You are on page 1of 11

International Journal of Heat and Mass Transfer 165 (2021) 120635

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Numerical study of sorption-enhanced methane steam reforming over


Ni/Al2 O3 catalyst in a fixed-bed reactor
Amira Neni a, Yacine Benguerba a,∗, Marco Balsamo b, Alessandro Erto c, Barbara Ernst d,
Djafer Benachour a
a
Laboratoire des Matériaux Polymères Multiphasiques (LMPMP), Ferhat ABBAS University of Setif-1, Sétif, Algeria
b
Dipartimento di Scienze Chimiche, Università degli Studi di Napoli Federico II, Complesso Universitario di Monte Sant’Angelo, 80126 Napoli, Italy
c
Dipartimento di Ingegneria Chimica, dei Materiali e della Produzione Industriale, Università degli Studi di Napoli Federico II, P.leTecchio, 80, 80125 Napoli,
Italy
d
Université de Strasbourg, CNRS, IPHC UMR 7178, Laboratoire de Reconnaissance et Procédés de Séparation Moléculaire (RePSeM), ECPM 25 rue Becquerel
F-67000 Strasbourg, France

a r t i c l e i n f o a b s t r a c t

Article history: The present work deals with the Sorption-Enhanced Methane Steam Reforming (SE-MSR), an interesting
Received 12 June 2020 and energy-efficient hydrogen production route with in situ CO2 capture. A computational fluid dynamics
Revised 17 October 2020
(CFD) model for an industrial-scale fixed-bed reactor, with Ni/Al2 O3 as catalyst and CaO as an adsorbent
Accepted 21 October 2020
for CO2 capture, is developed taken into consideration also the coke deposition. Temperature is shown
to be the key parameter of the SE-MSR chemical process at large scales. H2 production is constant and
Keywords: maximum until the saturation of CaO sorbent occurs, after which the concentrations of all the other com-
Sorption-enhanced reaction pounds start to vary, and the efficiency of the process begins to drop. When the exothermic carbonation
Methane steam reforming reaction stops, an alteration of the thermal regimes is observed. The absence of the contribution of the
Hydrogen
exothermic carbonation reaction results in a decrease of the temperature, which in turn determines a
CFD
lower conversion of CH4 and H2 O, according to the endothermic reforming reactions.
Catalytic reactor
Coke deposition The maximum H2 outlet mole fraction (dry basis) is 0.8, and it occurs in the presence of CO2 sorption;
the value drops to 0.42 once the adsorbent reaches its maximum conversion degree. The molar selectivity
in hydrogen relative to the quantity of CH4 fed to the reactor is of the order of 1.75 (with CO2 -capture)
and 0.8 (without CO2 capture).
The molar fluxes obtained and the kinetics of the system model show the excellent choice of the operat-
ing conditions of the catalyst to produce a large quantity of hydrogen as well as of the adsorbent, which
eliminates the CO2 responsible of coke deposition.
© 2020 Elsevier Ltd. All rights reserved.

1. Introduction mary industrial process for H2 production. To mitigate MSR-related


CO2 emissions and to increase hydrogen productivity, the sorption
Hydrogen is one of the most efficient among the potential en- enhanced methane steam reforming (SE-MSR) has been proposed
ergy carriers currently exploitable [1, 2]. Still, its production gen- as an innovative technique consisting in a pre-combustion process
erates large amounts of carbon dioxide, which is an issue to be to convert fuel (natural gas) into a higher-heating-value and high-
carefully accounted for sustainable energy production. In fact, CO2 purity fuel (H2 ) with in situ CO2 capture [6-10]. In a conventional
emissions related to human activities are significantly contributing MSR process, three main reactions occur as follows [11]:
to global climate change [3,4].
CH4 (g ) + H2 O(g )  CO(g ) + 3H2 (g ) H298K = 206.2 kJ/mol (1)
New and highly efficient energy production systems with inte-
grated CO2 capture processes are being developed worldwide with
CH4 (g ) + 2H2 O(g )  CO2 (g ) + 4H2 (g )
particular attention to minimizing carbon dioxide adverse environ-
mental impacts [5]. Methane steam reforming (MSR) is the pri- H298K = 164.9 kJ/mol (2)

∗ CO(g ) + H2 O(g )  CO2 (g ) + H2 (g ) H298K = −41.1 kJ/mol


Corresponding author.
E-mail address: Benguerbayacine@yahoo.fr (Y. Benguerba). (3)

https://doi.org/10.1016/j.ijheatmasstransfer.2020.120635
0017-9310/© 2020 Elsevier Ltd. All rights reserved.
A. Neni, Y. Benguerba, M. Balsamo et al. International Journal of Heat and Mass Transfer 165 (2021) 120635

The formation of coke is given by the following three reactions


Nomenclature [12-17]:

