MR)

Figure 8 displays CO2 and H2 conversions along both RWGS-R and RWGS-MR at the optimized

conditions. CO2 and H2 as the reactants are consumed to produce CO and H2O during the

PT
reactions. In RWGS-R, the conversions of CO2 and H2 are increased along the reactor length and

RI
after reaching to the equilibrium follow a constant trend. While in the RWGS-MR, by removing

water during the reaction, thermodynamic equilibrium is shifted toward more CO2 and H2

SC
consumption and thus the higher conversions are achieved compared to the RWGS-R.

U
Figure 8
N
The products mole fractions at the optimized conditions were depicted in Fig. 9. H2O mole
A
fraction in the RWGS-R is increased and then follows a monotonous profile, while in the
M

RWGS-MR, it first rises up to a certain point at the beginning of the reactor (because of the
D

RWGS reaction) and then it diminishes along the rest of the reactor. This is due to the fact that
TE

from the peak point to the end of the reactor, water permeation rate through the membrane is

higher than water production rate of the RWGS reaction, thus it is reduced. On the other hand, as
EP

illustrated in part (b), more CO is produced during water penetration through the membrane
CC

based on the Le Chatelier's principle (see Eq. (1)).

Figure 9
A

Figure 10 shows H2O and H2 mole fractions distributions in the permeation side of the RWGS-

MR. As before said, in addition to H2O, some amounts of H2 also diffuses through the H-SOD

membrane at 389 ºC. This is due to the fact that H2O and H2 kinetic diameters at this temperature
(0.265 nm and 0.29, respectively) are smaller than the pore diameters of the zeolites; both

molecules are separated. The permeation of H2O in the zeolite diminishes with increasing

temperature, and H2O/H2 permselectivity remarkably drops at high temperatures. H2O

permeance is between 1×10-7 and 10-6 mol/(s m2 Pa) in the zeolite membranes. At 389°C, a

PT
H2/H2O selectivity equals to 0.85 was taken into account based on the Rohde et al. work [23].

Therefore, H2O and H2 concentrations are increased in the permeation side during the reactor

RI
length and a lower concentration for H2 is also observed.

SC
Figure 10

U
8.3.2. A comparison among scenario 1, scenario 2 and the conventional route (CR)

N
Indirect CO2 conversion process is a green methanol production route, since the feedstock
A
contains CO2 and renewable H2. Two different scenarios were investigated in this research:
M

scenario 1 and scenario 2. In the first, a RWGS reactor was applied to produce syngas, while in

the second, a water permselective membrane was employed in the RWGS reactor. Here scenario
D

1 and 2 as two proposed routes of indirect CO2 conversion process were compared with the
TE

conventional route (CR) of methanol synthesis from natural gas with the same operating
EP

conditions. Figure 11 compares methanol and water production rates for CR and scenario 1 and

scenario 2. Methanol is produced 217, 313 and 326 ton/day, while water production rate is 88,
CC

149 and 124 ton/day in CR, scenario 1 and scenario 2, respectively. Scenario 2 produces 109 ton

of methanol per day more than CR (50.23 % increase in methanol production rate), which is a
A

considerable amount. In comparison with scenario 1, a lower water production is also formed in

scenario 2 (17% reduction). These contribute to the fact that a higher CO concentration in the

produced syngas for scenario 2 is achieved.


Figure 11

8. Conclusions

In the present work, the indirect CO2 hydrogenation process for green methanol production in an

industrial scale was numerically investigated and optimized via two different routes. In the first

PT
route (scenario 1), the produced syngas in a RWGS reactor passes a separator to remove water as

RI
the catalyst poisoning of methanol synthesis reactor. While in the second route (scenario 2), a

water permselective membrane was applied in the RWGS reactor to separate water during the

SC
reaction and eliminate the intended separator in the process. Also the plausibility of these two

U
scenarios to produce methanol was compared with the conventional methanol synthesis route

N
(CR). The results indicated the superiority of scenario 2 to scenario 1, due to the following
A
reasons:
M

- A higher CO2 conversion and CO yield were achieved in scenario 2.

- A more suitable synthesis gas composition was gained in scenario 2.


