Energies 13 00610 v2 PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

energies

Article
Modeling and Design of a Multi-Tubular Packed-Bed
Reactor for Methanol Steam Reforming over a
Cu/ZnO/Al2O3 Catalyst
Jimin Zhu *, Samuel Simon Araya , Xiaoti Cui , Simon Lennart Sahlin and
Søren Knudsen Kær
Department of Energy Technology, Aalborg University, Pontoppidanstræde 111, 9220 Aalborg Ø, Denmark;
[email protected] (S.S.A.); [email protected] (X.C.); [email protected] (S.L.S.); [email protected] (S.K.K.)
* Correspondence: [email protected]; Tel.: +45-9356-2443

Received: 6 December 2019; Accepted: 26 January 2020; Published: 31 January 2020 

Abstract: Methanol as a hydrogen carrier can be reformed with steam over Cu/ZnO/Al2 O3 catalysts.
In this paper a comprehensive pseudo-homogenous model of a multi-tubular packed-bed reformer
has been developed to investigate the impact of operating conditions and geometric parameters on
its performance. A kinetic Langmuir-Hinshelwood model of the methanol steam reforming process
was proposed. In addition to the kinetic model, the pressure drop and the mass and heat transfer
phenomena along the reactor were taken into account. This model was verified by a dynamic model in
the platform of ASPEN. The diffusion effect inside catalyst particles was also estimated and accounted
for by the effectiveness factor. The simulation results showed axial temperature profiles in both tube
and shell side with different operating conditions. Moreover, the lower flow rate of liquid fuel and
higher inlet temperature of thermal air led to a lower concentration of residual methanol, but also a
higher concentration of generated CO from the reformer exit. The choices of operating conditions
were limited to ensure a tolerable concentration of methanol and CO in H2 -rich gas for feeding into a
high temperature polymer electrolyte membrane fuel cell (HT-PEMFC) stack. With fixed catalyst
load, the increase of tube number and decrease of tube diameter improved the methanol conversion,
but also increased the CO concentration in reformed gas. In addition, increasing the number of baffle
plates in the shell side increased the methanol conversion and the CO concentration.

Keywords: methanol steam reforming; multi-tubular packed-bed reformer; hydrogen production;


temperature profile; geometric parameter

1. Introduction
The contribution of hydrogen to the promotion of green energy is mainly driven by recent
achievements, especially polymer electrolyte membrane (PEM) fuel cells, where hydrogen is used
as the fuel. One obstacle for the development of the hydrogen economy is the safe storage and
transportation challenge presented by hydrogen. Therefore, special attention has been paid to the
development of economical hydrogen production methods by chemically converting hydrocarbons or
alcohols to a hydrogen-rich synthesis gas stream. This process, generically called reforming, requires
oxidizing agents. When water is used as the oxidant, the process is known as steam reforming [1].
Among various hydrogen carriers, methanol stands out because of its properties of being liquid at
ambient conditions and infinitely miscible with water. Moreover, methanol has a low boiling point
(65 ◦ C) for vaporization, a relatively high H/C ratio (4:1), a low reforming temperature (200–300 ◦ C)
owing to the absence of a strong C–C bond, and is producible from various carbon-based feedstocks,
such as natural gas, biomass, and CO2 [2–4]. Under appropriate conditions, the most favored reaction

Energies 2020, 13, 610; doi:10.3390/en13030610 www.mdpi.com/journal/energies


Energies 2020, 13, 610 2 of 25

stoichiometry is the methanol steam reforming (MSR) reaction. One of the major advantages of the MSR
reaction is that 1/3 of the hydrogen product can be derived from water. In addition to the MSR reaction,
there are normally other two reactions that happen during the reforming process: the water-gas shift
(WGS) reaction and the methanol decomposition (MD) reaction. The three main reactions that take
place within the methanol steam reformer are shown in the following equations [5]:
Methanol steam reforming reaction (MSR):
 
CH3 OH + H2 O ←→ CO2 + 3H2 ∆H = + 49.7 kJ mol−1 , (1)

Water-gas shift reaction (WGS):


 
CO + H2 O ←→ CO2 + H2 ∆H = − 41.2 kJ mol−1 , (2)

Methanol decomposition reaction (MD):


 
CH3 OH ←→ CO + 2H2 ∆H = + 90.2 kJ mol−1 , (3)

Although hydrogen is the only desired product, other by-products are inevitably formed in
reformate gas mixture, such as carbon dioxide (CO2 ), small amounts of carbon monoxide (CO),
unconverted water and methanol vapor. The fractions of methanol, CO2 and especially CO in the
reformate gas should be minimized because of their poisoning effect on fuel cells [6,7]. Catalysts used
for methanol steam reforming are supposed to have as main properties good activity and fast kinetics
at low temperature, high selectivity to suppress CO production, good stability and long lifetime [5].
The most widely used commercial catalysts for the MSR process are Cu-based catalysts, especially
Cu/ZnO/Al2 O3, due to their relatively high activity and selectivity [8].
Methanol reforming methods carried out experimentally and industrially in packed-bed reactors
will inevitably result in high investment and operating costs. Therefore, there are numerous studies on
the kinetics and mechanisms of MSR over commercial Cu/ZnO/Al2 O3 catalysts. Jiang et al. [9] proposed
an expression of reaction rates based on power rate law kinetics. It is assumed that there was only one
kind of active site for reactions and the methyl formate was the intermediate. Peppley et al. [10] studied
the reaction network of MSR on the catalyst BASF K3-110. They assumed two distinct types of active
sites, one type for MSR and WGS reactions and the other for MD reaction. And a comprehensive kinetic
model was developed considering the surface mechanism of the catalyst. Agrell et al. [11] investigated
the MSR over a Cu/ZnO/Al2 O3 catalyst from Süd-Chemie (G-66 MR) and developed a kinetic model.
With operating temperatures below 220 ◦ C, an Arrhenius-type function was used; and with higher
temperatures, the mass transport hindered the reaction kinetics, hence a fifth degree polynomial was
used instead of the Arrhenius expression. Sandra et al. [12] and Herdem et al. [13] compared several
kinetic rate expressions of the MSR process. They found that a kinetic Langmuir-Hinshelwood model
which was developed by Peppley et al. [10] presented the best fit to the experimental data.
Another dominant factor for the reforming process is the reformer design. Conventional
packed-bed reformers use catalyst particles in the form of pellets or cylinders, which are versatile
for application at both the laboratory and industrial scale owing to their relatively low cost and easy
operation [14]. Nevertheless, one disadvantage of packed-bed reformers is the radial temperature
gradient in the catalyst bed [3]. Recent progress in micro-processing make it possible to manufacture
wall-coated micro-channel reactors and membrane reactors, which present fewer heat and mass
transfer limitations, less pressure drop, better selectivity, but also a drawback of lower specific catalyst
load [2,9,15,16]. However, potential barriers for the commercialization of micro-channel and membrane
reactors, such as high costs and low mechanical resistance, make packed-bed reformers still the most
widely used types in the chemical industry for extracting hydrogen from methanol.
Because of the endothermic characteristic of the reforming process, an external heating source is
needed to activate the reaction sites and prevent temperature drops in the catalyst bed. For on-site
Energies 2020, 13, 610 3 of 25