C Inertial loss coefficient (m−1 ) CH4 (g )  (s ) + 2H2 (g ) H298K = 74.9 kJ/mol (4)
Dp Mean particle diameter (m)
Dr Internal reactor diameter (m) C(s ) + H2 O(g )  CO(g ) + H2 (g ) H 298K = 131.3 kJ/mol (5)
Dj,m Diffusion coefficient in the mixture of j-th species
(m2 s−1 ) C(s ) + CO2 (g )  2CO(g ) H298K = 172.0 kJ/mol (6)
Djk Binary mass diffusion coefficient of j-th species (m2
s−1 ) The advantages of coupling reaction systems with a certain
Eai Activation energy of the i-th reaction (J mol−1 ) form of in situ separation for one or more undesired reaction prod-
Fi External body force component (kg m−2 s−2 ) ucts have been widely reported in the literature, mainly through
G Gravity acceleration (m s−2 ) modeling works [8,9]. The use of CaO-based sorbents to capture
hf Fluid enthalpy (kJ kg−1 ) CO2 has become of special interest in recent years since it is one
of the most promising and low-cost methods to reduce CO2 emis-
hs Solid enthalpy (kJ kg−1 )
sions from flue gases [18]. Many researchers have reported high
kj Rate constant of j-th reaction
CO2 capture capacities for calcium oxide-based sorbents [19,20],
Jj Diffusion flux of j-th species (kmol m−2 s−1 )
and the carbonation process (highly exothermic) is described by
Ki Equilibrium constant i-th reaction (bar−1 ) the following chemical reaction:
K ji Adsorption constant of j-th species in the i-th reac-
tion CaO(s ) + CO2 (g )  CaCO3 (s ) H298K = − 175.7 kJ/mol (7)
L Reactor length (m)
The reverse reaction (calcination) allows recovering carbon
Mi Molecular weight of i-th species (kg kmol−1 )
dioxide in a concentrated form, realizing the simultaneous regen-
pi Partial pressure of i-th species (Pa)
eration of CaO. In the literature, CaO-based sorbents were selected
P Pressure (Pa)
on the basis of several parameters including high selectivity and
R Universal gas constant (kJ kmol−1 K−1 )
CO2 capture capacity, high capture rate, stable capacity for cyclic
Ri Reaction rate of i-th species (kmol m−3 s−1 )
operation, good mechanical strength in particular for fluidized beds
T Temperature (K)
operated at high pressure in the presence of steam.
Tin Feed gas temperature (K)
For the MSR process, the catalyst most commonly employed in
T0 Initial solid temperature (K)
the industrial practice is the commercially available nickel sup-
Tw Wall temperature (K)
ported on alumina (Ni/Al2 O3 ) because of its low cost and high
Tf Fluid temperature (K)
porosity support coupled with large pores, which enables a fast
U Gas velocity (m/s)
diffusion process [21-23]. Specific studies have been carried out,
X Reactor coordinate (m)
also investigating the effect of the catalyst particle shape on the
Xmax Maximum CaO conversion degree (-)
catalytic activity, measured in terms of methane conversion, but
Yj Mass fraction of species j (-)
any significant difference among a simple cylinder, a Raschig ring,
Greek letters a 7-holes cylinder, and a 7-holes sphere was observed [24].
A Permeability coefficient (m2 ) High-pressure SMR operations are commonly adopted in large-
ε Fixed-bed void fraction (-) scale applications to reduce the size of the reactor and the H2 pro-
G 0 i Gibbs free energy for the i-th reaction (kJ kmol−1 ) duction costs [25]. Fixed-bed catalytic reactors represent a possible
Hi Heat of i-th reaction (kJ kmol−1 ) alternative to fluidized beds because they are more easily operated
H ji Adsorption heat for the j-th species involved in the at high pressures, without the need for the gas/solid separation
i-th reaction (kJ kmol−1 ) step, and because of the operational problems caused by attrition
h Effectiveness factor (-) and elutriation of the solid material are negligible [26,27]. Balasub-
λe f f Medium effective heat conduction coefficient (J s−1 ramanian et al. [28] observed that a flow of high purity H2 gas (dry
m−1 K−1 ) yield of H2 ~ 95%) could be obtained from a SE-MSR process using
λf Gas heat conduction coefficient (J s−1 m−1 K−1 ) a fixed-bed reactor. Barelli et al. [29] evaluated the performance
λs Solid heat conduction coefficient (J s−1 m−1 K−1 ) of a new sorbent based on the incorporation of CaO particles into
μ Viscosity (Pa s) calcium aluminates as a part of a SE-MSR process carried out for
ρf Fluid Density (kg m−3 ) several cycles in a fixed-bed reactor. The authors claimed that the
ρcat Catalyst density (kg m−3 ) new sorbent determined higher H2 purity and CO2 adsorption than
ρcaO CaO density (kg m−3 ) conventional materials. In the numerical study performed by Fer-
ρs Solid density (kg m−3 ) nandez et al. [6] about the carbonation/reforming of SE-MSR in an
τi j Stress tensor (kg m−1 s−2 ) autothermal fixed-bed, the authors obtained an outlet H2 molar
fraction and a methane conversion of 95% and 85%, respectively.
This article aims to deepen the analysis of the feasibility of
The first reaction, called the steam reforming reaction (MSR), the SE-MSR process in a fixed-bed reactor with a Ni/Al2 O3 catalyst
describes the conversion of methane to carbon monoxide. The by performing computational fluid dynamics (CFD) simulations to
third one, called water gas-shift reaction, regulates the composi- achieve a better understanding of the interactions between mass
tion in terms of carbon monoxide and carbon dioxide ratio. This re- and heat transport and chemical reactions in an adiabatic catalytic
action is exothermic and is more thermodynamically favored than reactor, based on a fixed-bed scheme [30,31]. More specifically, the
the MSR reaction Eq. (1)) at temperatures lower than 723 K [5]. effect of CO2 capture onto a CaO sorbent on the performances of
Reaction (2) is taken into account because CO2 can also be pro- the SE-MSR process was investigated. One innovative element of
duced directly from methane reaction with water vapor [6]. The the present work relies on considering an effective medium con-
MSR (Eq. 1) and Global MSR (Eqs. 1-(2) reactions are highly en- ductivity for the case of thermal equilibrium between the solid
dothermic, as indicated by the molar enthalpy of reaction values. (sorbent+catalyst) and gaseous phase, which has never been con-

2
A. Neni, Y. Benguerba, M. Balsamo et al. International Journal of Heat and Mass Transfer 165 (2021) 120635

Table 1 (11)
Reactor characteristics and operating conditions [6]
The coke formation in the reforming of hydrocarbons is known
Parameter Value
to cause catalyst deactivation, and its kinetics can be expressed as:
Feed gas temperature, Tin 923 K  
Initial solid temperature, T0 923 K PH22
Total pressure, P 3.5 MPa k4 KCH4 PC H4 −
K4
Steam to carbon (S/C) molar ratio 5 R4 =  2 (12)
Inlet gas mass flux 3.5 kg/m2 s
1
Maximum CaO conversion, Xmax 0.4 1 + KC H4 ,4 PC H4 + PH12.5
Effectiveness factor, η 0.3 KH2 ,4
Catalyst apparent density, ρ cat 330 kg/m3  
Sorbent apparent density, ρ cao 1100 kg/m3 k5 PH2 O PCO
Solid density, ρ s 1675 kg/m3