D

- There is no need for further removal of water from syngas in separator in scenario 2.
TE

- Methanol production is increased 13 ton/day (% 4.15 increase) in scenario 2 compared to


EP

scenario 1.

- Water production is reduced 17% in scenario 2 compared to scenario 1.


CC

Also in comparison with CR, methanol production was increased 109 ton/day (% 50.23 increase)
A

in scenario 2. Although the work is theoretical in nature, does show strongly the feasibility of the

proposed process. However the costs of the respective implementation will need to be addressed

in future work.

Nomenclature
AC Cross sectional area (m2)
AP Perimeter area (m)
C pg Specific heat of the gas at constant pressure (J K-1mol−1)

C pw Specific heat of water vapor in the reaction side (J K-1mol−1)

dp Particle diameter (m)

F Molar flow rate (mol s-1)

PT
Jw Water permeation rate (mol m-2 s-1)

hf Gas-solid heat transfer coefficient (W m−2 K−1)

RI
hi Heat transfer coefficient between fluid phase and reactor wall in reaction side (W m−2 K−1)

ho Heat transfer coefficient between fluid phase and reactor wall in the permeation side (W m-2

SC
K-1)
ki Rate constants for the ith reaction

U
Ki Adsorption equilibrium constant

Kpi Equilibrium constants N


A
n Fraction of feedstock to adjust syngas composition
M

P Pressure (bar)
Q Volumetric flow rate (m3 s-1)
Qw H2O permeance (mol s-1 m-2 Pa-1)
D

ri Rate of reaction of ith component (mol kg-1 s-1)


TE

T Temperature (K)
U Overall heat transfer coefficient (W m−2 K−1)
y Gas phase mole fraction
EP

z Variable represent length of reactor (m)


CC

Greek letters
 Effectiveness factor
 Density of gas phase (kg m-3)
A

B Bulk density (kg m-3)


p Particle density (kg m-3)
 Viscosity of gas phase (kg m-1 s-1)
s Sphericity

j Thiele modulus
 Void fraction of catalyst bed
H f ,i Enthalpy of formation of component i (J mol−1)

Subscripts

BW Boiling water
perm Permeation side
i Chemical species

PT
0 Inlet conditions
w Water

RI
Abbreviations

SC
CR Conventional route
DE Differential evolution
H-SOD Hydroxy sodalite
MeOH Methanol

U
RWGS Reverse water gas shift
RWGS-R Reverse water gas shift reactor
RWGS-MR
SN
N
Reverse water gas shift membrane reactor
Stoichiometric number
A
WGS Water gas shift
M
D
TE
EP
CC
A
References:

[1] A. Ateka, P. Pérez-Uriarte, M. Gamero, J. Ereña, A.T. Aguayo, J. Bilbao, A comparative


thermodynamic study on the CO2 conversion in the synthesis of methanol and of DME, Energy
120 (2017) 796-804.
[2] N.v.d. Assen, L.J. Müller, A. Steingrube, P. Voll, A. Bardow, Selecting CO2 sources for CO2

PT
utilization by environmental-merit-order curves, Environmental science & technology 50(3)
(2016) 1093-1101.
[3] A. Rafiee, K.R. Khalilpour, D. Milani, M. Panahi, Trends in CO2 conversion and utilization:

RI
A review from process systems Perspective, Journal of Environmental Chemical Engineering
(2018).
[4] P. Kumar, P. With, V.C. Srivastava, R. Glaeser, I.M. Mishra, Conversion of carbon dioxide

SC
along with methanol to dimethyl carbonate over ceria catalyst, Journal of Environmental
Chemical Engineering 3(4) (2015) 2943-2947.
[5] W. Wisaijorn, Y. Poo-arporn, P. Marin, S. Ordóňez, S. Assabumrungrat, P. Praserthdam, D.

U
Saebea, S. Soisuwan, Reduction of carbon dioxide via catalytic hydrogenation over copper-based
catalysts modified by oyster shell-derived calcium oxide, Journal of Environmental Chemical
Engineering 5(4) (2017) 3115-3121. N
[6] P. Akhter, M.A. Farkhondehfal, S. Hernández, M. Hussain, A. Fina, G. Saracco, A.U. Khan,
A
N. Russo, Environmental issues regarding CO2 and recent strategies for alternative fuels through
photocatalytic reduction with titania-based materials, Journal of Environmental Chemical
M

Engineering 4(4) (2016) 3934-3953.