applications, it means that the reformer should be integrated with a heat supply unit, usually called
a catalytic combustor or burner [17–19]. In this system, a flow of thermal fluid is needed to transfer
heat from the burner to the catalyst bed. Reaction rates of methanol reforming predominately depend
on the local concentration and temperature correlated to heat and mass transfer mechanisms, which
should be investigated when designing a reactor. Yoon et al. [20] analyzed the dominant limiting
mechanisms (heat transfer, mass transfer and chemical kinetics) in the methanol steam reformer
theoretically and experimentally. Results showed that with the diminishing of catalyst size, the heat
transfer limitation increased and the mass transfer limitation decreased. Also with the diameter of
the reactor diminished, the heat and mass transfer were enhanced. Vadlamudi et al. [21] analyzed a
packed-bed reactor for autothermal reforming of methanol to produce sufficient hydrogen for a 100 W
fuel cell stack. They developed a 1-D non-isothermal model considering the steady state mass and
energy balance. The simulated results agreed well with experimental data and the pressure drop was
considered to be negligible. Ma et al. [22] investigated the hydrogen output and thermal behavior of a
plant-scale fixed-bed reformer for methanol steam reforming based on a 2-D pseudo-homogenous
model. The results showed that there was no obvious concentration gradients in the radial direction.
Moreover, with a larger tube diameter, the limited heat transfer would lead to a larger radial temperature
gradient in the catalyst bed. A similar result has been reported in [23] that the small ratio of tube
to particle diameter (D/dp) and low reactant velocity introduced a large heat transport resistance
between the tube wall and the catalyst particles. Mears et al. [23] developed a criterion to evaluate the
importance of radial temperature gradients in fixed bed catalytic reactors. Results showed that the
heat transfer resistance between the wall and the catalyst bed cannot be neglected when D/dp > 100.
Furthermore, the increased porosity of the bed near the wall caused a limited number of contact points
between catalyst particles and the reactor wall. Hence, the major cause of the heat transport resistance
between the catalyst bed and the reactor wall could be regarded as the gas film [24], which has been
considered in this study. Vázquez et al. [25] employed a tubular-quartz reactor and a multichannel
micro packed-bed reactor to perform the kinetic model of methanol steam reforming. The results
represented both axial and radial temperature gradients in the catalyst bed. But for a small-scale
multichannel reactor with a large length to width ratio (L/W = 50), the radial temperature gradients
in the catalyst bed can be considered negligible at an almost isothermal condition. The study proved
that convective heat transfer properties in the catalyst bed could be improved by increasing the length
to width ratio of reactors to increase the gas velocity and the contact surface of gas flow in reactors.
Montebelli et al. [26] compared the performance of two highly conductive structured multi-tubular
reactors with a commercial multi-tubular packed-bed reactor for methanol synthesis. They concluded
that the packed-bed reactor had a better performance than structured systems due to the effective
convective heat transfer mechanism in the catalyst bed, which is shown as lower hot-spot temperatures
and higher radial heat transfer rates. The effectiveness factor for catalysts in commercial size has been
widely investigated owing to the strong effect of internal diffusion on reaction rates. Lee et al. [27]
estimated the effectiveness factor of catalyst particles to investigate the effect of the particle internal
diffusion limitations and obtain the intrinsic kinetics of methanol steam reforming over Cu/ZnO/Al2 O3
catalyst. Tesser et al. [28] tested different kinetics of steam reforming of methanol in packed bed reactor,
considering both mass and heat balance along the length of reactor and inside the catalyst particles.
The multi-tubular packed-bed reformer is normally represented as an entire tube bundle immersed
in an external heating source with a uniform and constant temperature when developing a mathematical
model. However, in practical applications, the temperature of thermal fluid in inter-tubular space is
variable along the length and has significant effect on the performance of the reformer. Therefore,
the temperature profiles of both tube side and shell side along the reactor should be taken into
consideration. Compared to the large heat transfer resistance between the tube wall and the catalyst
bed, the convective and conductive heat transfer inside the catalyst bed is relatively efficient. Hence,
with a large L/W of the reactor, we took into account the heat transfer resistance of gas film, and neglected
the radial temperature gradient inside catalyst bed to simplify the model. The catalyst effectiveness
Energies 2020, 13, 610 4 of 25

factor is generally introduced for taking into account the internal diffusion resistance in commercial
catalyst particles, especially when large size particles are used. In this paper, the Weisz-Prater Criterion
was used to check if there were diffusion limitations, and the effectiveness factor was estimated for
the reaction.
In this work, a one-dimensional pseudo-homogenous model for multi-tubular packed-bed
reformer was established in MATLAB taking into account the main chemical reactions, and the mass
and heat transfer phenomena in both tube side and shell side. In the radial direction, the overall
heat transfer coefficient between catalyst bed and external heating source was considered, including
the conductive heat transfer through the tube wall, the convective heat transfer from reactant to
the inner wall of the tube, and the convective heat transfer from the outer tube wall to the outside
fluid. Effects of the pressure drop along the reactor and the intraparticle diffusion limitation were
also included in this model. A dynamic model of the multi-tubular methanol steam reformer was
developed in ASPEN to verify the MATLAB model. The thermal behavior of both tube side and
shell side was represented in the term of temperature profile. The influence of operating conditions
such as flow rate of methanol-and-water mixture and inlet temperature of external thermal air on the
methanol conversion and CO concentration of reformed gas was investigated. In addition, the impact
of geometric parameters of reactor design, such as the diameter and number of tubes as well as the
spacing and number of baffles, has been investigated on the reformer performance. This model of
MSR is expected to be integrated with the high temperature polymer electrolyte membrane fuel cell
(HT-PEMFC) in a combined stack arrangement to investigate the thermal integration of the system for
further study.