KH2 O PH2 K5
Catalyst particle diameter, Dp 0.01 m R5 =  2 (13)
Reactor length, L 7m 1 PH2O 1 1. 5
Reactor diameter, Dr 0.3 m 1 + KC H4 ,5 PCH4 + + P
Bed void fraction, ε 0.5 KH2 O,5 PH2 KH2 ,5 H2
Solid thermal conductivity, λs 1.75 W/m K  
Fluid thermal conductivity, λ f 0.0454 W/m K k6 PC O2 PCO

KCO KC O2 PCO K6
R6 =  2 (14)
1 PC O2
sidered in the literature to the best of our knowledge. This is 1 + KCO,6 PCO +
deemed to be crucial for a correct analysis of the transient energy KCO,6 KC O2 ,6 PCO
evolution in the fixed-bed, which significantly affects the reaction PC H4 , PC O2 , PCO , PH2 and PH2 O [bar] are the partial pressures of CH4 ,
rates, and consequently, the composition of the obtained gaseous CO2 , CO, H2, and H2 O gas species, respectively; ki is the kinetic rate
phase. Overall, the proposed numerical analysis allows determin- constant of the reaction i:
ing the parameters necessary for the design and development of  
the assumed system for effective methane steam reforming under ki = k0i × exp −Eai/RT (15)
conditions of energy self-sufficiency.
Ki is the equilibrium constant for reaction i:
2. Materials and methods  
Ki = K0i × exp −G0i/RT (16)
2.1. Reference apparatus and operative conditions
K ji is the adsorption constant for the j-th species involved in the
i-th reaction:
The data related to operating conditions, reactor, sorbent, and  
catalyst used in the simulations are reported in Table 1. The ref- K ji = K 0ji × exp −H ji/RT (17)
erence reactor and its parameters (L, Dr, ε , etc.) were taken from
the literature [6]. Under the operating conditions of the carbon- In the Eqs. reported above, Eai (J/mol) is the energy of activation,
ation stage, the value of the inlet gas mass flux allows reaching R (J/mol K) is the universal gas constant; T (K) is the gas temper-
a superficial velocity of about 0.5 m/s, which is close to the nor- ature in the reaction zone. G0i and H ji are the Gibbs free energy
mal operation range commonly adopted in industrial SMRs [6]. The for the i-th reaction and the heat of adsorption for the j-th species
solid forming the fixed-bed is composed of 30 wt % Ni-based cata- involved in the i-th reaction.
lyst and 70 wt % of the CaO-based sorbents, as in a typical SE-MSR Kinetic and thermodynamic parameters, with the rate constants
reactor [6,28]. of all the chemical reactions, are given in Table 2.
And the final kinetic for sorption/desorption reaction of CaO
2.2. Kinetic models with CO2 is given by [6]:
η dX
The simulations of the methane catalytic steam reforming were R7 = (18)
conducted in a fixed-bed reactor. The reactor model has been de- MCaO dt
signed for adiabatic conditions so that it can benefit from the ther- The conversion X is given by the following differential equation
mal neutrality of the overall SER by using CaO as a regenerable CO2 [6]:
sorbent (reaction 7) [6]. This means that the heat necessary for
dX  
the endothermic reactions is generated by the exothermic calcina- = kcarb (Xmax − X ) νCO2 − νCO2.eq (19)
dt
tion process. The kinetic expressions for the steam reforming and
water gas shift reactions were obtained by adopting the Langmuir- where kcarb is the reaction rate constant of active CaO sorbent (de-
Hinshelwood equations reported by [32]: termined as 0.35 s−1 [6]), Xmax is the maximum carbonation conver-
  sion of CaO. νCO2 and νC O2 .eq are the gas-phase mole fraction and the
1 k1 PH32 PCO equilibrium mole fraction of CO2 in the reactor, respectively. T is
R1 = PC H4 PH2 O − (8)
DE N 2 PH2.5 K1 the fluid temperature; MCaO is the molecular weight of CaO, and ηis
 
2
the effectiveness factor for all the seven chemical reactions [6].
1 k2 PH42 PC O2
R2 = PC H4 PH22 O − (9)
DE N 2 PH3.5 K2 2.3. Simulation of a fixed-bed reactor
 
2

1 k3 PH2 PC O2 2.3.1. Governing equations


R3 = PCO PH2 O − (10)
DE N 2 PH2 K3 A 3D computational domain was assumed as any other differ-
ent choice would exert a significant effect on the retrieved results
The denominator (DEN) is given by the following expression:
[33]. In the 3D unsteady state CFD model formulation, the compu-
PH2 O tational domain representing the packed catalytic bed is solved as
DEN = 1 + KCO,123 PCO + KH2 ,123 PH2 + KC H4 ,123 PC H4 + KH2 O,123 a continuum (Fig. 1).
PH2

3
A. Neni, Y. Benguerba, M. Balsamo et al. International Journal of Heat and Mass Transfer 165 (2021) 120635

Table 2
Kinetic and equilibrium parameters of SE-MSR reactions [kmol, bar, s]
 240100   0

k1 1.17 × 1012 exp − KC H4 ,5 3.49 × exp −
RT RT
 243900   97770 
k2 2.83 × 1011 exp − KH2 O,5 4.73 ×10−6 exp +
RT RT
 67130   216145 
k3 5.43 × 102 exp − KH2 ,5 1.83 1013 × exp −
RT RT
 58893   100395 
−6
k4 6.95 × 10 exp −
3
KCO,6 7.34 10 × exp +
RT RT
 166397   104085 
k5 5.55 × 109 exp − KC O2 ,6 2.81 10 exp −
7
RT RT
 243835   2240 0 0 
k6 1.34 × 1015 exp − K1 exp − + 30.114
RT RT
 38280   36580 
KC H4 ,123 6.66 × 10−4 exp K3 exp − 4.036
RT RT
 88680 
KH2 O, 123 1.77 × 10+5 exp − K2 Kp1 ∗ Kp3
RT
 82900   84400 
KH2 ,123 6.15 × 10−9 exp + K4 2.98 × 105 exp −
RT RT
 70650   125916 
KCO,123 8.25 × 10−5 exp K5 1.3827 × 107 exp −
RT RT
 567   168527 
KC H4 ,4 0.21 × exp − K6 1.9393 × 109 exp −
RT RT
 133210   20474 
KH2 ,4 5.18 × 10 e xp −
7
νCO2 .eq 4.137 10 exp −7
RT T

and
3.5 ( 1 − ε )
C= (24)
Dp ε3
In the case of thermal equilibrium between the porous medium
and the fluid flow, an effective conductivity is considered, and the
transient term includes the thermal inertia of the solid region on
the medium. The conservation of energy for an adiabatic reactor is
given by the following equation:
     