[7] A. Ranjbar, A. Irankhah, S.F. Aghamiri, Reverse water gas shift reaction and CO2
mitigation: nanocrystalline MgO as a support for nickel based catalysts, Journal of
D

Environmental Chemical Engineering 6(4) (2018) 4945-4952.


[8] F. Samimi, M.R. Rahimpour, Direct Methanol Fuel Cell, Methanol, Elsevier2017, pp. 381-
TE

397.
[9] O.Y. Abdelaziz, W.M. Hosny, M.A. Gadalla, F.H. Ashour, I.A. Ashour, C.P. Hulteberg,
Novel process technologies for conversion of carbon dioxide from industrial flue gas streams
EP

into methanol, Journal of CO2 Utilization 21 (2017) 52-63.


[10] N. Meiri, R. Radus, M. Herskowitz, Simulation of novel process of CO2 conversion to
liquid fuels, Journal of CO2 Utilization 17 (2017) 284-289.
[11] S.G. Jadhav, P.D. Vaidya, B.M. Bhanage, J.B. Joshi, Catalytic carbon dioxide
CC

hydrogenation to methanol: a review of recent studies, Chemical Engineering Research and


Design 92(11) (2014) 2557-2567.
[12] M. De Falco, S. Giansante, G. Iaquaniello, L. Barbato, Methanol Production from CO 2 Via
A

Reverse-Water–Gas-Shift Reaction, CO2: A Valuable Source of Carbon, Springer2013, pp. 171-


186.
[13] F. Samimi, M. Feilizadeh, M. Ranjbaran, M. Arjmand, M.R. Rahimpour, Phase stability
analysis on green methanol synthesis process from CO2 hydrogenation in water cooled, gas
cooled and double cooled tubular reactors, Fuel Processing Technology 181 (2018) 375-387.
[14] O.-S. Joo, K.-D. Jung, I. Moon, A.Y. Rozovskii, G.I. Lin, S.-H. Han, S.-J. Uhm, Carbon
dioxide hydrogenation to form methanol via a reverse-water-gas-shift reaction (the CAMERE
process), Industrial & engineering chemistry research 38(5) (1999) 1808-1812.
[15] S.-W. Park, O.-S. Joo, K.-D. Jung, H. Kim, S.-H. Han, Development of ZnO/Al2O3 catalyst
for reverse-water-gas-shift reaction of CAMERE (carbon dioxide hydrogenation to form
methanol via a reverse-water-gas-shift reaction) process, Applied Catalysis A: General 211(1)
(2001) 81-90.
[16] Z. Cao, L. Guo, N. Liu, X. Zheng, W. Li, Y. Shi, J. Guo, Y. Xi, Theoretical study on the
reaction mechanism of reverse water–gas shift reaction using a Rh–Mo 6 S 8 cluster, RSC
Advances 6(110) (2016) 108270-108279.
[17] S.S. Kim, H.H. Lee, S.C. Hong, A study on the effect of support's reducibility on the reverse

PT
water-gas shift reaction over Pt catalysts, Applied Catalysis A: General 423 (2012) 100-107.
[18] C.S. Chen, J.H. Wu, T.W. Lai, Carbon dioxide hydrogenation on Cu nanoparticles, The
Journal of Physical Chemistry C 114(35) (2010) 15021-15028.

RI
[19] B. Anicic, P. Trop, D. Goricanec, Comparison between two methods of methanol
production from carbon dioxide, Energy 77 (2014) 279-289.
[20] F. Samimi, D. Karimipourfard, M.R Rahimpour, Green methanol synthesis process from

SC
carbon dioxide via reverse water gas shift reaction in a membrane reactor, Chemical Engineering
Research and Design (2018).
[21] G. Bercic, J. Levec, Intrinsic and global reaction rate of methanol dehydration over.
gamma.-alumina pellets, Industrial & engineering chemistry research 31(4) (1992) 1035-1040.