2. Description of Methanol Steam Reformer


Generally, a reformed methanol fuel cell (RMFC) system is composed of a burner, an evaporator,
several thermal fluid circuits, a methanol steam reformer and a fuel cell stack. Firstly, the mixture
of methanol and water is pumped into the evaporator where the fuel is evaporated. The vapor fuel
is then fed into the catalyst bed in the reformer and converted into hydrogen-rich gas by chemical
reactions. The hydrogen-rich gas from the reformer is sent to the anode side of the fuel cell stack.
The fuel cells utilize the hydrogen from anode side together with the oxygen from the cathode side to
generate electricity by electrochemical reactions. The exhaust gas from the fuel cell stack is directed to
the burner, where the residual hydrogen and methanol react with air in the burner. The generated heat
is transferred into the catalyst bed by thermal fluid circuit. In this work, the methanol steam reformer
as a subsystem of the RMFC system has been studied.
The reformer for MSR in this study is a multi-tubular packed-bed reformer. The structure of the
heat-exchanger type reformer, including baffles, tubes and a shell, is presented in Figure 1. The reactor
shell is usually surrounded by thermal insulation materials to avoid any significant amount of heat
loss. Tubes packed with Cu/ZnO/Al2 O3 catalyst are arranged in equilateral triangle tube bundles and
installed inside the shell. The baffle plates are used to support the tube bundles, increase the flow
distribution in the inter-tubular space and for an effective heat transfer between the tube side and
shell side.
The heating gas from the burner flows through the shell side of the reformer, thereby providing
an external heat source for the reactions. In the tube side, reactants flow through the catalyst bed,
where the steam reforming reactions occur.
The reformer for MSR in this study is a multi-tubular packed-bed reformer. The structure of the
heat-exchanger type reformer, including baffles, tubes and a shell, is presented in Figure 1. The
reactor shell is usually surrounded by thermal insulation materials to avoid any significant amount
of heat loss. Tubes packed with Cu/ZnO/Al2O3 catalyst are arranged in equilateral triangle tube
bundles and installed inside the shell. The baffle plates are used to support the tube bundles, increase
Energies 2020, 13, 610 5 of 25
the flow distribution in the inter-tubular space and for an effective heat transfer between the tube
side and shell side.

Figure 1.
Figure Structure of
1. Structure of the
the multi-tubular
multi-tubular packed-bed
packed-bed reformer
reformer for
for methanol
methanol steam
steam reforming.
reforming.
3. Model of the Methanol Steam Reformer

3.1. Kinetic Model of Reaction Rates


In this study, the process of methanol steam reforming consists of three reversible overall reactions:
MSR, WGS and MD. For kinetic analysis of the reforming process over Cu/ZnO/Al2 O3 catalyst,
the Langmuir-Hinshelwood macro kinetic model based on the study of Peppley et al. [10] is used.
This classic and comprehensive kinetic model is developed on the basis of several assumptions:
hydrogen and oxygen-containing species adsorb on different active sites; the active sites for MD
reaction are different from those for SRM and WGS reaction; the rate-determining step for both MSR
and MD is the dehydrogenation of the adsorbed methoxy, while for WGS reaction the RDS is the
formation of an intermediate formate species [12]. According to the Langmuir-Hinshelwood model,
the rate expressions for three key reactions involved in the process can be expressed as follows:
1
 
eq

kR K ∗ ( 1 )
pCH3 OH /p H
2
1 − p3H pCO2 /KR pCH3 OH pH2 O CTS CTS
CH3 O 2 2 1 1a
rR =   1
 1
 1
 1 1
, (4)
1 + K∗ (1)
pCH3 OH /pH + K 2 ∗ 2 ∗
p p + K (1) pH2 O /pH
(1) CO2 H
2 2 2
1 + K (1a) pH
CH3 O 2 HCOO 2 OH 2 H 2

1
 
eq
 2
kW K∗ (1)
p CO pH 2 O /pH
2
1 − pH2 pCO2 /KW pCO pH2 O CTS
OH 2 1
rW =  2 , (5)
1 1 1
  
1 + K∗ (1)
p CH 3 OH /pH
2
+ K∗ pCO2 pH + K∗ (1)
2
pH2 O /pH2
CH3 O 2 HCOO(1) 2 OH 2

1
 
eq

kD K∗ pCH3 OH /pH2 1 − p2H pCO /KD pCH3 OH CTS CTS
CH3 O(2) 2 2 2 2a
rD =   1
  1
 1 1
, (6)
1 + K∗ (2)
pCH3 OH /p 2
H
+ K ∗
(2)
pH2 O /pH
2
1 + K 2
p
(2a) H
2
CH3 O 2 OH 2 H 2

eq
where k j and Kj are the rate constant and equilibrium constant of reaction j (j =
R, W, D) respectively, K∗ is the adsorption coefficient, pi is the partial pressure of component
i (i = CO2 , CO, H2 , CH3 OH and H2 O), CTS , CTS , CTS and CTS are the total site concentrations of
1 1a 2 2a
site ‘10 , ‘1a’, ‘20 , and ‘2a’ respectively. The required parameters for the comprehensive kinetic model of
methanol steam reforming can be found in [10]. The temperature dependence of each constants can be
expressed using the Arrhenius expression [10,29,30]:

−ER
 

kR = kR exp , (7)
RT
Energies 2020, 13, 610 6 of 25

−ED
 

kD = kD exp , (8)
RT
−EW
 

kw = kR exp , (9)
RT
50240 − 170.98T − 2.64 × 10−2 T2
!
eq
KR = exp − , (10)
RT
−41735 + 46.66T − 7.55 × 10−3 T2
!
eq
KW = exp − , (11)
RT
eq
eq KR
KD = eq , (12)
KW

 ∆SCH3 O(1) ∆HCH3 O(1) 


 ∗ ∗ 
K∗ = exp − , (13)
CH3 O(1) R RT


 ∆SHCOO(1) ∆HHCOO(1) 
 ∗ ∗ 
K∗ = exp − , (14)
HCOO(1) R RT 

 ∆SOH(1) ∆HOH(1) 
 ∗ ∗ 

K (1) = exp  − , (15)
OH R RT 
∆SH(1a) ∆HH(1a)
!
KH(1a) = exp − , (16)
R RT

 ∆SCH3 O(2) ∆HCH3 O(2) 


 ∗ ∗ 
K∗ = exp − , (17)
CH3 O(2) R RT 

 ∆SOH(2) ∆HOH(2) 
 ∗ ∗ 
K ∗ (2)
= exp − , (18)
OH R RT 
∆SH(2a) ∆HH(2a)
!
KH(2a) = exp − , (19)
R RT

To calculate the rate ri mol s−1 (kg of catalyst)−1 of production i per time per mass of catalyst, it
 
 
is necessary to combine the rate expressions r j mol s−1 m−2 for each individual reaction j and multiply
 
by the surface area per unit mass of fresh catalyst Sc m2 kg−1 :

rCO2 = (rR + rW )Sc , (20)

rCO = (rD − rW )Sc , (21)

rH2 = (3rR + 2rD + rW )Sc , (22)

− rCH3 OH = (rR + rD )Sc , (23)

− rH2 O = (rR + rW )Sc , (24)

3.2. Pressure Drop in Catalyst Bed of the Packed-Bed Reformer


As a fluid passes through a packed bed, it experiences pressure loss. Especially when the size of
tablets is small, the pressure drop in the catalyst bed has to be taken into consideration. In this paper,
we assume that a set of porous cylindrical catalyst particles of uniform size are packed in cylindrical
Energies 2020, 13, 610 7 of 25

tubes. It is widely accepted that the pressure drop can be approximated by the semiempirical Ergun
equation [31]:
G 1 − φ 150(1 − φ)ηt
!" #
dP
=− + 1.75G , (25)
dz ρt Dp φ3 Dp

 bed, the viscosity of the gas mixture ηt (Pa s) as well as the density of the gas
Inside the packed
mixture ρt kg m−3 can be described as a function of reactor length z (m); P (Pa) is the pressure in
the catalyst bed; φ is the void fraction of the catalyst bed; Dp (m) is the diameter of catalyst particles;
 