Fig. 1. Geometric configuration of the reactor module. ∂ ε ρ f h f + ( 1 − ε ) ρs h s ∂ ui ρ f h f ∂ ∂T
+ = λe f f
∂t ∂ xi ∂xj ∂ xi
The mass conservation principle leads to the mass continuity 
Nr
+ (Hi ηi Ri ) (25)
equation defined as follows:
i=1
   
∂ ρf ∂ ρ f ui where h f is the total fluid enthalpy; hs is the total solid medium en-
+ =0 (20)
∂t ∂ xi thalpy; ρs is the solid medium density; Hi is the reaction enthalpy;
λe f f is the effective thermal conductivity of the medium, expressed
in which ρ f is the mass density; xi (i = 1, 2, 3) are the Cartesian co- as:
ordinates, and ui are the velocity components [34,35,39]. The equa-
tion for conservation of momentum in the direction i and a non- λe f f = ε λ f + (1 − ε ) λs (26)
accelerating reference frame is given as follows:
    in which λ f is the thermal conductivity of the gas and λs is the cor-
∂ ρ f ui ∂ ρ f ui u j ∂ P ∂ τi j responding value of the solid.
+ =− + + ρ f gi + Fi (21)
∂t ∂xj ∂ xi ∂ x j The conservation of chemical species is given by the following
equations:
In this balance, P is the static pressure, τi j is the stress tensor,    
and gi is the gravitational body force. Fi is an external body force ∂ ρY j ∂ ρ ui Y j ∂    Nr
 
component; it can include forces from the interaction between the + =− J j,i + α ji ηi Ri (27)
∂t ∂ xi ∂xj i=1
phases, centrifugal forces, Coriolis force, and user-defined forces.
The body force F defined for the porous zone (catalyst/sorbent ∂Yj
J ji = −ρ D jk (28)
zone) is given by [36]: ∂xj
 μU  ρ f U U
F =− +C (22) J ji is the diffusion flux of species j; Ri is the i-th reaction rate; Y j is
α 2 the mass fraction of the j-th species in the mixture (Y j = mj ) be-
m

F is 0 for gaseous zones. In the above equation, the perme- ing mthe total mass and D jk the binary mass diffusion coefficient of
ability, α , and the inertial loss coefficient, C, were derived by the the j-th component in the k-th component. The fluid density was
Blake-Kozeny equation [37]as: computed from the Peng-Robinson real gas equation for a multi-
component mixture.
D2p ε3 All the kinetics have been implemented using a User Defined
α= (23)
Function (UDF) programmed in C++ [37].
150 (1 − ε )2

4
A. Neni, Y. Benguerba, M. Balsamo et al. International Journal of Heat and Mass Transfer 165 (2021) 120635

Fig. 2. Dynamic temperature profiles.

2.3.2. Numerical methods and computational tools 3. Results and discussion


The resolution scheme uses the pressure-velocity coupling so-
lution method with a second-order upwind discretization for the Figs. 2(a) and 2(b) show the dynamic temperature profiles at
system of species equations, first-order upwind for momentum and different axial positions in the fixed-bed. It can be noticed that
energy, and second-order for pressure. Monitoring the residuals at t=12 s, the hot spot located at a distance of 0.5 m from the
of all variables and the mass flow of species is the key parame- reactor inlet moves towards the reactor exit as the process pro-
ter for the convergence of the calculations. The solution was as- ceeds. The temperature of the moving hot spot reaches a maxi-
sumed to have been reached when these residuals are less than mum value at the exit of the reactor (L=6.99 m) at approximately
or equal to 10−8 with constant mass flow values. The resolution 615 s (Tmax =968.73 K). It is noticed in Fig. 2(b) that whatever the
of the model equations set is a three-dimensional map of the ve- time of the process, the temperature drops from 923 K (inlet tem-
locity and concentration of all gas species in each of the calcula- perature) to a value that depends on the intensity of the involved
tion cells. To ensure that the solution at every position for differ- SE-MSR chemical reactions Eqs. (1)-((7)) (Fig. 3). Moreover, the
ent times was independent of the mesh size, the mesh was re- thermal profile inside the reactor is strictly dependent on the car-
fined more and more by increasing the number of cells. If the re- bonation reaction, and the flat zones are associated with a block of
sults stop changing with these successive mesh refinements, the this reaction, as explained in the following.
solution was considered to be independent of the mesh. The nu- This temperature-drop highlights that the beginning of the
merical results are compared with the available experimental data chemical process is globally endothermic. In Fig. 3, the kinetic rates
[6] for a detailed quantitative analysis. The kinetics were estimated of the occurring reactions 1-7 are reported. The exothermic reac-
using user-defined functions (UDFs). The experimental setup made tions 3 and 7 (see Fig. 3(c) and 3(g)) are not able to compensate for
it possible to obtain the concentration of the target compounds at the heat consumed by the main steam reforming reactions 1 and
the inlet and the outlet of the fixed bed catalytic reactor. Therefore, 2 (Fig. 3(a) and 3(b)). It is to be noted that the energy required
simulations using CFD were performed to simulate these profiles. for the production of CO2 by reaction 2 is satisfied by the removal
The geometric domain was subdivided into several control volumes of carbon dioxide via the exothermic carbonation (reaction 7). It is
(Fig. 1). For each of these control volumes, the governing equa- also observed that, after a certain length, reaction 7 starts to take
tions that express the conservation of mass, momentum, and en- place at a significant rate, and the temperature thus begins to in-
ergy were discretized. This leads to a system of algebraic equations crease until reaching a maximum (hot spot).
connecting the flow parameters at characteristic points associated After the consumption of most of the CO2 released (15.064
with the control volumes. Here, the least-squares cell-based tech- mol/m3 , see Figs. 4(c)), reaction 7 weakens and can no longer com-
nique was used. The resulting set of algebraic equations was solved pensate for the energy consumed by endothermic reactions 1 and
iteratively. 2, which in turn, weakens with the decrease of CH4 and H2 O con-