U
[22] S. Khajavi, J.C. Jansen, F. Kapteijn, Application of a sodalite membrane reactor in
esterification—Coupling reaction and separation, Catalysis Today 156(3-4) (2010) 132-139.
N
[23] M. Rohde, G. Schaub, S. Khajavi, J. Jansen, F. Kapteijn, Fischer–Tropsch synthesis with in
situ H2O removal–Directions of membrane development, Microporous and Mesoporous
A
Materials 115(1-2) (2008) 123-136.
[24] H. SAN SHWE, D. Park, G. DUFFY, J. EDWARDS, D. ROBERTS, A. ILYUSHECHKIN,
M

L. MORPETH, T. Nguyen, Kinetics of high-temperature water-gas shift reaction over two iron-
based commercial catalysts using simulated coal-derived syngases, Chemical engineering journal
146(1) (2009) 148-154.
D

[25] C. Singh, D.N. Saraf, Simulation of high-temperature water-gas shift reactors, Industrial &
Engineering Chemistry Process Design and Development 16(3) (1977) 313-319.
TE

[26] G. Graaf, E. Stamhuis, A. Beenackers, Kinetics of low-pressure methanol synthesis,


Chemical Engineering Science 43(12) (1988) 3185-3195.
[27] B. Lommerts, G. Graaf, A. Beenackers, Mathematical modeling of internal mass transport
EP

limitations in methanol synthesis, Chemical Engineering Science 55(23) (2000) 5589-5598.


[28] F. Samimi, M.R. Rahimpour, A. Shariati, Development of an Efficient Methanol Production
Process for Direct CO2 Hydrogenation over a Cu/ZnO/Al2O3 Catalyst, Catalysts 7(11) (2017)
CC

332.
[29] B. Babu, R. Angira, Modified differential evolution (MDE) for optimization of non-linear
chemical processes, Computers & chemical engineering 30(6-7) (2006) 989-1002.
[30] R. Storn, K. Price, Differential evolution–a simple and efficient heuristic for global
A

optimization over continuous spaces, Journal of global optimization 11(4) (1997) 341-359.
[31] F. Samimi, M. Bayat, D. Karimipourfard, M. R. Rahimpour, P. Keshavarz, A novel axial-
flow spherical packed-bed membrane reactor for dimethyl ether synthesis: simulation and
optimization, Journal of Natural Gas Science and Engineering 13 (2013) 42-51.
[32] F. Samimi, S. Kabiri, A. Mirvakili, M.R. Rahimpour, Simultaneous dimethyl ether synthesis
and decalin dehydrogenation in an optimized thermally coupled dual membrane reactor, Journal
of Natural Gas Science and Engineering 14 (2013) 77-90.
[33] F. Samimi, S. Kabiri, A. Mirvakili, M. R. Rahimpour, The concept of integrated thermally
double coupled reactor for simultaneous production of methanol, hydrogen and gasoline via
differential evolution method, Journal of Natural Gas Science and Engineering 14 (2013) 144-
157.
[34] A. Mirvakili, F. Samimi, A. Jahanmiri, Optimization of Ammonium Nitrate Thermal
Decomposition in a Fluidized Bed Reactor Using Differential Evolution (DE) Method, Chemical
Engineering Communications 202(5) (2015) 557-568.
[35] F. Samimi, M. Bayat, M. R. Rahimpour, P. Keshavarz, Mathematical modeling and

PT
optimization of DME synthesis in two spherical reactors connected in series, Journal of Natural
Gas Science and Engineering 17 (2014) 33-41.
[36] F. Samimi, S. Kabiri, M.R. Rahimpour, The optimal operating conditions of a thermally

RI
double coupled, dual membrane reactor for simultaneous methanol synthesis, methanol
dehydration and methyl cyclohexane dehydrogenation, Journal of Natural Gas Science and
Engineering 19 (2014) 175-189.

SC
[37] F. Samimi, A. Ahmadi, O. Dehghani, M. R. Rahimpour, DE Approach in Development of a
Detailed Reaction Network for Liquid Phase Selective Hydrogenation of Methylacetylene and
Propadiene in a Trickle Bed Reactor, Industrial & Engineering Chemistry Research 54(1) (2014)
117-129.

