You are on page 1of 15

Journal of Natural Gas Science and Engineering 20 (2014) 132e146

Contents lists available at ScienceDirect

Journal of Natural Gas Science and Engineering


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

Syngas production in a novel methane dry reformer by utilizing of trireforming process for energy supplying: Modeling and simulation
Mehdi Farniaei a, Mohsen Abbasi b, Hamid Rahnama c, Mohammad Reza Rahimpour c, *,
Alireza Shariati c
a
b
c

Department of Chemical Engineering, Shiraz University of Technology, Shiraz 71555-313, Iran


Department of Chemical Engineering, School of Chemical and Petroleum Engineering, Persian Gulf University, Bushehr 75169, Iran
Chemical Engineering Department, School of Chemical and Petroleum Engineering, Shiraz University, Shiraz 71345, Iran

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 7 May 2014
Received in revised form
13 June 2014
Accepted 15 June 2014
Available online

In this study, tri-reforming process has been utilized as an energy source for driving highly endothermic
process of methane dry reforming process in a multi-tubular recuperative thermally coupled reactor
(TCTDR). 184 two-concentric-tubes have been proposed for this conguration. Outer tube sides of the
two-concentric-tubes have been considered for the tri-reforming reactions while dry reforming process
takes place in inner tube sides. Simulation results of co-current mode have been compared with corresponding predictions of thermally coupled tri- and steam reformer (TCTSR); in which the tri-reforming
process has been coupled with steam reforming of methane in same conditions. A mathematical heterogeneous model has been applied to simulate both dry and tri-reforming sides of the TCTDR. Results
showed that methane conversion at the output of dry and tri-reforming sides reached to 63% and 93%,
respectively. Also, molar ow rate of syngas at the output of DR side of TCTDR reached to 7464 kmol h1
in comparison to 3912 kmol h1 for SR side of TCTSR.
2014 Elsevier B.V. All rights reserved.

Keywords:
Syngas production
Dry reforming of methane
Tri-reforming of methane
Steam reforming of methane

1. Introduction
Synthesis gas (syngas), a fuel gas mixture consists of hydrogen
and carbon monoxide is important intermediate for production of
several materials such as; methanol, hydrogen, ammonia and
synthetic petroleum for use as a fuel through FischereTropsch
process (Halmann and Steinfeld, 2004; Stelmachowski and
Nowicki, 2003; Van der Drift and Boerrigter, 2006; Zahedi nezhad
et al., 2009).
One of the basic process to produce syngas with desired ratio of
H2/CO is the reforming of methane or natural gas. There are three
main reforming processes; steam reforming (SRM), carbon dioxide
reforming (CDR) and partial oxidation of methane (POM). Also,
recently a new process named as tri-reforming of methane (TR) has
been proposed by Song and Pan in 2001, for production of syngas
from ue gases (Arena et al., 1996; Choi et al., 2004; Lee et al.,
2003).

Dry reforming is the reaction of carbon dioxide with methane


that converting two major greenhouse gases with high global
warming potential into valuable chemicals. Produced syngas from
dry reforming reaction has low H2/CO ratio that is suitable for use in
FischereTropsch synthesis (Kang et al., 2007; Lee et al., 2003;
Nematollahi et al., 2011; Pino et al., 2011; Yin et al., 2007).
Dry reforming of methane reaction (DR):

CH4 CO2 42CO 2H2 ;

DH298 247:3 kJ mol1




TR is a new process that synergistically combines three main


catalytic reforming reactions; CO2 reforming (Eq. (1)), steam
reforming (Eqs. (2)e(4)) and oxidation of methane (Eqs. (5) and (6))
in a single reactor (Song and Pan, 2004).
Steam reforming of methane reactions (SR):

SR 1 : CH4 H2 O4CO 3H2 ;


SR 2 : CH4 2H2 O4CO2 4H2 ;

* Corresponding author. Tel.: 98 711 2303071; fax: 98 711 6287294.


E-mail address: rahimpor@shirazu.ac.ir (M.R. Rahimpour).
http://dx.doi.org/10.1016/j.jngse.2014.06.010
1875-5100/ 2014 Elsevier B.V. All rights reserved.

(1)

DH298 206:3 kJ mol1




(2)

DH298 164:9 kJ mol1

Wateregas shift reaction (WGSR):

(3)

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

DH298 41:1 kJ mol1




CO H2 O4CO2 H2 ;

(4)

Partial oxidation of methane (POM):

1
CH4 O2 4CO 2H2 ;
2

DH298 35:6 kJ mol1




(5)

Complete oxidation of methane (COM):

CH4 2O2 4CO2 2H2 O;

DH298 880 kJ mol1




(6)