G kg m−2 s−1 is the superficial mass velocity. The first-order solution for viscosity of pure gas ηi (Pa s)
can be expressed by using the Chun et al. method [32,33]:
1
Fc (Mi T ) 2
ηi = 40.785 2
, (26)
Vc3 Ωv

Fc = 1 − 0.2756ω + 0.059035µ4r + κ, (27)

κ = 0.682 + 4.704[(number o f − OH groups)/(molecular weight)], (28)


µ
µr = 131.3 1
, (29)
( Vc T c ) 2
T
T∗ = 1.2593 , (30)
Tc
Ωv = A(T∗ )−B + C[exp(−DT∗ )] + E[exp(−FT∗ )],
h i
(31)

where A = 1.16145, B = 0.14874, C = 0.52487, D = 0.77320, E = 2.16178 and F = 2.43787.


In this study, Herning and Zipperer’s method [34] was utilized to estimate the gas mixture
viscosity ηm (Pa s) in the methanol steam reformer:

5
X xi ηi
ηm = P5 , (32)
i=1 j=1 x j Φij

! 12
Mj
Φij = = Φ ji −1 , (33)
Mi
The properties of each gas component are listed in Table 1.

Table 1. Basic constants values for the calculation of viscosity of gases.

Gas Components Vc (cm3 mol−1 ) Tc (K) ω µ (D)


CH3 OH 118.00 512.64 0.565 1.70
H2 O 55.95 647.14 0.344 1.84
H2 64.20 32.98 −0.217 0
CO 93.10 132.85 0.045 0.122
CO2 94.07 304.12 0.225 0

3.3. Effectiveness Factor


The use of large catalyst size can reduce the pressure drop in the catalyst bed along the length,
but may lead to a significant effect of intraparticle diffusion limitation on reaction rates. The reforming
mixture in catalyst bed consists of five components: carbon dioxide (CO2 ), hydrogen (H2 ), water (H2 O),
methanol (CH3 OH), carbon monoxide (CO). In catalyst bed, reactants transfer from the bulk fluid
to the external surface of catalyst particle with reforming reactions taking place. Then the reactants
will diffuse from the external surface with a higher concentration to reach the pores surface inside
Energies 2020, 13, 610 8 of 25

the catalyst through the pore tortuosity. The effective diffusivity is defined to describe the diffusion,

which affects the chemical reactions inside catalyst particles. The effective diffusivity Di,eff m2 s−1 is
estimated using the Maxwell-Stefan equation [35]:
n
xj xi N j
!
1 X 1
= 1− + , (34)
Di,eff Dij x j Ni Di,K
j=1
j,i
 
where Dij m2 s−1 is the diffusivity for a binary mixture of i and j. The Chapman-Enskog equation is
used for the binary diffusivity at low density:
 1/2
T3/2 1/Mi + 1/M j
Dij = 0.0018583 , (35)
Pσ2ij ΩD

σi + σ j
σij = , (36)
2
σ = 0.841Vc1/3 , (37)
ε
= 0.75Tc , (38)
k
 0.10
ΩD = 44.54Tij∗−4.909 + 1.911Tij∗−1.575 , (39)

kT
Tij∗ = , (40)
εij
 1/2
εij = εi ε j , (41)
 
where T (K) is the operating temperature, P (atm) is the pressure, Mi kg mol−1 is the molar mass
 
of component i, ΩD is the collision integral for diffusion. The Knudsen diffusivity Di,K m2 s−1 of
component i is calculated by:
r
d 8RT
Di,K = , (42)
3 πMi
 
where d (m) is the average pore diameter 6.4 × 10−9 m [36].
The effectiveness factor η (ranging from 0 to 1) is defined to describe the relative importance of
diffusion and reaction limitations. For a first order reaction, the expression of the effectiveness factor is:

ηφ21 = 3(φ1 cothφ1 − 1), (43)

where φ1 is the Thiel modulus for a first-order reaction. The left-hand side is also called the Weisz-Prater
parameter CWP , which is used to determine whether the diffusion is limiting the reactions. The shape
of catalyst particles used in this study is cylindrical.

CWP = ηφ21 , (44)

−r js ρc r2p
φ21 = , (45)
4Di,eff Cis

where r js mol s−1 ( g o f catalyst)−1 and Cis mol m−3 are the rate of reaction j. and the concentration
   

of component i. if the entire interior surface were exposed; rp (m) is the radius of catalyst particle.
Energies 2020, 13, 610 9 of 25

If CWP  1, the diffusion limitation and concentration gradient within the catalyst particle is supposed
to be negligible. However, if CWP  1, the internal diffusion limitation should be considered.

3.4. Mass and Energy Balance in Tube Side and Shell Side of the Packed-Bed Reformer
In this study, a comprehensive mathematical model is developed to evaluate the performance
of the MSR process in the reformer. Tubes packed with catalyst particles were placed inside the
reformer shell. Chemical reactions take place on the surface of Cu/ZnO/Al2 O3 catalyst in the tube side
of reformer. There is also a sweep of heating gas through the shell side, where no reaction happens.
The model developed for the performance of the MSR process in the catalyst bed of one single tube can
be extended to all tubes. The transport phenomena in the reactor-heat supply system can be described
by the concentration of reactants and thermal profiles along the length of thereactor. In the radial
direction of the methanol steam reformer, the convective heat transfer on both surfaces of the tube
and the conductive heat transfer through the tube wall are considered. To simplify the mathematical
model, major assumptions can be listed as follows:

• steady-state conditions;
• ideal gas behavior;
• the radial mass and thermal dispersions in the tube side and shell side are negligible;
• the methanol steam reaction is regarded as a first-order reaction;
• the reformer-heat supply system is adiabatic from the surrounding.