U
[38] D. Karimipourfard, N. Nemati, M. Bayat, F. Samimi, M.R. Rahimpour, Mathematical
Modeling and Optimization of Syngas Production Process: A Novel Axial Flow Spherical
N
Packed Bed Tri-Reformer, Chemical Product and Process Modeling.
A
M
D
TE
EP
CC
A
List of Figures:

Fig. 1: A schematic diagram of the indirect CO2 conversion process, (a) scenario 1 versus (b)
scenario 2

Fig. 2: A schematic of RWGS membrane reactor

Fig. 3: A comparison between the RWGS reactor model and Hla et al. (2009) data. CO

PT
conversion during WGS reaction for dry-feed coal-derived syngas (65% CO, 30% H2, 2% CO2,
3% N2), slurry-feed coal-derived syngas (44% CO, 37% H2, 16% CO2, 3% N2), backend of
catalytic membrane reactor (7% CO, 12% H2, 78% CO2, 3% N2), backend of WGS reactor (4%

RI
CO, 55% H2, 38% CO2, 3% N2)

Fig. 4: CO2 conversion as a function of inlet temperature and pressure of RWGS reactor

SC
Fig. 5: (a) CO2 conversion, (b) H2O and (C) CO mole fractions of RWGS membrane reactor a
function of inlet temperature and pressure

U
Fig. 6: The effect of inlet temperature of RWGS reactor (in the absence of membrane) on the
N
mole fractions of different components of methanol synthesis reactor (a) MeOH mole fraction,
(b) H2O mole fraction, (c) CO mole fraction, (d) CO2 mole fraction and (e) H2 mole fraction
A
Fig. 7: The effect of RWGS reactor inlet temperature on the temperature profile of methanol
M

synthesis reactor

Fig. 8: A comparison between CO2 and H2 conversions along RWGS-R and RWGS-MR
D

Fig. 9: A comparison of (a) H2O mole fraction and (b) CO mole fraction along RWGS-R and
TE

RWGS-MR

Fig. 10: H2 and H2O mole fractions in permeation side of RWGS-MR


EP

Fig. 11: A comparison between the performance of CR, scenario 1 and scenario 2 in MeOH and
water production rate
CC
A
A
CC
EPT
ED
M
A
N
U
1(a)
SC
RI
PT
A
CC
EPT
ED
M
A
N
U
1(b)

Fig. 1
SC
RI
PT
A
CC
EPT
ED
M
A
N
U
SC
RI
PT
A
CC
EP
TE
D
M

33
Fig. 2
A
N
U
SC
RI
PT
PT
RI
8
Hla et al. (2009)

SC
7 Model prediction

R2=0.987
6
CO conversion (%)

U
5

4
N
A
M
3

2
D

1
TE

0
Dry-feed Slurry-feed Backend of MR Backend of WGSR
EP

Different inlet gas composition


CC

Fig. 3
A

34
PT
RI
SC
U
0.3

N
A
0.4 0.25
M

0.3 0.2
CO2 conversion

0.2 0.15
TE

0.1
0.1
EP

0 30
800 0.05
700 20
600 10
500
CC

400
300 0
Inlet temperature (K) Inlet pressure (bar)
A

Fig. 4

35
PT
RI
SC
0.7

U
0.6
0.8 N
A
0.5
0.6
M
CO2 conversion

0.4

0.4
0.3
D
TE

0.2 0.2

30
0 0.1
EP

800 20
700
600 10
500
400
300 0
CC

Inlet temperature (K) Inlet pressure (bar)


A

5(a)

36
PT
RI
SC
U
0.055

0.06
N 0.05
A
0.05 0.045
M
H2O mole fraction

0.04
0.04
0.035
0.03
D

0.03
0.02
TE

0.025
0.01
0.02
0
0.015
EP

800
600 0.01
30
0.005
CC

400 20
10
Inlet temperature (K) 200 0
Inlet pressure (bar)
A

5(b)

37
A
CC
EP
TE
D
M

38
A
N
U
SC
RI
PT
0.3

0.4
0.25

0.3

PT
CO mole fraction

0.2

0.2
0.15

RI
0.1
0.1

SC
0 30
800 0.05
700 20

U
600
500 10
400
Inlet temperature (K)
300
N0
Inlet pressure (bar)
A
M

5(c)
D

Fig. 5
TE
EP
CC
A

39
0.16

0.14

0.2
0.12
MeOH mole fraction

0.15 0.1

PT
0.08
0.1

RI
0.06
0.05

SC
0.04
0
1 0.02
700

U
0.5 600
Dimensionless 500 0
reactor length 0
N
400
RWGS reactor inlet temperature (K)
A
M