WGSR is an important chemical reaction taking place during the


TR process which has a signicant impact on the whole process
(Fiaschi and Baldini, 2009). Syngas from TR process has the H2/CO
ratio about 1.5e2, suitable for production of other chemicals such
as methanol, dimethyl ether and liquid hydrocarbons (Jiang et al.,
2007).
TR concept allows effective utilization and conversion of CO2 in
ue gases (that contain CO2 H2O, O2 and N2) from power plants
without any pre-separation. On the other hand, because of existence of methane oxidation reaction, heat generated in the TR
reactor is as high as it can be used to drive another endothermic
reaction like dry reforming reaction (Arab Aboosadi et al.,
2011a,b).
By the concept of thermally coupled reactor the generated heat
from TR process can be transferred to dry reforming reaction and
thus both processes proceed simultaneously. Generally, in thermally coupled reactors, two reactions are coupled together in a
single reactor. One of the reactions is considered as heat source
(exothermic reaction) for another reaction (endothermic reaction).
Hence, endothermic reaction is driven by generated heat from
exothermic reaction in a reactor. Thermally coupled reactors
generally are classied into three main categorizes; direct coupling,
regenerative coupling and recuperative coupling. Much attention
has been paid to recuperative coupled reactors in much literature.
Recuperative coupled reactors consist of two vertical concentric

tubes for each of reactions separated by a wall. Generally, inner tube


side has been considered for exothermic reaction and outer tube
side is for exothermic one. Heat indirectly is transferred from outer
tube side to inner tube side through the wall (Rahimpour et al.,
2012).
In the present work, a recuperative coupled reactor has been
proposed in which TR process is considered as heat source for
highly endothermic dry reforming reaction. This reactor is consisted of two xed beds separated by a wall where heat is transferred across the wall from outer tube side to inner tube side of the
reactor. Thus, in this way both reactions are driven simultaneously
in single reactor.

2. Processes description
2.1. Conventional steam reformer (CSR)
Conventional steam reformer has been used for syngas production in methanol synthesis unit of Zagros Petrochemical
Company, Assaluyeh, Iran. CSR is consisted of a xed bed reactor
and a huge top red-furnace for supplying needed energy for SR
equal to 69 MWth in xed bed reactors. A conventional
auto-thermal reformer (CAR) has been applied at the outlet of
CSR for complete conversion of methane and adjustment of H2/
CO ratio that is suitable for methanol production (Rahnama
et al., 2014). Operational conditions of the CSR with catalyst
characteristics and feed compositions have been presented in
Table 1.

2.2. Thermally coupled tri- and dry reformer (TCTDR)


Fig. 1 presents a schematic diagram of thermally coupled triand dry reformer reactor (TCTDR). 184 two concentric tubes has
been considered for the reactor conguration in which tri- and dry
reforming processes occur in outer and inner tube sides of the
reactor, respectively.

Table 1
Operating conditions of steam reforming side of TCTSR (same as the conventional
steam reformer (CSR)).
Parameter
Total feed gas ow (kmol h

Value
1

9129.6

Feed composition (mol %)


CO2
CO
H2
CH4
N2
H2O

1.72
0.02
5.89
32.59
1.52
58.26

Steam to methane ratio

1.79

Inlet temperature (K)


Design temperature (K)
Inlet pressure (kPa)
Design pressure (kPa)
Inside diameter (mm)
Heated length (m)
Number of tubes
Catalyst name
Catalyst volume (m3)
Catalyst shape
Particle size (mm)
Void fraction (e)
Heat load on tube (100% of design case)
(MJ m2 h1) (based on tube ID)
Reformer duty (100% of design case)
[GJ h1] net

793.15
1063.15
4000
4100
125
12
184
NiOeMg/CeeZrO2/Al2O3
27.8
10-HOLE rings
19  16
0.4
287.6

133

248.2
Fig. 1. A schematic diagram of TCTDR.

134

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146
Table 2
Optimized parameters for tri-reformer reactor (Arab Aboosadi et al.,
2011a,b).
Inlet temperature (K)

1100

Composition (mol %)
CO2
CO
H2
CH4
O2
N2
H2O

24.81
0.01
1.53
18.7
8.78
0.01
46.18

Or reactant ratios
Steam/methane ratio
Oxygen/methane ratio
Carbon dioxide/methane ratio

2.46
0.47
1.30

2.3. Thermally coupled tri- and steam reformer (TCTSR)


Schematic diagram of thermally coupled tri- and steam
reformer (TCSTR) is similar to TCTDR.
This multi-tubular reactor has 184 two-concentric-tubes and
endothermic steam reforming of methane occurs in the inner tube
side, while methane tri-reforming takes place in the outer tube
side. Tri-reforming reaction supply sufcient energy for driving
itself and also, steam reforming reaction in inner tube side. Specications of steam reforming side have been considered same as the
CSR reactor. Ni-based catalysts are loaded in vertical tubes of steam
reforming side. Natural gas and steam mixed together and entered
to the inner tube side (steam reforming side). Generated heat from
tri-reforming side transfers to the inner tube side and reaction take
places.
Effective length of TCTSR is 12 m similar to TCTSR based on
specication of CSR in Zagros Petrochemical Company, Assaluyeh,
Iran (see Table 1).
3. Kinetic of reactions

Table 3
Simulation conditions of tri-reforming (TR) side of TCTDR reactor (Arab
Aboosadi et al., 2011a,b).
Total feed gas ow (kmol h1)
Methane feed rate (kmol h1)
Inlet pressure (bar)
Catalyst shape
Particle size (mm)
Shell inner diameter (m)

28,115.4
9264.4
20
10-HOLE rings
19  16
2

3.1. Dry reforming reactions


During production of synthesis gas by dry reforming, beside dry
reforming reaction (Eq. (1)), reverse wateregas shift reaction
(reverse of Eq. (4)) occurs, too. Kinetic of reaction for Eqs. (1) and (4)
over Rh/Al2O3 catalyst is considered based on Richardson and
Paripatyadar model (Richardson and Paripatyadar, 1990), as
follows:

"
2.2.1. Dry reforming side (inner tube side)
Dry reforming side of the reactor has been loaded with Rh/Al2O3
catalysts and size of reactor is similar to CSR as presented in Table 1.
Rh/Al2O3 is an effective catalyst for high dry reforming reaction
rate with CO2/CH4 ratio near 1 in feed. Also, in dry reforming experiments, no signs of coke formation on the catalyst surface have
been found (Richardson and Paripatyadar, 1990).

rDR kDR


rRWGS 
2.2.2. Tri-reforming side (outer tube side)
Pre-reformed gas is mixed with natural gas stream and proper
ratio of steam, CO2 and O2. Then pre-reformed gas is sent through
tri-reforming side that is loaded with NiOeMg/CeeZrO2/Al2O3
catalysts. This type of catalyst reduces coke formation on the
reactor wall and surface of the catalyst (Cho et al., 2009). Size of the
catalysts is same as the conventional steam reformer. The optimized inlet parameters of tri-reformer reactor have been considered for inlet parameters of tri-reforming side of the reactor which
are listed in Table 2 (Arab Aboosadi et al., 2011a,b).
Table 3 shows other simulation conditions for tri-reforming side
of the reactor.

KCO2 KCH4 PCO2 PCH4


2
1 KCO2 PCO2 KCH4 PCH4

2 !#
PCO PH2
1
KDR PCO2 PCH4
kRWGS KCO2 KH2 PCO2 PH2

1 KCO2 PCO2 KH2 PH2

"
2


2 #
PCO PH2 O
1
KRWGS PCO2 PH2

(7)

(8)

The Arrhenius kinetic parameters, constants of reaction equilibrium and adsorption equilibrium constants are summarized in
Table 4.
3.2. Tri-reforming side reactions
Eqs. (2)e(4) and also complete oxidation of methane (Eq. (6))
have been considered to describe tri-reforming process. Among trireforming reactions, dry reforming reaction (Eq. (1)) is a dependent
reaction as it can be written as SRM reaction minus WGS reaction.
Hence, considering kinetic model for Eqs. (2) and (4) is enough and

Table 4
Arrhenius kinetic parameters, constants of reaction equilibrium and adsorption equilibrium constants for DR and RWGS reactions.

Rate constant of reaction (mol

g1
cat

1

Equilibrium constants of reaction


Equilibrium constants of adsorption (Pa1)

DR

RWGS



kDR 1290 exp 102;065
RT


KDR exp 420:3T773
RT


KCO2 2:64  103 exp 37;641
RT


KCH4 2:63  103 exp 40;684
RT



kRWGS 1:856  105 exp 73;105
RT


KRWGS exp 4400
T  4:036


KCO2 5:6955  106 exp 9262
RT


KH2 1:4705  105 exp 6025
RT

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

no need to consider kinetic model for CO2 reforming reaction (Eq.


(1)). The kinetic model of Xu and Froment over Ni-based catalysts is
used for Eqs. (2)e(4). This model has been extensively tested under
lab scale and is more general for steam reforming reactions. Xu and
Froment kinetic model for reactions (2)e(4) is as follows (Xiu et al.,
2002; Xu and Froment, 1989):

p3H2 pCO

k1
R1 2:5
pH2
R2

1
pCH4 pH2 O 
 2
KI
f
!
p4H pCO2
1
pCH4 p2H2 O  2
 2
KII
f

k2
p3:5
H2

k
R3 3
pH2

pH pCO2
pCO pH2 O  2
KIII

(9)

k4a pCH4 pO2


1

KOC2 pO2

1
 2
f

(11)
pH2 O
pH2

(12)

k4b pCH4 pO2


1

Components

Koi (bar1)

DHi (J mol1)

CH4
CO
H2
H2O
CH4 (combustion)
O2 (combustion)

6.65  104
8.23  105
6.12  109
1.77  105

38,280
70,650
82,900
88,680

C p
KCH
CH4
4

KOC2 pO2


(13)

Table 5 shows reaction equilibrium constant and Arrhenius


parameters. Van't Hoff parameters for species adsorption are listed
in Table 6. The consumption or formation rate of species i, ri
(mol kg1 s1), is determined by summing up the reaction rates of
species i (Rj (mol kg1 s1)). To account intra-particle transport
limitations, effectiveness factor, hi, are used (De Groote and
Froment, 1996; Gosiewski et al., 1999). Therefore the reaction rate
of each species is as below:

rCH4 h1 R1  h2 R2  h4 R4

(13-a)

rO2 2h4 R4

(13-b)

rCO2 h2 R2 h3 R3 h4 R4

(13-c)

rH2 O h1 R1  2h2 R2  h3 R3 2h4 R4

(13-d)

rH2 3h1 R1 4h2 R2 h3 R3

(13-e)

rCO h1 R1  h3 R3

(13-f)

Table 5
Reaction equilibrium constants and Arrhenius kinetic parameters for TRM reactions.
1
1
Reaction, j Equilibrium constant, Kj
koj (mol kg1
)
cat s ) Ej (J mol


15
0.5
1.17

10
bar
240,100
1
30:114 bar2
KI exp 26;830
Ts

KII KI $KIII bar2




KIII exp 4400
Ts  4:036

4
kj koj  expEj =RT.