In the catalyst bed, it is assumed that the reforming reactions are affected by internal diffusion of
reactants. The effectiveness factor is set to be η for the catalyst. Therefore, the continuity equations for
specie i is given by the following mole-balance equations:

dFi
= ηri ρc Ac , (46)
dz
   
where Fi mol·s−1 is the molar flow rate of component i, ρc kg·m−3 is the density of catalyst, Ac (m2 )
is the area of cross section of the catalyst bed.
A steady state energy balance along the axis of catalyst bed leads to the following equation [31]:

r j ∆H j Sc ρc
P
dTt Ut a∆T +
= P Ac , (47)
dz Fi Cpi
 
where Cpi J mol−1 K−1 is the specific heat of gas component i, which can be calculated from Table 2;
 
∆H j J mol−1 is the enthalpy change of reaction j, which can be calculated from Table 3; ∆T (K) is
the temperature difference between shell side and tube side at length z (m); a is the
 ratio of the heat
−2 −1 −1 −2 −1
transfer area inside the reactor to the reactor volume; Ut Wm K or Js m K is the overall heat
transfer coefficient of tube side.
A steady state energy balance along the axis of the shell side is written as:

dTs Nt Us Ao ∆T
= , (48)
dz Fs Cps L
   
where Us W m−2 K−1 or J s−1 m−2 K−1 is the overall heat transfer coefficient outside the tube, Ao m2
 
is the heat transfer area outside the reactor tube, Fs mol∆s−1 is the molar flow rate of heating air,
 
Cps J kg−1 K−1 is the specific heat of heating air, L (m) is the total length of the packed-bed reactor,
and Nt is the number of tubes in the reformer.
Energies 2020, 13, 610 10 of 25

Table 2. Specific heat of gases.

Species Specific Heat (J/mol·K)


T 2 T 3
       2
T
H2 CPH2 = a1 + b1 1000 + c1 1000 + d1 1000 + e1 1000
T ,
T 2 T 3 1000 2
       
H2 O T
CPH2 O = a2 + b2 1000 + c2 1000 + d2 1000 + e2 T ,
T 2 T 3 1000 2
       
CO2 T
CPCO2 = a3 + b3 1000 + c3 1000 + d3 1000 + e3 T ,
T 2 T 3 1000 2
       
CO T
CPCO = a4 + b4 1000 + c4 1000 + d4 1000 + e4 T ,
CH3 OH CPCH3 OH = 63.4.
a1 = 33.066178 a2 = 30.92000 a3 = 24.99735 a4 = 25.56759
b1 = −11.363417 b2 = 6.832514 b3 = 55.18696 b4 = 6.096130
c1 = 11.432816 c2 = 6.7934356 c3 = −33.69137 c4 = 4.054656
d1 = −2.772874 d2 = −2.534480 d3 = 7.948387 d4 = −2.671301
e1 = −0.158558 e2 = 0.0821398 e3 = −0.136638 e4 = 0.131021

Table 3. Enthalpy change of reactions.


 
∆HR = 4.95 × 104 + CPCO2 + 3CPH2 − CPCH3 OH − CPH2 O (T − 298) J/mol
 
∆HD = 9.07 × 104 + CPCO + 2CPH2 − CPCH3 OH (T − 298) J/mol
 
∆HW = −4.12 × 104 + CPCO2 + CPH2 − CPH2 O − CPCO (T − 298) J/mol
Energies 2019, 12, x FOR PEER REVIEW 10 of 26

Figure 2 shows the cross section of a single tube and the radial temperature profile near the tube
Figure 2 shows the cross section of a single tube and the radial temperature profile near the tube
wall. There are three regions, where the temperature varies sharply, corresponding to three resistances
wall. There are three regions, where the temperature varies sharply, corresponding to three
to heat transfer: (1) the fluid film of the inner side of the tube, (2) the tube wall and (3) the fluid film
resistances to heat transfer: (1) the fluid film of the inner side of the tube, (2) the tube wall and (3) the
outside the tube.
fluid film outside the tube.

Figure 2. Cross section of single tube and radial temperature profile.


Figure 2. Cross section of single tube and radial temperature profile.
Therefore, the overall driving force can be decomposed into the sum of three separate temperature
Therefore, the overall driving force can be decomposed into the sum of three separate
drops in each region:
temperature drops in each region:
∆T∆𝑇==Ts𝑇−−Tt𝑇==(T(𝑇 wo ) )++(T
s −−T𝑇 wi ) )++(T
(𝑇wo −−T𝑇 t ),),
(𝑇wi −−T𝑇 (49)
(49)
coefficients hℎt and
Film coefficients andhs ℎareare used
used to describe
to describe the convective
the convective heat-transfer
heat-transfer rate between
rate between the
the fluid
fluid flowtube
flow and andwall
tubeinwall in #1
#1 and #3and #3films
fluid fluidseparately.
films separately. The
The film film heat-transfer
heat-transfer ht of the ℎfluid
coefficient
coefficient of
the fluid film in tube side is defined as a proportionality constant between the heat flux 𝑑𝑞/𝑑𝐴 and
driving force (𝑇 − 𝑇 ):
𝑑𝑞 = ℎ 𝑑𝐴 (𝑇 − 𝑇 ), (50)
Similarly, the film heat-transfer coefficient of the fluid film in shell side ℎ is defined as:
𝑑𝑞 = ℎ 𝑑𝐴 (𝑇 − 𝑇 ), (51)
Energies 2020, 13, 610 11 of 25

film in tube side is defined as a proportionality constant between the heat flux dq/dAi and driving
force (Twi − Tt ):
dq = ht dAi (Twi − Tt ), (50)

Similarly, the film heat-transfer coefficient of the fluid film in shell side hs is defined as:

dq = hs dAo (Ts − Two ), (51)

The heat transfer occurs in the region #2 is pure conduction through the tube wall. This process
can be described as:
kw
dq = dA (Two − Twi ), (52)
xw lm
where xw (m) is the thickness of tube wall; Ai and Ao are the internal and external surface areas of the
tube wall separately; dAlm is the log-mean of dAi and dAo :

dAo − dAi π(Do − Di )dL


dAlm =   =   = πDlm dL, (53)
ln AAo ln D
D
o
i i

Assuming the steady-state heat transfer, the overall heat-transfer coefficient of the shell side Us
and of the tube side Ut can be described by [37]:

1 1 Do xw Do 1
= + + , (54)
Us ht Di kw Dlm hs

1 1 xw Di 1 Di
= + + , (55)
Ut ht kw Dlm hs Do
To calculate the overall heat-transfer coefficients, the heat transfer coefficient of the inner film ht
can be estimated as: 
1 2
 1 − φ km
ht = 0.4Re 2 + 0.2Re 3 Pr 0.4 , (56)
φ Dp
where Dp (m) is the diameter of the catalyst particle, φ is the void fraction, Pr is the Prandtl number,
 
km W m−1 K−1 is the average thermal conductivity of the gas mixture inside tubes.
For packed beds, the Reynolds number is defined by:

Dp G 1
Re = , (57)
ηm 1 − φ

For a conventional packed-bed reformer in the form of shell-and-tube heat exchanger, the outer
film coefficient hs is determined by the geometric parameters of reformer tubes and baffles inside the
shell, which is shown in Figure 3. The bundle of tubes in a heat-and-tube heat exchanger reformer can
be stacked in the triangular pitch which allows the tubes to be more tightly packed. The center-to-center
distance between adjacent tubes is called the tube pitch, pt (m). For tubes packed in triangular pitch,
the general requirement is that, pt ≥ 5/4 Di . A commonly used technique for increasing the heat
transfer coefficient hs is to install baffle plates inside the shell, which partially block the cross-sectional
area. The reduction of the cross-section area available for flow will increase the mass flux or velocity of
the flow, also can prevent the formation of large stagnant regions, and thereby enhance hs . The baffle
plate is a disc whose diameter is equal to the inner diameter of the reformer shell Ds (m) with holes for
tubes to pass through. The part of the baffle plate called baffle window is formed by cutting off this
part of plate to make a cross section available for the shell-side flow. Series of baffle plates are arranged
inside shell along the length of the reformer with the baffle window alternately placed on top and
bottom. The spacing between the baffle plates is baffle pitch, Pb (m). Typical baffle pitch is a fraction
block the cross-sectional area. The reduction of the cross-section area available for flow will increase
the mass flux or velocity of the flow, also can prevent the formation of large stagnant regions, and
thereby enhance ℎ . The baffle plate is a disc whose diameter is equal to the inner diameter of the
reformer shell 𝐷 (m) with holes for tubes to pass through. The part of the baffle plate called baffle
window
Energies is formed
2020, 13, 610 by cutting off this part of plate to make a cross section available for the shell-side
12 of 25
flow. Series of baffle plates are arranged inside shell along the length of the reformer with the baffle
window alternately placed on top and bottom. The spacing between the baffle plates is baffle pitch,
𝑃 (m).
of the shellTypical
diameter, 0.2 <
baffle pitch is sa<fraction
Pb /D 1. In theofdesign of reformer,
the shell 0.2 geometric
diameter,many 𝑃 /𝐷 1.parameters should
In the design of
reformer,
be taken intomany geometric
account parameters
with respect should be solution
to the optimum taken into
of account with respect
arrangements to the
of baffles andoptimum
tubes.
solution of arrangements of baffles and tubes.

Figure 3. Arrangement of tubes and baffles in the shell of reformer.


Figure 3. Arrangement of tubes and baffles in the shell of reformer.

A generally used equation for estimating the heat transfer coefficient in the shell-side film hs in a
A generally used equation for estimating the heat transfer coefficient in the shell-side film ℎ in
shell-tube heat exchanger is the Donohue equation:
a shell-tube heat exchanger is the Donohue equation:
𝐷e 𝐺0.6 . Cps𝐶ηs 𝜂0.33. ks𝑘
! !
Do G
hs = = 0.2
ℎ 0.2 ,, (58)
(58)
ηs 𝜂 ks 𝑘 Do𝐷
where mass velocity 𝐺 (kg m s  ) is calculated by:
where mass velocity Ge kg m−2 s−1 is calculated by:
𝐺 = 𝐺 𝐺 , (59)
q
Ge = Gb Gp , (59)
𝐺 (kg m s ) is the mass velocity through the baffle window:
 
Gb kg m−2 s−1 is the mass velocity through the baffle window:
.
ms
Gb = , (60)
Sb
 
where Sb m2 is the area available for shell-side fluid flow through the baffle window:

πD2s πD2o
! !
Sb = fb − Nb , (61)
4 4
 
Gp kg m−2 s−1 is the mass velocity for crossflow perpendicular to the tubes:
.
ms
Gp = , (62)
Sp
 
where Sp m2 is the interstitial area available for crossflow perpendicular to the bank of tubes at the
widest point in the shell: !
Do
Sp = Pb Ds 1 − , (63)
pt
  .  
where ks W m−1 K−1 is average thermal conductivity of shell-side gas, ms kg s−1 is the mass flow
rate of shell-side gas, ηs (Pa s) is the average viscosity of shell-side gas.
Energies 2020, 13, 610 13 of 25

4. Results and Discussion

4.1. Comparison betweeen Counter-Current and Co-Current Reactor


In a methanol steam reformer, the packed bed reactor for endothermic reactions is always coupled
with a combustor, which provides an external heating source by the flow of thermal air passing
through the adjacent shell side. When the reactant and thermal air flow in opposite directions,
the reactor is known as counter-current reactor. For co-current reactor, the flows are in the same
direction. The thermal performance of both co-current and counter-current reactors was simulated
with the inlet methanol-and-steam-mixture temperature of 433K and the inlet thermal air temperature
of 673K. Results of methanol conversion are both above 95%. The CO concentration in the exit gas
is 0.99% of the counter-current reactor and 0.34% of the co-current reactor. As Figure 4a represents,
the methanol and steam mixture is fed from the entrance of the reactor in a low temperature, and the
thermal air flows into the shell side from the opposite side of the reformer with a high temperature.
The small temperature difference in the “cold side” leads to a diminished heat transfer so that a lower
tube-side temperature. Therefore, the reaction rates of methanol reforming in the former part are
limited. In the “hot side” of reactor, the temperature of catalyst bed increases rapidly. The increased
operating temperature results to a high methanol conversion but also a higher CO concentration in the
exit gas. In Figure 4b, the temperature profiles show that the tube-side temperature increases sharply
after entering into the reactor. Due to the boosted endothermic reactions and reduced temperature
difference, the temperature of tube side decreases gradually after the peak temperature around 500 K.
The maximum temperature in co-current reactor is lower than counter-current reactor, which leads to a
better performance in CO control and less catalyst deactivation. Moreover, the larger temperature
difference drives a more efficient heat transfer, which benefits the higher reaction rate in the front part
Energies 2019, 12, x FOR PEER REVIEW 13 of 26
of the reactor. Therefore, co-current heat exchanger reformers are more favored.

Counter-current reactor Co-current reactor


700 700

Tube side Tube side


650 650
Shell side Shell side

600 600 Flow of thermal air


Temperature (K)

Temperature (K)

Flow of thermal air

550 550

500 500

450 450
Flow of reactants
Flow of reactants
400 400
0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
Reactor Length (m) Reactor Length (m)

(a) (b)
Figure
Figure 4. 4. Temperature
Temperature profiles
profiles ofof both
both tube
tube side
side and
and shell
shell side
side inin a (a)
a (a) counter-current
counter-current reactor,
reactor, and
and aa
(b)(b) co-currentreactor
co-current reactoralong
alongthe
thelength
lengthofofthe
thereactor.
reactor.