5
A. Neni, Y. Benguerba, M. Balsamo et al. International Journal of Heat and Mass Transfer 165 (2021) 120635

Fig. 3. Reaction kinetics distribution along the reactor at different times: R1 (a), R2 (b), R3 (c), R4 (d), R5 (e), R6 (f), R7 (g)

6
A. Neni, Y. Benguerba, M. Balsamo et al. International Journal of Heat and Mass Transfer 165 (2021) 120635

Fig. 4. Axial species molar concentration profiles: a) CH4 , b) H2 O, c) CO2

centrations (Figs. 4(a), 4(b)). Then, a tendency towards an equi- probably due to the fact that they did not account for the thermal
librium gas temperature slightly higher than the inlet tempera- conductivity of the solid (adsorbent + catalyst), which may slow
ture is noticed. Moreover, the outlet temperature at L=6.99 m (see down the axial heat transfer towards the reactor exit. In Fig. 2(b),
Fig. 2(a)) increases to a maximum of 968.73 K at 615 s and then the maximum outlet temperature value is obtained at 615 s, while
drops, which is related to the heat evolution and transfer along in the work of Fernandez et al. [6] the peak was found at around
the reactor axial coordinate. In practice, the above temperature 900 s. In general, the molar concentration values at the reactor exit
changes may be due to the much higher thermal capacity and ther- are in good agreement with those calculated by Fernandez et al.
mal conductivity performance of the solid phase (sorbent and cat- [6]. These results show that, in the investigated simulation condi-
alyst) with respect to the one of the gas phase. tions, the time-scale of heat transfer phenomena is far larger than
Fig. 5(a) shows the time evolution of the concentrations of each those related to mass transfer, specifically referred to as the CO2
component at the outlet of the adiabatic fixed-bed reforming reac- capture step, which is the key step of the proposed process. Hence,
tor. Due to the occurring of chemical reactions, CH4 and H2 O are the simulation of the latter was limited to a smaller time-scale, af-
continuously consumed, while H2 , CO2 , CO, and a minimal coke- ter which the methane steam reforming would not be carried out
amount is produced. However, they undergo further reactions that with the beneficial CO2 capture integration.
tend to alter (mainly diminishing) their concentrations. During the The analysis of the simulation results should mainly focus on
first 500 s, the molar concentration profiles are almost constant the principal target of the SE-MSR process, i.e., H2 , methane con-
with time, which corresponds to the set of parameters of a station- version, and H2 /CO. The outlet concentration of hydrogen strongly
ary process (in the gas side). Just after, the CO2 concentration starts depends on the efficiency of the sorbent (see Fig. 5(b)), as it starts
to increase again (a first increase is observed after 100 s and a first to decrease when the sorbent adsorption capacity starts to be satu-
plateau is attained for times ranging from about 200 and 500 s), rated. The MSR takes place in parallel with the capture of CO2, and
likely due to the saturation of a significant fraction of CaO (mainly, the latter leads to improve the hydrogen equilibrium concentration
for the sorbent located in the first half of the total length of the in the outlet gas. Thus, it is reasonable to say that the main factor
reactor, cf. Fig. 6) devoted to its capture. In turn, the concentration enhancing hydrogen production is the CO2 capture. The value cal-
of the other compounds starts to vary, as a cascade process orig- culated by Fernandez et al. [6] for the outlet hydrogen mole frac-
inating from CaO, which is significantly reducing its carbonation tion (dry basis) was around 0.9, while in our work, it was of the
activity. Also, the concentration of coke tends to a minimum value, order of 0.8. This difference, as explained above, is ascribable to the
which is a negative factor in the conduction of the MSR process. thermal conductivity of the solid that our model takes into consid-
This pre-breakthrough period was found to be about 500 s, while eration, which determines lower temperatures that affect the en-
in Fernandez et al. [6], this period was found to be 720 s. This is dothermic reactions. This value drops to 0.42 once the adsorbent

7
A. Neni, Y. Benguerba, M. Balsamo et al. International Journal of Heat and Mass Transfer 165 (2021) 120635

Fig. 5. Dynamic Outlet profiles: a) Species concentrations, b) Hydrogen yield, c) Methane conversion and H2 /CO ratio