DHCi (J mol1)

1.26  101
7.78  107

27,300
92,800

where h1 0.07, h2 0.06, h3 0.7, h4 0.05 (De Groote and


Froment, 1996).
4. Mathematical model
A one dimensional heterogeneous catalytic reaction model has
been developed in order to determine the concentration and
temperature distribution inside both sides of the reactor. This
model will accounts heat and mass transfer resistance along the
reactor. Some assumptions are considered to model both sides of
the reactor:





The reactor operates at steady state conditions.


Gas phase is ideal.
Axial diffusions of mass and heat are negligible.
One dimensional plug ow pattern is considered in both sides of
the reactor.
 Bed porosity in axial and radial directions is constant.
 Heat loss to surrounding is neglected.
Mass and energy balance equations are obtained by considering
a differential element along the axial direction of the reactor.
4.1. Balance equation for solid phase
The mass and energy balance equations for solid phase have
been developed for both sides of the reactor as follows:

kgi yi  yis hri rB a 0

av hf T  Ts rB a

N
X
i1

KCoi (bar1)

(10)

The kinetic model of Trimm and Lam (1980) is used for methane
combustion (Eq. (6)) as a rigorous study, but it was driven over supported Pt-based catalyst. Hence, the model adsorption parameters are
adjusted for Ni-based catalysts as below (De Smet et al., 2001):

C p
KCH
CH4
4

Table 6
Van't Hoff parameters for species adsorption.

Ki Koi  expDH=RT.
C  expDH C =RT.
KiC Koi
i

f 1 KCO pCO KH2 pH2 KCH4 pCH4 KH2 O

R4 

135

2.83  1014 bar0.5 243,900


5.43  105 bar1

67,130

8.11  105 bar2


6.82  105 bar2

86,000
86,000



hri DHf i 0

(14)

(15)

where Ts and yis are temperature and the solid phase mole fraction
of component i inside of reactor, respectively and h is effectiveness
factor (the ratio of the reaction rate observed to the real rate of
reaction), which is obtained from a dusty gas model calculations
(Graaf et al., 1990). Surface areas per volume of catalyst pellet are
shown by av (m2 m3).
4.2. Balance equation for uid phase
Mass and energy balance equations for both sides have been
developed as follows:

Ft vyi
av ct kgi yis  yi 0

Ac vz

(16)

1 g vFt T
av hf Ts  T 0
 Cp
Ac
vz

(17)

136

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

Table 7
Auxiliary correlations.
Parameter

Equation

Reference

Component heat capacity


Mixture heat capacity
Viscosity of reaction mixtures
Mixture thermal conductivity
Mass transfer coefcient between gas and solid phases

Cp a bT cT 2 dT 3
Based on local compositions
Based on local compositions
Based on local compositions

Barbieri and Di Maio (1997)

ug  103
kgi 1:17Re0:45 Sc0:67
i
2Rp ug r
Re
m
m
Sci
rDim  104
1yi
Dim P
yi

Van Ness et al. (2001)


Cussler (1997)

isj Dij

Dij

107 T 3=2

3=2

1=Mi 1=Mj

o =Di
h1 Ai lnD
AAoi h1
2pLKw
o
i
 2=3

0:407
Cp m
h
0:458 rudp

m
B
Cp rm
k

Overall heat transfer coefcient

1
U

Heat transfer coefcient between gas phase and reactor wall

Reid et al. (1977)

3=2

Pvci vcj 2

Van Ness et al. (2001)


Smith (1980)

where T and yi are temperature and the uid phase mole fraction of
component i inside of the reactor, respectively. kgi is the mass
transfer coefcient between gas and solid phase. First and second
terms in Eq. (17) show heat transferred by convection and the heat
exchanged between uid phase and solid particle, respectively.

4.3. Pressure drop


The Ergun momentum equation is used to consider the pressure
drop variations along the reactor axis:
2

1  ug r
1  2 mug
dp
150
1:75
dz
3 d p
3 d2p

(18)

where the pressure drop is in Pa.

4.4. Boundary conditions


Temperature, pressure and gas compositions are known at the
entrance of the reactor. Hence, the following boundary conditions
are applied at the inlet of reactor:

z 0;

yi yi0 ;

T T0 ;

P P0

(19)

For analyzing performance of TCTDR and TCTSR, syngas quality


(H2/CO) and CH4 conversion is calculated by following equations:

CH4 conversion

FCH4 ;in  FCH4 ;out


FCH4 ;in

(20)

Table 8
Comparison between model prediction and plant data.
Parameter

Plant data

CSR

Methane conversion %

26.5

26

Composition (mol %)
CO2
CO
H2
CH4
N2
H2O

5.71
3.15
31.39
20.41
1.29
38.05

5.72
3.19
31.53
20.33
1.30
37.94

Fig. 2. Variations in mole fraction of all components along (a) tri-reforming (TR) side
and (b) dry reforming (DR) side of the TCTDR reactor.