4.2.Pressure
4.2. Pressure Drop
Drop ininCatalyst
CatalystBedBed

It Itisisshown
shownthat thatthe
thesmaller-sized
smaller-sizedcatalyst
catalystperforms
performs better
better than
than larger-sized
larger-sized catalyst
catalyst and
and suffers
suffers
from less deactivation, owing to the increase in the surface area and number
from less deactivation, owing to the increase in the surface area and number of pore entrances causedof pore entrances caused
bybythethedecrease
decrease in in the
the particle
particle size
size[38].
[38].However,
However,inin small-scale packed
small-scale packedbedbed
reformer, especially
reformer, when
especially
smaller
when smaller size size
of catalyst tablet
of catalyst is chosen,
tablet the effect
is chosen, of pressure
the effect dropdrop
of pressure becomes significant
becomes andand
significant has has
to be
taken into consideration during the construction of the reactor. The effect of the
to be taken into consideration during the construction of the reactor. The effect of the particle size on particle size on the
pressure
the pressure along thethe
along catalyst bedbed
catalyst is investigated by changing
is investigated by changingthe particle diameter
the particle from 0.5
diameter frommm0.5tomm
2.0 mm,
to
which is shown in Figure 5. The particles packed in reformer tubes are nearly
2.0 mm, which is shown in Figure 5. The particles packed in reformer tubes are nearly monodispersed monodispersed in size
in size and cylindrical in shape. As expected, the pressure decreases along the length of catalyst bed,
and the decrease is favored at smaller particle sizes of catalyst. Additionally, the pressure drop is
calculated based on Ergun equation without taking into account the increased porosity of the bed
near the wall and the viscous friction at the wall. It is indicated that the value of actual pressure drop
may reach 20% higher than the pressure drop calculated by Ergun equation when the tube-to-
It is shown that the smaller-sized catalyst performs better than larger-sized catalyst and suffers
from less deactivation, owing to the increase in the surface area and number of pore entrances caused
by the decrease in the particle size [38]. However, in small-scale packed bed reformer, especially
when smaller size of catalyst tablet is chosen, the effect of pressure drop becomes significant and has
to be taken
Energies into
2020, 13, 610consideration during the construction of the reactor. The effect of the particle size 14 of on
25
the pressure along the catalyst bed is investigated by changing the particle diameter from 0.5 mm to
2.0 mm, which is shown in Figure 5. The particles packed in reformer tubes are nearly monodispersed
and cylindrical
in size in shape.
and cylindrical inAs expected,
shape. the pressure
As expected, decreases
the pressure along thealong
decreases lengththeoflength
catalyst
of bed, andbed,
catalyst the
decrease is favored at smaller particle sizes of catalyst. Additionally, the pressure
and the decrease is favored at smaller particle sizes of catalyst. Additionally, the pressure drop is drop is calculated
based on Ergun
calculated basedequation
on Ergun without taking
equation into account
without the increased
taking into account theporosity of theporosity
increased bed nearofthe
thewall
bed
and the viscous friction at the wall. It is indicated that the value of actual pressure drop
near the wall and the viscous friction at the wall. It is indicated that the value of actual pressure drop may reach 20%
higher
may reachthan the20%pressure
higher drop
than calculated
the pressure by Ergun equation when
drop calculated the tube-to-particle-diameter
by Ergun equation when the tube-to- ratio
D/d p is quite smallratio
particle-diameter [39]. 𝐷/𝑑
In thisisstudy, the particle
quite small size
[39]. In thisofstudy,
1.5 mmtheis particle
used. As recommended
size of 1.5 mm isby BASF
used. As
Catalyst Selectra ® , when the particle size of catalyst is 1.5 mm, the pressure drop should be taken into
recommended by BASF Catalyst Selectra®, when the particle size of catalyst is 1.5 mm, the pressure
consideration
drop should be during
takenthe
intodevelopment
consideration of during
the reformer model.
the development of the reformer model.

80

70 0.50 mm
0.75 mm
60 1.00 mm
Pressure Drop (mbar)

1.25 mm
50 1.50 mm
1.75 mm
40
2.00 mm

30

20

10

0
0.0 0.1 0.2 0.3 0.4 0.5
Reactor Length (m)

Figure 5. Pressure drop along the length of reactor with different particle sizes.
Figure 5. Pressure drop along the length of reactor with different particle sizes.
4.3. Model Verification and the Effectiveness Factor
4.3. The
Model Verification
reformer usedand
fortheMSR
Effectiveness Factor is a multi-tubular fixed bed reactor. The external
in this paper
thermal air in the shell side is used for providing heat source for reforming reactions inside tubes.
The overall heat transfer between shell side and tube side gas is similar to the case of shell-and-tube
heat exchanger. The geometrical parameters and operating conditions of the reformer are shown in
Table 4. The development of the reformer model is of high importance to predict the performance of the
small-scale methanol steam reformer and consequently optimize the design. Essentially, the simulation
model needs to be verified. Without taking into consideration of the effectiveness factor of reforming
reactions, the simulation model built in MATLAB-Simulink platform was verified using data from a
dynamic model of the methanol steam reformer in ASPEN. The comparison of methanol conversions
calculated by Simulink model and ASPEN model is shown in Table 5, where a good agreement is
be observed.

Table 4. Reactor specification and operating conditions of the reformer in this study.

Parameters Value

 Dp (mm
Catalyst particle diameter,  ) 1.5
Catalyst density, ρb kg m−3 1300
 
Surface area of fresh catalyst, Sc m2 kg−1 102,000
Void fraction φ 0.38
Operating temperature of catalyst (K) 503–563
Molar ratio of steam to methanol S/C 1.5
Pressure, P (bar) 1
Tube dimension (m) 0.016 × 0.001
Number of tubes, Nt 36
Inner diameter of reformer shell, Di (m) 0.24
Reactor length, L (m) 0.48
Energies 2020, 13, 610 15 of 25

Table 5. Comparison of the methanol conversions calculated by Simulink model and Aspen model.

Value
Parameters
Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Case 7 Case 8 Case 9
Mass flow rate of
4.5 6 7.5 4.5 4.5 4.5 4.5 4.5 4.5
liquid fuel (L/h)
Mass flow rate of
6 6 6 4 2 6 6 6 6
thermal air (g/s)
Inlet temperature of
433 433 433 433 433 453 413 433 433
vapor fuel (K)
Inlet temperature of
673 673 673 673 673 673 673 623 573
thermal air (K)
Result of Methanol Conversion
Simulink model (%) 95.34 80.03 66.75 76.92 45.74 96.63 93.85 81.52 62.71
ASPEN model (%) 95.18 79.10 65.71 77.21 46.38 96.42 93.78 81.31 62.81
Relative error (%) 0.17 1.16 1.56 −0.38 −1.40 0.22 0.07 0.26 −0.16

As the simulation results can be affected by internal diffusion of the catalyst particles,
the effectiveness factor should be taking into consideration. In this paper, the cylinder catalyst
particles with the size of 1.5 mm are used. To estimate the effectiveness factor by Equation (43), it is
necessary to check the value of Weisz-Prater parameter CWP to learn if the diffusion limitation is
significant within the catalyst particle. For both WGS and MD reactions, the value of CWP simulated
Energies 2019, 12, x FOR PEER REVIEW 15 of 26
under normal operating conditions is shown to be much smaller than 1. However, for the MSR
reactionsCWP
reaction, canranges from 0.5
be ignored andto it6.45. Therefore,
of the the intraparticle
MSR reaction should bediffusion limitation
considered in thisofmodel.
the WGS The
and MD reactions
effectiveness factorcan be ignored
profiles of MSR and it of the
reaction areMSR
shownreaction should
in Figure 6 as abe considered
function of theininlet
thisflow
model.
rate
The effectiveness
of fuel factor profiles
and the position of MSR
in the length ofreaction are shown
the catalyst bed. in Figure 6 as a function of the inlet flow
rate of fuel and the position in the length of the catalyst bed.