8
A. Neni, Y. Benguerba, M. Balsamo et al. International Journal of Heat and Mass Transfer 165 (2021) 120635

a drop in temperature, as already stated. Reactions 4, 5 and 6


(Fig. 3(d), 3(e) and 3(f)) were found to give very low and insignif-
icant rates compared to reactions 1, 2, 3 and 7 (between 10−3 and
10−5 mol/(m3 s1 )). The trends of the reaction rates with reactor
length are similar, showing that temperature is the most influ-
ential and limiting parameter in reagents and adsorbent conver-
sion. Reactions 5 and 6 are found to prevailingly occur towards
reagents production, which is indicative of high produced hydro-
gen quantities (due to SE-MSR). The following inequalities are sat-
PH O PCO PC O PCO ,
isfied: 2
PH < K5 and PCO
2
< K6 which show that the reverse reac-
2
tions prevail (see Eqs. (13) and ((14)). It’s concluded that the quan-
tity of H2 O is not sufficient to convert the coke by reaction 5 to
more hydrogen and, in turn, led to the conversion of CH4 to C(s)
by reaction 4.
In Fig. 6, the dynamic behavior of the CaO conversion along
the reactor length and for subsequent time spots is reported. Af-
ter the first 600s, CaO conversion is maximum in the middle of
the reactor, while lower figures are retrieved at the outlet. Note
that a CaO saturation front advances along the length of the reac-
Fig. 6. Dynamic behavior of the CaO conversion
tor as a function of time. Fig. 6 shows that saturation begins in the
area of the reactor inlet and progresses towards the reactor out-
let. This behavior seems to be related again to the crossing effect
reaches saturation, and CO2 is not captured yet (0.58 is reported in
between reaction courses and temperature profiles. The advancing
Fernandez et al. [6] as an example). Further essential parameters
of endothermic reforming reactions provides the reactants for the
can be calculated in order to better describe the performances of
exothermic CO2 capture reaction. Moreover, when times goes on,
the system as far as the CO2 capture step is active. For instance, the
it is evident that the decreasing of CO2 concentration, which is ad-
selectivity in hydrogen relative to the quantity of CH4 initially in-
sorbed in the first part of the reactor length, determines a lower
troduced into the reactor is of the order of 1.75 (with CO2 -capture)
saturation of the final part of the reactor that, as expected, start
and 0.8 (without CO2 capture), which provides an idea of the very
to saturate only in the final steps of the process. The flat part of
high importance of using an effective CO2 sorption step to im-
the curves is obtained when dx/dt=0, which is possible only if
prove the performance of the MSR. The corresponding hydrogen
νCO2 = νCO2.eq . The calculated values of νCO2.eq are in the range be-
produced molar fluxes are approximately 7.56 mol s−1 (with CO2
tween 0.009 (T=923K) and 0.0247 (T=968.7K). Thus, in this part of
capture) and 3.46 mole s−1 (without CO2 capture), respectively,
the reactor, it is likely that, due to the complex thermal dynamics
once again confirming the significant contribution provided by the
discussed above, the molar fraction of CO2 is equal to the molar
CO2 sorption step.
fraction of CO2 at equilibrium and the carbonation reaction tem-
In the breakthrough period between 200 s and 500 s, a decreas-
porarily blocked down.
ing quantity of all the outlet products H2 , CO, and C concentration
A matching between temperature profiles (Fig. 2) and SE-MSR
was noticed, with an increasing outlet concentration of CO2 and of
process trends (Figs. 3–6) shows the significant contribution of the
the reagents (CH4 and H2 O). This is due to the saturation of the
thermal conductivity inside the fixed-bed, whose limitation could
adsorbent that, as already stated above, in turn, determines alter-
contribute to further enhance the hydrogen production.
ation of the process stationary state (gas side). Fig. 5(c) shows a
Reynolds number distribution in the reactor is an important
decrease in the CH4 conversion after the breakthrough time. The
parameter shown by the CFD visualization technique. As shown
H2 /CO ratio increases considerably due to the low concentration of
in Fig. 7(a), the Reynolds number’s average value is about 880,
CO, even though the production of hydrogen weakens. Moreover,
showing a laminar flow. At all the longitudinal positions, the drop
the absence of the exothermic reaction also determines alteration
in the gas mixture density-value (Fig. 7(b)), due to the tempera-
of the thermal regimes since there is no contribution of the heat
ture changes in the reactor (Figs. 2(a) and 2(b)), is compensated
deriving from carbonation reaction (7), and the decrease of the
by the increase in the gas mixture velocity (Fig. 7(c)). This com-
temperature determines a lower conversion of CH4 and H2 O, ac-
pensation effect between gas mixture density and velocity leads
cording to the endothermic reforming reactions. When CaO reaches
to the almost constant Reynolds number along the reactor length
its maximum conversion (R7 = 0), which can be clearly observed
(L) over time. Fig. 7(d) shows the absolute pressure calculated by
in both Fig. 2(b) and Fig. 3(g), the endothermicity of reactions (1)
the Peng-Robinson equation of state, where the pressure-drop was
and (2) prevails over the exothermicity of reaction 3, leading to a
calculated by eq. (22). The total pressure, which is the sum of
temperature drop in the reaction zone. Despite the temperature-
the absolute and the dynamic pressure, is related to temperature,
drop at the reactor inlet, the temperature remains higher in the
gas mixture velocity, and density. An increase in the fluid veloc-
rest of the reactor, likely due to the poor thermal conduction of
ity determines augmented pressure drops. It simultaneously deter-
the fixed-bed. The kinetics R1 , R2 , and R3 (Figs. 3(a), 3(b) and 3(c))
mines an advantage in terms of enhanced heat and mass trans-
show a continuous conversion of CH4 and H2 O into H2 , CO, CO2
fer rates. This analysis aims to validate the hypothesis of a lam-
(which is no longer adsorbed) and coke.
inar regime flow, considering the fairly average Reynolds values
Concerning the difference along the reactor length, the kinet-
calculated beside the relatively low-pressure drop. The obtained
ics R1, R2 , and R3 Fig. 3(a), 3(b) and 3(c)) resulted in being more
pressure drops values (the maximum pressure drop was 10 kPa)
intense at the inlet where the temperature was around 923 K.
were found comparable with those of typical industrial SMR fixed-
At the same time, they slowed down with the decreasing tem-
bed reactors [32, 38]. Furthermore, Fig. 7(e) shows a Prandtl num-
perature along the reactor, which was due to the endothermic-
ber value between 1.8 and 2.07, which indicates that the dissipa-
ity of the global reaction. Reactions 1 and 2 (endothermic) were
tion of both fluid momentum and heat occurs at about the same
more intense than reactions 3 and 7 (exothermic), and the re-
rate.
sult was an overall endothermic thermal balance that determined