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

Syngas quality

FH2
FCO

(21)

4.5. Auxiliary correlations


To complete the simulation, some auxiliary correlations must be
applied into the developed model equations. The correlation estimations for heat and mass transfer between two phases and also,
overall heat transfer coefcient between wall and gas phase in the
inner tube side must be considered. Table 7 shows applied correlations for physical properties, mass and heat transfer coefcient.
The heat transfer coefcient between gas phase and reactor wall is
applicable for heat transfer coefcient between gas bulk phase and
solid phase (hf).
5. Numerical solution
The developed model contains a set of differential equations by
using the backward nite difference approximation for solving the
model; the differential equations alter to a non-linear algebraic set

Fig. 3. Variations in reaction rates along (a) TR side, and (b) DR side of TCTDR reactor.

137

of equations. Then length of the reactor is divided into 100


separate sections. Non-linear algebraic equations of each section
are solved by GausseNewton method in MATLAB programming
environment. Validation of modeling results has been presented in
Table 8. Results identify that simulation results at the output of
CSR are similar to plant data in Zagros Petrochemical Company,
Assaluyeh, Iran.
6. Results and discussion
Fig. 2(a) and (b) represents variations of all components mole
fractions along tri- and dry reforming sides of the TCTDR, respectively. At the entrance of TR side, all oxygen is rapidly consumed by
COM reaction and CO2 is produced at rst but then CO2 mole
fraction is decreased because of consuming by reforming reactions.
Behavior of prole of H2O mole fraction is almost similar to CO2
mole fraction to length of 2.8 m of the TR side of reactor but after
that RWGS reaction controls process conditions and causes
increasing H2O mole fraction at the rest of the reactor. Proles of
component mole fractions along dry reforming (DR) side, Fig. 2(b),
can be explained via stoichiometric coefcient components in dryreforming process. Methane and carbon dioxide as reactants have

Fig. 4. (a) Changes in temperature of TCTDR along reactor axis and (b) variation of heat
generation in TR side, heat consumption in DR side and heat transfer between sides of
TCTDR along reactor axis.

138

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

Fig. 6. Molar ow rate of syngas along TR side of TCTDR and TCTSR reactors.

Fig. 5. Comparison between mole fraction proles of (a) CO2 and H2, (b) H2O and (c)
CH4 and CO along TR side of TCTDR and TSTSR reactors.
Fig. 7. Variations of (a) CH4 conversion and (b) H2 yield along TR side of TCTDR and
TCTSR reactors.

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

139

Fig. 8. Changes of (a) temperature and (b) generated heat along TR side of TCTDR and
TCTSR reactors.

Fig. 9. Comparison between mole fraction of (a) H2 and CH4, (b) H2O, and (c) CO and
CO2 along DR side of TCTDR and SR side of TCTSR.

140

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

Fig. 10. Changes of (a) temperature and (b) consumed heat along DR side of TCTDR and
SR side of TCTSR reactors.

Fig. 12. Variations in (a) syngas quality along DR and TR sides of TCTDR, (b) syngas
quality along SR and TR sides of TCTSR and (c) CH4 conversion along DR side of TCTDR
and SR side of TCTSR reactors.

Fig. 11. Molar ow rate of syngas along DR side of TCTDR and SR side TCTSR reactors.

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

same stoichiometric coefcients and also both products of dry


reforming reaction; carbon monoxide and hydrogen have stoichiometric coefcient of 2. Small amount of H2O is produced by
RWGS.
Fig. 3(a) and (b) shows how reaction rates change along the triand dry reforming side of reactor, respectively. Fig. 3(a) reveals
that complete oxidation of methane controls conditions of TR
process kinetically at the entrance of TR side. When most of oxygen is consumed by COM reaction (see Fig. 2(a)); rate of reaction
reaches to its equilibrium value, so endothermic reactions of TR
process control conditions (equilibrium controlling). Reaction rate
of WGS is negative because of high reaction rate of RWGS reaction
that is due to high temperature and concentration of CO2 in feed
gas.
Fig. 3(b) shows that rate of DR and RWGS reaction is high at the
entrance of DR side because of high temperature and concentration
of CH4 and CO2 at rst. Then reaction rates decreases because of
temperature reduction at the rest of DR side. Reaction rate of RWGS
reaches near zero at the entrance of DR side of TCTDR reactor. Fig. 4
demonstrates temperature proles in TR and DR sides of the
reactor. Also, generated heat from TR side, consumed heat by DR
side and transferred heat between both sides are presented in
Fig. 4(b), for more explanation about thermal behavior of the
reactor. Temperature increases along TR side at rst, because a large
amount of heat is generated by methane combustion reaction (see
Fig. 4(b)) at the entrance of TR side. This generated heat is spent for
endothermic reactions in TR side and temperature along rest of TR
side decreases gradually.
6.1. Comparison between tri-reforming side of two different
reactors
This section discus about comparison between two thermally
coupled tri- and dry reforming (TCTDR) reactor and second is
thermally coupled tri- and steam reforming (TCTSR) reactor; in
which tri-reforming process is coupled with steam reforming
process instead of dry reforming process.
Fig. 5(a) and (b), represents comparison between proles of
mole fraction of components along TR side of TCTDR and TSTSR
reactors. As shown in these gures, there are no signicant difference between behavior of component mole fraction in TR side of
reactors; because of same feed and conditions of TR sides.
As shown in Fig. 6, molar ow rate of syngas along TR side of
TCTSR is higher than TCTDR one in the half part of TR side,
although its nal value is lower. Flow rate of syngas along TR side
of TCTSR increases to 56,860 kmol h1 at length of 2.6 m then after
that begin to decrease and reaches to 38,360 kmol h1 at the end
of the reactor (12 m length) while molar ow rate of syngas along
TR side of TCTDR increases to 45,440 kmol h1 at the end of the
reactor.
Fig. 7 shows variations of CH4 conversion along TR side of TCTDR
and TCTSR reactors. CH4 conversion proles along TR side of both
reactors are similar and nal values at the end of the reactor are
near together; these are due to same feed conditions of TR side of
both reactors.
Fig. 8(a) and (b) represents thermal behavior of TR side of both
reactors. Since all components and conditions in TR side of TCTDR
and TCTSR reactors are considered same, the amount of oxygen that