6(a)
D
TE
EP
CC
A

40
0.08

0.1 0.07

0.06
0.08
H2O mole fraction

0.05

PT
0.06
0.04

RI
0.04
0.03

SC
0.02 0.02

1
0 0.01

U
300 0.5
400 500 600 700 0

RWGS reactor inlet temperature (K)


800
N Dimensionless reactor length
A
M

6(b)
D
TE
EP
CC
A

41
PT
0.3
0.35
0.25

RI
0.3

0.25
CO mole fraction

0.2

SC
0.2

0.15 0.15

U
0.1
0.1
0.05 N
A
0 0 0.05
700
600 0.5
M

500
400 1
Dimensionless reactor length
RWGS reactor inlet temperature (K)
D
TE

6(c)
EP
CC
A

42
0.24

0.23

0.26 0.22
0.24 0.21
CO2 mole fraction

PT
0.22 0.2
0.2 0.19

RI
0.18 0.18
0.16

SC
0.17
1
0.16
400 0.5

U
500 0.15
600
700

RWGS reactor inlet temperature (K)


N
0 Dimensionless reactor length
A
M

6(d)
D
TE
EP
CC
A

43
0.7
0.8

0.65
0.7

PT
H2 mole fraction

0.6 0.6

RI
0.5 0.55

0.4

SC
0.5

0 0.45

U
400 0.4
0.5

700
N600
500
A
1
Dimensionless reactor length RWGS reactor inlet temperature (K)
M
D

6(e)
TE

Fig. 6
EP
CC
A

44
580

580 570
570
560

PT
560
Temperature (K)

550 550

RI
540
540
530

SC
530
520

510 520

U
500 1
300 510
400
500
600
700
N 0
0.5
A
800
Dimensionless reactor length
RWGS reactor inlet temperature (K)
M

Fig. 7
D
TE
EP
CC
A

45
PT
0.7
RWGS-R CO
2

RI
0.6 RWGS-MR

H
2

SC
0.5
Conversion

0.4

U
CO
2
0.3
N
A
0.2
M
H2
0.1
D

0
0 0.2 0.4 0.6 0.8 1
TE

Dimensionless reactor length

Fig. 8
EP
CC
A

46
PT
RI
SC
0.09

0.08

U
0.07
RWGS-R
N RWGS-MR
H2O mole fraction

0.06
A
0.05 0.035
M

0.04 0.025
D

0.03
0.015
0 0.02 0.04 0.06 0.08
TE

0.02

0.01
EP

0
0 0.2 0.4 0.6 0.8 1
Dimensionless reactor length
CC

9(a)
A

47
PT
RI
SC
U
0.25

N
A
0.2
M
CO mole fraction

0.15 RWGS-R
RWGS-MR
D
TE

0.1
EP

0.05
CC

0
0 0.2 0.4 0.6 0.8 1
Dimensionless reactor length
A

9(b)

Fig. 9

48
PT
RI
SC
U
0.25
N
A
0.2
M
Components mole fraction
in permeation side

0.15
TE

0.1
EP

H O
0.05 2
CC

H2

0
A

0 0.2 0.4 0.6 0.8 1


Dimensionless reactor length

Fig. 10

49
PT
RI
SC
U
N
A
350
M

CR 326
313
Scenario 1
300 Scenario 2
Production rate (ton/day)

250
TE

217
200
149
EP

150 124
88
CC

100

50
A

0
Water MeOH

Fig. 11

50
List of Tables:

Table 1: The operating conditions and reactor specifications of RWGS reactor

Table 2: The operating conditions and reactor specifications of RWGS membrane reactor

Table 3: The operating conditions and reactor specifications of methanol synthesis reactor

Table 4: The kinetic parameters of methanol synthesis reaction

PT
Table 5: Auxiliary correlations in the mathematical modeling

RI
Table 6: Comparison between model results and plant data of methanol synthesis reactor