9
A. Neni, Y. Benguerba, M. Balsamo et al. International Journal of Heat and Mass Transfer 165 (2021) 120635

4. Conclusion

In this work, a Sorption-Enhanced Steam Methane Reforming


(SE-MSR) process was simulated using a fixed-bed reactor filled
with 70% Ni/Al2 O3 catalyst and 30% CaO sorbent for CO2 , aimed
at investigating the effect on the H2 production. The reactor was
simulated in CFD environment as adiabatic, exploiting the ther-
mal neutrality of the combination of the overall SER reaction
(endothermic) and the CO2 capture by CaO regenerable sorbent
(exothermic). The model accounted for the effective thermal con-
ductivity in both fluid and solid, which significantly influenced
the retrieved results. More in general, the correlation between the
course of the reactions involved in the process and the trend of
the temperature, both along the reactor and during the time-steps,
played a crucial role in determining the dynamics of the simulated
process. The base-idea of the process is that the energy require-
ment for methane steam reforming is satisfied by the removal of
carbon dioxide via the exothermic carbonation reaction.
Due to the balance between these two opposite tendencies, the
temperature profiles at different axial positions in the fixed-bed
showed the existence of a hot spot located at a distance of 1.5 m
from the reactor inlet, which moves towards the reactor exit as the
process proceeds. The outlet temperature increases to a maximum
of 968.73 K after 615 s and then drops related to the heat evolution
and transfer along the reactor axial coordinate.
H2 production was constant and maximum during the first
500s. The CO2 concentration starts to increase, likely due to the
beginning of the saturation of CaO sorbent, which starts to reduce
its CO2 capture capacity. In turn, the concentration of the other
compounds starts to vary, as a cascade process originating from
CaO saturation. When the exothermic reaction stops, an alteration
of the thermal regimes was observed. The absence of the contribu-
tion of the exothermic carbonation reaction resulted in a decrease
of the temperature, which in turn determined a lower conversion
of CH4 and H2 O, according to the endothermic reforming reactions.
The maximum H2 outlet molar fraction (dry basis) experienced
was around 0.8 in correspondence of the presence of the CO2 sorp-
tion step contribution. The value dropped to 0.42 once the adsor-
bent reached the saturation. The selectivity in hydrogen relative to
the quantity of CH4 initially introduced into the reactor was of the
order of 1.75 (with CO2 -capture) and 0.8 (without CO2 capture),
which testifies the very high importance of using an effective CO2
sorption step to improve the performance of the MSR.
In conclusion, CO2 sorption is able to significantly enhance the
methane steam reforming. However, the thermal conductivity of
the solid phase (catalyst + sorbent) resulted in being a limiting
factor in the heat transfer to the reactor outlet. This should drive
to store the heat produced by exothermic reactions and, thus, to
be able to produce a higher amount of hydrogen.

Declaration of Competing Interest

The authors declare that they have no known competing finan-


cial interests or personal relationships that could have appeared to
influence the work reported in this paper.

CRediT authorship contribution statement

Amira Neni: Investigation, Writing - original draft. Yacine


Benguerba: Conceptualization, Methodology, Software, Writing -
Fig. 7. Dynamics plots in the reactor at different axial positions (L), (a) Reynolds review & editing, Supervision. Marco Balsamo: Conceptualization,
number, (b) Gas mixture density, (c) Velocity, (d) Total pressure, (e) Molecular
Prandtl number
Methodology, Writing - review & editing, Supervision. Alessandro
Erto: Conceptualization, Methodology, Writing - review & editing,
Supervision. Barbara Ernst: Writing - original draft. Djafer Bena-
chour: Writing - original draft, Supervision.