141

is reacted with methane is same, too. Hence prole of heat generation from TR side of TCTDR is overlapped to the TCTSR one.
Temperature prole along TR side of TCTDR is lower than TCTSR
one at the second part of reactor.
6.2. Comparison between DR side of TCTDR and SR side of TCTSR
reactors
Fig. 9(aec) presents a comparison between proles of component mole fractions along steam reforming (SR) side of TCTSR and
dry reforming (DR) side of TCTDR reactors. High effect of WGS reaction on SR process in second part of SR side of TCTSR reactor
(after reaching to equilibrium) led to a lower prole of H2 in
comparison with H2 prole in DR side of TCTDR reactor, as shown in
Fig. 9(a).
Since H2O is one of the feed components of SR side of
TCTSR, prole of H2O mole fractions higher than H2O mole
fraction in DR side of TCTDR reactor (see Fig. 9(b)). Small
amount of H2O in DR side of TCTDR reactor is produced by
reverse WGS.
Variations of CO and CO2 mole fraction along SR side of
TCTSR and DR side of TCTDR are illustrated in Fig. 9(c). CO2 in
DR side of TCTDR is reactant while in SR side of TCTSR is
product, so CO2 decreases along DR side and increases along SR
side.
Thermal behavior of both DR side of TCTDR and SR side of TCTSR
reactors are compared in Fig. 10(a) and (b). Heat consumption along
DR side of TCTDR is higher than transferred heat so temperature
decreases along DR side.
Fig. 11 shows syngas molar ow rate along DR and SR side of
TCTDR and TCTSR reactors, respectively. Molar ow rate of syngas
reaches to 7464 kmol h1 at the output of DR side of TCTDR while
this value is 3912 kmol h1 for SR side of TCTSR reactor.
Fig. 12(a) and (b) represents proles of syngas quality (H2/
CO) along sides of both reactors. Results identify that syngas
quality is equal to 9.19 and 1.12 at the outlet of SR and TR sides
of TCTSR respectively. Also, syngas is produced with a quality of
1.08 and 0.958 in TR and DR sides of TCTDR respectively.
Therefore, by considering molar ow rate of syngas in each side
for determination of reactors performance, TCTDR and TCTSR
can produce total syngas with a quality of 1.063 and 1.848
respectively.
CH4 is more converted by DR side of TCTDR in comparison with
SR side of TCTSR as shown in Fig. 12(c). CH4 conversion reached to
63% and 26% at the end of endothermic side of TCTDR and TCTSR
reactors, respectively.
Finally, the pros and cons of the TCTDR can be addressed as
follows to be overcome for this to become a more widely adopted
method.
Production of two types of syngas and reduction of energy
consumption is advantages of TCTSR and TCTDR. In CSR, huge
amount of energy is needed. For preparation of this energy
methane should be consumed in red-furnace for energy production with CO2 generation. But in TCTSR and TCTDR, energy is prepared from TR reactions and red-furnace is removed. Elimination
of a low performance red-furnace and replacing it with a high
performance reactor causes a reduction of full consumption with
production of a new type of syngas.

142

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

One of the problems with dry reforming reaction is formation of


coke on the active surface and leading to deactivation of the
catalyst rapidly. Carbon deposition is dependent on temperature,
type of catalyst and existence of oxygen, steam, etc. in process. By
adjusting one or all of these parameters and choosing an appropriate catalyst carbon deposition can be signicantly eliminated
during reaction.
Molar ow rate of the feed gases to TR side should be sufcient
for preparation of enough heat for DR and SR endothermic reactions. Coke formation on the catalyst and reactor wall is eliminated in TR side because of applying some operating conditions of
the reactor such as high temperature, high steam to carbon ratio in
feed composition, appropriate Ni-based catalysts and presence of
steam and oxygen in the feed.
Of course, economic feasibility such as capital and operating cost
of the process is necessary in order to consider commercialization
of the proposed conguration.
7. Conclusions
Tri-reforming process was supplied for providing necessary heat
of endothermic dry reforming process plus production of two types
of syngas by different quality in a multi-tubular thermally coupled
reactor (TCTDR). 184 two-concentric-tubes were simulated
numerically by a one dimensional heterogeneous model and results
of each side compared with thermally coupled steam and trireformer (TCSTR) reactor. Results identied that CH4 conversion at
the output of dry reforming (DR) side of TCTDR reached to 63% that
is very higher than steam reforming (SR) side of TCTSR reactor with
26%. Also, syngas quality of DR side in TCTDR is lower than SR side
of TCTSR (0.958 vs. 9.19). Molar ow rate of syngas at the output of
DR side in TCTDR reactor reached to 7464 kmol h1 while this value
was 3912 kmol h1 for SR side in TCTSR. CH4 conversion at the
output of TR side of TCTDR reached to 93%. Syngas molar ow rate
at the output of TR side of TCTDR reached to 5.544  104 kmol h1.
Some advantages of TCTDR reactor are: both sides produced syngas
with different quality, needed energy for highly dry reforming
process was supplied by another process (tri-reforming) instead of
huge furnaces.
Nomenclature
av
Ac
Cp
dp
Di
Dij
Dim
Ft
Fi
hf
hi
ho
ki