Table 7: Optimization results of both scenarios

SC
U
N
A
M
D
TE
EP
CC
A

51
Table 1

Parameter Values
Particle density (kg/m3) 2628
Particle size (m) 6×6×10-3
Length of reactor (m) 7

PT
Diameter of reactor (m) 0.1292
Number of tubes 2962
Inlet pressure (bar) 1-30

RI
Inlet temperature (ºC) 100-450
Feed flow rate per tube (mol/s): 0.64
Feed composition (mole fraction):

SC
CO2 0.25
H2 0.75

U
N
A
M
D
TE
EP
CC
A

52
Table 2

Parameter Values
Length of reactor (m) 7
Inner tube diameter (m) 0.1292
Outer tube diameter (m) 0.25

PT
Number of tubes 2962
Sweep gas composition (mole fraction):
H2 0

RI
H2O 0
N2 1
Sweep gas inlet temperature (ºC) 27

SC
Sweep gas inlet pressure (bar) 1
Sweep gas flow rate per tube (mol/s): 0.1

U
Table 3
N
A
Parameter Values
Particle diameter (m) 5.47 × 10-3
M

Bed void fraction 0.39


Density of catalyst bed (kg/m3) 1140
Length of reactor (m) 7.022
3.8× 10-2
D

Diameter of reactor (m)


Number of tubes 2962
TE

Inlet pressure (bar) 76.98


Inlet temperature (K) 503
Feed flow rate (mol/s) 0.64
Feed composition (mole fraction)a ,b:
EP

CH3OH 0.005
CO2 0.094
CO 0.046
CC

H2O 0.0004
H2 0.659
N2 0.093
A

CH4 0.1026

a feed composition of conventional methanol synthesis reactor from natural gas


bfeed composition of methanol synthesis reactor in the indirect CO2 conversion process is
obtained in the present work after process modeling

53
Table 4

Rate constants [mol/s kg bar]

PT
B A B
k  A exp( )
RT
k1 (4.89  0.29) 107 63,000  300

RI
k2 (9.64  7.3)  1011 152,900  6800
k3 (1.09  0.07) 105 87,500  300

SC
Adsorption equilibrium constants
B A B
K  A exp( )
RT

U
KCO (2.16  0.44) 105 46,800  800
61,700  800
KCO 2
K H 2O
N
(7.05  1.39) 107
(6.37  2.88) 109 84,000  1400
A
K 1/H 22
M

Equilibrium constants
A A B
( )
K p  10 T  B
D

K p1 5139 12.621
K p2 3066 10.592
TE

K p3 -2073 -2.029
EP
CC
A

54
Table 5

PT
Parameter Equation
Component heat capacity 2 2
 C3 / T   C5 / T 
C p  C1  C2    C4 

RI

 sinh(C3 / T )   sinh(C5 / T ) 
Viscosity of reaction mixtures

SC
C1T C2

C C
1  3  42
T T

U
Binary gas diffusion
10 7 T 3 / 2 1 M i  1 M j
Dij 
N P(vci3 2  vcj3 2 )
A
Overall heat transfer coefficient
1 1 Ai ln( Do Di ) Ai 1
  
U hi 2LK w Ao ho
M

Heat transfer coefficient between the gas 0.407


 Cp  0.458  ud p
23
phase and reactor wall h 
    
C p   K   B  
D


TE
EP

Table 6
CC

Plant data Model Error %


Reactor Reactor Reactor
inlet outlet outlet
A

Feed flow rate (mol s-1) 0.565 0.51 0.5065 -0.68


Temperature (K) 503 528 523.08 -0.93
Composition (mol%):
CO2 3.45 2.18 2.17 -0.45
H2 79.55 75.71 75.51 -0.26
CO 4.66 1.44 1.09 -24.30
H2O 0.08 1.74 1.76 1.14

55
CH3OH 0.032 5.49 5.81 5.82
CH4 11.72 12.98 13.07 0.69

Table 7

PT
Decision variables Scenario 1 Scenario 2
nRWGS 0.76 0.5
Inlet temperature (˚C) 450 389
Inlet pressure (bar) 23 30

RI
SC
U
N
A
M
D
TE
EP
CC
A

56

You might also like