10
A. Neni, Y. Benguerba, M. Balsamo et al. International Journal of Heat and Mass Transfer 165 (2021) 120635

References [20] Z. Zhou, Y. Qi, M. Xie, Z. Cheng, W. Yuan, Synthesis of CaO-based sorbents
through incorporation of alumina/aluminate and their CO2 capture perfor-
[1] V. Dupont, B. Ross A, E. Knight, I. Hanley, M.V. Twigg, Production of hydro- mance, Chemical Engineering Science 74 (2012) 172–180.
gen by unmixed steam reforming of methane, Chemical Engineering Science [21] A. Kaengsilalai, A. Luengnaruemitchai, S. Jitkarnka, S. Wongkasemjit, Poten-
63 (2008) 2966–2979. tial of Ni supported on KH zeolite catalysts for carbon dioxide reforming of
[2] J.M. Ogden, Review of Small Stationary Reformers for Hydrogen Production, methane, J. Power Sources 165 (2007) 347–352.
The international Energy Agency (2001) Golden IEA/H2/TR-02/002. [22] A. Lemonidu, M.A. Goula, I.A. Vasalos, Carbon dioxide reforming of methane
[3] JM. Alarcón, JR. Fernández, CaCO3 calcination by the simultaneous reduction over 5 wt.% nickel calcium aluminate catalysts ± effect of preparation method,
of CuO in a Ca/Cu chemical looping process, Chemical Engineering Science 137 Catalysis Today 46 (1998) 175–183.
(2015) 254–267. [23] K. Takehira, T. Shishido, P. Wang, T. Kosaka, K. Takaki, Autothermal reforming of
[4] J. Adanez, P. Gayán, G. Grasa, L.F. De Diego, L. Armesto, A. Cabanillas, Circu- CH4 over supported Ni catalysts prepared from Mg–Al hydrotalcite-like anionic
lating fluidized bed combustion in the turbulent regime: modelling of carbon clay, Journal of Catalysis 221 (2004) 43–54.
combustion efficiency and sulphur retention, Fuel 80 (2001) 1405–1414. [24] D. Pashchenko, Experimental investigation of reforming and flow characteris-
[5] L. Eduardo, G. Oliveira, Sorption enhanced steam methane reforming process tics of a steam methane reformer filled with nickel catalyst of various shapes,
for continuous production of hydrogen in pressure swing adsorptive reactors, Energy Conversion and Management 185 (2019) 465–472.
Chemical Engineering Science 84 (2009) 12–20. [25] D.P. Harrison, Sorption-enhanced hydrogen production: a review, Ind. Eng.
[6] J.R. Fernandez, J.C. Abanades, R. Murillo, Modeling of sorption enhanced steam Chem. Res. 47 (2008) 6486–6501.
methane reforming in an adiabatic fixed bed reactor, Chemical Engineering [26] JR Fernández, I Martínez, JC Abanades, MC Romano, Conceptual design of a
Science 84 (2012) 1–11. Ca–Cu chemical looping process for hydrogen production in integrated steel-
[7] JR. Fernandez, J.C. Abanades, G. Grasa, Modeling of sorption enhanced steam works, Int J Hydrogen Energy 42 (2017) 11023–11037.
methane reforming —Part II: Simulation within a novel Ca/Cu chemical loop [27] S Noorman, MVS Annaland, H. Kuipers, Packed bed reactor technology for
process for hydrogen production, Chemical Engineering Science 84 (2012) chemical looping combustion, Ind. Eng. Chem. Res. 46 (2007) 4212–4220.
12–20. [28] B Balasubramanian, AL Ortiz, S Kaytakoglu, DP. Harrison, Hydrogen from
[8] J.P. Jakobsen, E. Halmøy, Reactor modeling of sorption enhanced steam methane in a single-step process, Chem. Eng. Sci. 54 (1999) 3543–3552.
methane reforming, Energy Procedia 1 (2009) 725–732. [29] L Barelli, G Bidini, F. Gallorini, SE-SR with sorbents based on calcium alumi-
[9] T.J.R. Hendricus, D.E. Jurriaan, D.C. Paul, G.H. Wim, W.B. Ruud, Modeling study nates: process optimization, Appl Energy 143 (2015) 110–118.
of the sorption-enhanced reaction process for CO2 capture, I.Model develop- [30] Deutschmann O., 2008. Computational Fluid Dynamics Simulation of Catalytic
ment and validation. Industrial& Engineering Chemistry Research 48 (2009) Reactors. Handbook of heterogeneous catalysis (2nd Ed.), 1811-1828.
6966–6974. [31] Y. Benguerba, M. Virginie, C. Dumas, B. Ernst, Computational Fluid Dynamics
[10] Z.S. Li, N.S. Cai, Modelling of multiple cycles for sorption-enhanced steam Study of the Dry Reforming of Methane Over Ni/Al2O3 Catalyst in a Membrane
methane reforming and sorbent regeneration in fixed bed reactor, Energy and Reactor, Coke Deposition. Kinetics and Catalysis 58 (3) (2017) 328–338.
Fuels 21 (2007) 2909–2918. [32] J. Xu, G.F. Froment, Methane steam reforming methanation and water-gas shift,
[11] D. Ki Leea, Baekb Il., W. Yoonc, Modeling andsimulation for the methane steam Intrinsic kinetics. AlChE J. 35 (1989) 88–96.
reforming enhanced by in situ CO2 removal utilizing the CaO carbonation for [33] D. Pashchenko, Effect of the geometric dimensionality of computational do-
H2 production, Chemical Engineering Science 59 (2004) 931–942. main on the results of CFD-modeling of steam methane reforming, Interna-
[12] J.-W. Snoeck, G.F. Froment, M. Fowles, Filamentous carbon formation and gasi- tional Journal of Hydrogen Energy (2018) 8662–8673 43.18.
fication: Thermodynamics, driving force, nucleation and steady-state growth, [34] M. Coroneo, G. Montante, M.G. Baschetti, A. Paglianti, CFD modelling of in-
Journal of Catalysis 169 (1997) 240–249. organic membrane modules for gas mixture separation, Chem. Eng. Sci. 64
[13] L.A. Arkatova, The deposition of coke during carbon dioxide reforming of (2009) 1085–1094.
methane over intermetallides, Catal. Today 157 (2010) 170–176. [35] M. Coroneo, G. Montante, A. Paglianti, Numerical and experimental fluid-dy-
[14] M.C.J. Bradford, M.A. Vannice, CO2 reforming of CH4 over supported Pt cata- namic analysis to improve the mass transfer performances of Pd-Ag membrane
lysts, Journal of Catalysis 173 (1998) 157–171. modules for hydrogen purification, Ind. Eng. Chem. Res. 49 (2010) 9300–9309.
[15] K Nagaoka, K Seshan, K Aika, JA. Lercher, Carbon deposition during carbon [36] J. Xuan, M.K. Leung, D.Y. Leung, M. Ni, Integrating chemical kinetics with CFD
dioxide reforming of methane and comparison between Pt/Al2O3 and Pt/ZrO2, modelling for autothermal reforming of biogas, Int. J. Hydrogen Energy 34
Journal of Catalysis 197 (2001) 34–42. (2009) 9076–9086.
[16] A.D. Ballarini, S.R. de Miguel, E.L. Jablonski, O.A. Scelza, A.A. Castro, Reform- [37] C. Herce, C. Cortés, S. Stendardo, Computationally efficient CFD model for
ing of CH4 with CO2 on Pt-supported catalysts: effect of the support on the scale-up of bubbling fluidized bed reactors applied to sorption-enhanced
catalytic behaviour, Catalysis Today 107–108 (2005) 481–486. steam methane reforming, Fuel Processing Technology 167 (2017) 747–761.
[17] F. Guo, Y. Zhang, G. Zhang, H. Zhao, Syngas production by carbon dioxide re- [38] D. Pashchenko, Pressure drop in the thermochemical recuperators filled with
forming of methane over different semi-cokes, J, Power Sources 231 (2013) the catalysts of various shapes: A combined experimental and numerical in-
82–90. vestigation, Energy 166 (2019) 462–470.
[18] Deutschmann O., 2008. Computational Fluid Dynamics Simulation of Catalytic [39] Benguerba Y., Amer J., Ernst B., CFD modeling of the H2 /N2 separation with a
Reactors. Handbook of heterogeneous catalysis (2nd Ed.), 1811-1828. nickel/α -alumina microporous membrane, Chem. Eng. Sci. 123 (2015) 527–535,
[19] M.S. Yancheshmeh, R. Radfarnia H, C. Iliuta M, High temperature CO2 sorbents doi:10.1016/j.ces.2014.11.048.
and their application for hydrogen production by sorption enhanced steam re-
forming process, Chemical Engineering Journal 283 (2016) 420–444.

11

You might also like