specic surface area of catalyst pellet, m2 m3


cross section area of each tube, m2
specic heat at constant pressure, J mol1
particle diameter, m
tube inside diameter, m
binary diffusion coefcient of component i in j, m2 s1
diffusion coefcient of component i in the mixture, m2 s1
total molar ow rate, mol s1
ow rate of component i, mol s1
gasesolid heat transfer coefcient, W m2 K1
heat transfer coefcient between uid phase and reactor
wall in exothermic side, W m2 K1
heat transfer coefcient between uid phase and reactor
wall in endothermic side, W m2 K1
rate constant of reaction i, mol kg1 s1

k4a
k4b
Kw
L
Mi
N
P
pi
ri
R1
R2
R3
R4
R
Rp
Re
Sci
T
u
ug
U
vci
yi
z

rst reaction rate constant for the fourth rate equation,


mol kg1 s1
second reaction rate constant for the fourth rate equation,
mol kg1 s1
thermal conductivity of reactor
wall, W m1 K1
reactor length, m
molecular weight of component i, g mol1
number of components, e
total pressure, bar
partial pressure of component i, Pa
reaction rate of component i, mol kg1 s1
rst rate of reaction for steam reforming of methane,
mol kg1 s1
second rate of reaction for steam reforming of methane,
mol kg1 s1
rate of reversed wateregas shift reaction, mol kg1 s1
rate of complete oxidation of methane, mol kg1 s1
universal gas constant, J mol1 K1
particle radius, m
Reynolds number, e
Schmidt number of component, e
temperature, K
supercial velocity of uid phase, m s1
linear velocity of uid phase, m s1
overall heat transfer coefcient between
exothermic and endothermic sides, W m2 K1
critical volume of component i, cm3 mol1
mole fraction of component i, mol mol1
axial reactor coordinate, m

Greek letters
a
activity of catalyst (where a 1
for fresh catalyst)
DHf,i
enthalpy of formation of component i, J mol1
B
void fraction of catalytic bed, e
h
effectiveness factor used for the intra-particle transport
limitation, e
m
viscosity of uid phase, kg m1 s1
r
density of uid phase, kg m3
rB
density of catalytic bed, kg m3
Superscripts
g
in bulk gas phase
s
at surface catalyst
Subscripts
0
inlet conditions
i
chemical species
j
reactor side
Appendix A
The following source code has been applied for solving nonlinear algebraic equations by GausseNewton method in MATLAB
version:

144

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

145

146

M. Farniaei et al. / Journal of Natural Gas Science and Engineering 20 (2014) 132e146

References
Arab Aboosadi, Z., Rahimpour, M., Jahanmiri, A., 2011a. A novel integrated thermally
coupled conguration for methane-steam reforming and hydrogenation of
nitrobenzene to aniline. Int. J. Hydrogen Energy 4, 2960e2968.
Arab Aboosadi, Z., Jahanmiri, A., Rahimpour, M., 2011b. Optimization of tri-reformer
reactor to produce synthesis gas for methanol production using differential
evolution (DE) method. Appl. Energy 8, 2691e2701.
Arena, F., Frusteri, F., Parmaliana, A., Plyasova, L., Shmakov, A., 1996. Effect of
calcination on the structure of Ni/MgO catalyst: an X-ray diffraction study.
J. Chem. Soc. Faraday Trans. 3, 469e471.
Barbieri, G., Di Maio, F.P., 1997. Simulation of the methane steam reforming process
in a catalytic Pd-membrane reactor. Ind. Eng. Chem. Res. 6, 2121e2127.
Cho, W., Song, T., Mitsos, A., McKinnon, J.T., Ko, G.H., Tolsma, J.E., Denholm, D.,
Park, T., 2009. Optimal design and operation of a natural gas tri-reforming
reactor for DME synthesis. Catal. Today 4, 261e267.
Choi, J.-S., Kwon, H.-H., Lim, T.-H., Hong, S.-A., Lee, H.-I., 2004. Development of
nickel catalyst supported on MgOeTiO2 composite oxide for DIR-MCFC. Catal.
Today 93, 553e560.
Cussler, E.L., 1997. Diffusion: Mass Transfer in Fluid Systems. Cambridge University
Press, Cambridge.
De Groote, A.M., Froment, G.F., 1996. Simulation of the catalytic partial oxidation of
methane to synthesis gas. Appl. Catal. A Gen. 2, 245e264.
De Smet, C., De Croon, M., Berger, R., Marin, G., Schouten, J., 2001. Design of adiabatic xed-bed reactors for the partial oxidation of methane to synthesis gas.
Application to production of methanol and hydrogen-for-fuel-cells. Chem. Eng.
Sci. 16, 4849e4861.
Fiaschi, D., Baldini, A., 2009. Joining semi-closed gas turbine cycle and trireforming: SCGT-TRIREF as a proposal for low CO2 emissions powerplants.
Energy Convers. Manag. 8, 2083e2097.
Gosiewski, K., Bartmann, U., Moszczynski, M., Mleczko, L., 1999. Effect of the
intraparticle mass transport limitations on temperature proles and catalytic
performance of the reverse-ow reactor for the partial oxidation of methane to
synthesis gas. Chem. Eng. Sci., 4589e4602.
Graaf, G., Scholtens, H., Stamhuis, E., Beenackers, A., 1990. Intra-particle diffusion
limitations in low-pressure methanol synthesis. Chem. Eng. Sci. 4, 773e783.
Halmann, M., Steinfeld, A., July, 2004. Methanol, hydrogen, or ammonia production
by tri-reforming of ue gases from coal- and gas-red power stations. In:
Proceedings of the ECOS2004 Conference, pp. 7e9.
Jiang, H., Li, H., Zhang, Y., 2007. Tri-reforming of methane to syngas over Ni/Al2O3
dthermal distribution in the catalyst bed. J. Fuel Chem. Technol. 1, 72e78.
Kang, J.S., Kim, D.H., Lee, S.D., Hong, S.I., Moon, D.J., 2007. Nickel-based trireforming catalyst for the production of synthesis gas. Appl. Catal. A Gen. 1,
153e158.

Lee, S.-H., Cho, W., Ju, W.-S., Cho, B.-H., Lee, Y.-C., Baek, Y.-S., 2003. Tri-reforming of
CH4 using CO2 for production of synthesis gas to dimethyl ether. Catal. Today 1,
133e137.
Nematollahi, B., Rezaei, M., Khajenoori, M., 2011. Combined dry reforming and
partial oxidation of methane to synthesis gas on noble metal catalysts. Int. J.
Hydrogen Energy 4, 2969e2978.
, M., Recupero, V., 2011. Hydrogen production by
Pino, L., Vita, A., Cipit, F., Lagana
methane tri-reforming process over Nieceria catalysts: effect of La-doping.
Appl. Catal. B 1, 64e73.
Rahimpour, M., Dehnavi, M., Allahgholipour, F., Iranshahi, D., Jokar, S., 2012.
Assessment and comparison of different catalytic coupling exothermic and
endothermic reactions: a review. Appl. Energy 99, 496e512.
Rahnama, H., Farniaei, M., Abbasi, M., Rahimpour, M.R., 25 July 2014. Modeling of
synthesis gas and hydrogen production in a thermally coupling of steam and
tri-reforming of methane with membranes. J. Ind. Eng. Chem. 20 (4),
1779e1792. http://dx.doi.org/10.1016/j.jiec.2013.08.032.
Reid, R.C., Sherwood, T.K., Prausnitz, J., 1977. The Properties of Gases and Liquids,
third ed. McGraw-Hill, New York.
Richardson, J.T., Paripatyadar, S.A., 1990. Carbon dioxide reforming of methane with
supported rhodium. Appl. Catal. 1, 293e309.
Smith, J.M., 1980. Chemical Engineering Kinetics. McGraw-Hill, New York.
Song, C., Pan, W., 2004. Tri-reforming of methane: a novel concept for catalytic
production of industrially useful synthesis gas with desired H2/CO ratios. Catal.
Today 4, 463e484.
Stelmachowski, M., Nowicki, L., 2003. Fuel from the synthesis gasdthe role of
process engineering. Appl. Energy 1, 85e93.
Trimm, D.L., Lam, C.W., 1980. The combustion of methane on platinumealumina
bre catalystsdI: kinetics and mechanism. Chem. Eng. Sci. 6, 1405e1413.
Van der Drift, A., Boerrigter, H., 2006. Synthesis Gas from Biomass for Fuels and
Chemicals. ECN Biomass, Coal and Environmental Research.
Van Ness, H., Smith, J., Abbott, M., 2001. Introduction to Chemical Engineering
Thermodynamics. McGraw-Hill, Crawfordsville.
Xiu, G., Li, P., Rodrigues, A.E., 2002. Sorption-enhanced reaction process with
reactive regeneration. Chem. Eng. Sci. 18, 3893e3908.
Xu, J., Froment, G.F., 1989. Methane steam reforming, methanation and wateregas
shift: I. Intrinsic kinetics. AIChE J. 1, 88e96.
Yin, L., Wang, S., Lu, H., Ding, J., Mosto, R., Hao, Z., 2007. Simulation of effect of
catalyst particle cluster on dry methane reforming in circulating uidized beds.
Chem. Eng. J. 1, 123e134.
Zahedi nezhad, M., Rowshanzamir, S., Eikani, M., 2009. Autothermal reforming of
methane to synthesis gas: modeling and simulation. Int. J. Hydrogen Energy 3,
1292e1300.

You might also like