1.0

3.5 L/h
4.0 L/h
0.9
4.5 L/h
5.0 L/h
Effectiveness Factor

0.8

0.7

0.6

0.5
0.0 0.1 0.2 0.3 0.4 0.5
Reactor Length (m)

Figure 6. Effective factor of MSR reaction along the length of reactor with the inlet flow rate of fuel
ranging
Figure from 3.5 to 5.0
6. Effective L/h.of
factor The particle
MSR sizealong
reaction of catalyst is 1.5 mm.
the length of reactor with the inlet flow rate of fuel
ranging from 3.5 to 5.0 L/h. The particle size of catalyst is 1.5 mm.
4.4. Reformer Performance
4.4. The
Reformer Performance
developed mathematical model in this study allows predicting the mole fractions of different
speciesThe
exiting the reformer
developed mathematicalunder different
model conditions.
in this studyThe allows
mole fractions of different
predicting the mole species alongof
fractions
the length of
different the reactor
species arethe
exiting presented
reformerin Figure
under 7, with theconditions.
different inlet fuel flow
Therate
moleof methanol
fractions of and water
different
mixture at 4.5 L/h, the inlet temperature of thermal air at 673 K and the inlet temperature
species along the length of the reactor are presented in Figure 7, with the inlet fuel flow rate of vapor fuelof
atmethanol
433 K. The reforming process starts with only methanol and steam. As the reactions proceed
and water mixture at 4.5 L/h, the inlet temperature of thermal air at 673 K and the inlet along
the length of the
temperature of reactor, the at
vapor fuel mole
433fractions of methanol
K. The reforming and steam
process startsreduction,
with onlywhile the mole
methanol and fractions
steam. As
ofthe
reforming products including H , CO and CO increase continuously.
reactions proceed along the 2length of the2reactor, the mole fractions of methanol and steam
reduction, while the mole fractions of reforming products including H2, CO and CO2 increase
continuously.

70

60
methanol and water mixture at 4.5 L/h, the inlet temperature of thermal air at 673 K and the inlet
temperature of vapor fuel at 433 K. The reforming process starts with only methanol and steam. As
the reactions proceed along the length of the reactor, the mole fractions of methanol and steam
reduction, while the mole fractions of reforming products including H2, CO and CO2 increase
continuously.
Energies 2020, 13, 610 16 of 25

70

60

CH3OH
50
H2O

Mole Fraction (%)


H2
40
CO2
30 CO

20

10

0
0.0 0.1 0.2 0.3 0.4 0.5
Reactor Length (m)

Figure 7. Profiles of mole fraction of each components along the length of reactor.
Figure 7. Profiles of mole fraction of each components along the length of reactor.
In existing models of methanol steam reformer, the temperature of heating gas is usually considered
constant along themodels
In existing reactor of
length, so that
methanol a separate
steam energy
reformer, thebalance of the shell
temperature side can
of heating gasbeisignored
usually
toconsidered
simplify the mathematical model. In order to simulate the operating performance of the reformer
constant along the reactor length, so that a separate energy balance of the shell side can
more accurately,
be ignored the radial
to simplify theheat transfer and
mathematical thermal
model. In behavior of the shell
order to simulate theside have been
operating consideredof
performance
in this work. As shown in Figure 8a, the effect of fuel flow rate on axial temperature profiles in both
tube side and shell side was investigated. The flow rate of liquid water-and-methanol mixture varies
from 3 to 6 L/h, with the mass flow rate of thermal air at 6 g/s, the inlet temperature of vapor fuel and
thermal air at 433 K and 673 K, respectively. The temperature variation of catalyst bed inside tubes
is related to the endothermic reaction of methanol steam reforming and the external heat from the
shell side. With a lower flow rate of liquid fuel at 3 L/h, the tube-side temperature near the entrance of
the reactor increases more rapidly compared to the higher flow rate cases. Because less methanol is
participating in the reforming reactions, the energy absorbed by endothermic process is less than the
energy transferred to the tube side depending on the temperature difference. It is widely accepted that
the MSR process is favored at higher temperatures according to its thermodynamic characteristics.
Therefore, the enhanced endothermic reactions due to a higher tube-side temperature results in a
higher methanol conversion, which achieves nearly 100% at the exit of reactor as shown in Figure 8b.
When operating with higher flow rates of vapor fuel at 5 and 6 L/h, the tube-side temperatures
experience a momentary increase and then decrease along the length of reactor. With significant heat
consumption by reactions inside tubes, the radial heat transfer gradually becomes hard to compensate
for the heat loss because the drive force of temperature difference between the shell side and tube
side tends to be smaller. In addition, with a lower operating temperature in the catalyst bed and a
higher W/FCH3 OH (ratio of catalyst weight to molar flow rate of methanol, kg s/mol), the methanol
conversion with higher fuel flow rate is lower than other cases.
The effect of inlet temperature of the thermal air on the performance of reformer was investigated,
and the axial temperature profiles in both tube side and shell side were illustrate in Figure 9a.
The methanol conversion along the length of the reactor at different inlet temperature of the thermal air
is represented in Figure 9b. The inlet temperature of the thermal air varies in the range of 573−723 K
with the inlet temperature of vapor fuel at 433 K, the inlet flow rate of liquid fuel at 4.5 L/h, and the
inlet flow rate of the thermal air at 6 g/s. As can be seen from the temperature profiles, the increasing
temperature of the thermal air raises both the tube-side and the shell-side temperature along the
reactor length. There is a sharp and momentary increase of tube-side temperature near the entrance
of tube. In this part, the large driving force caused by temperature difference leads to a significant
heat transfer from the shell side to the tube side. After the tube-side temperature has been increased,
the increasing tendency slows down with the inlet temperature of thermal air at 723 K. It is known
that higher operating temperature favors MSR process, and hence with a higher inlet temperature of
profiles in both tube side and shell side was investigated. The flow rate of liquid water-and-methanol
mixture varies from 3 to 6 L/h, with the mass flow rate of thermal air at 6 g/s, the inlet temperature
of vapor fuel and thermal air at 433 K and 673 K, respectively. The temperature variation of catalyst
bed inside tubes is related to the endothermic reaction of methanol steam reforming and the external
heat from
Energies the610
2020, 13, shell side. With a lower flow rate of liquid fuel at 3 L/h, the tube-side temperature near
17 of 25
the entrance of the reactor increases more rapidly compared to the higher flow rate cases. Because
less methanol is participating in the reforming reactions, the energy absorbed by endothermic process
thermal air, the reaction rate of MSR is enhanced, which leads to more absorption of energy owing to
is less than the energy transferred to the tube side depending on the temperature difference. It is
the endothermic process. However, the driving force of radial heat transfer due to the temperature
widely accepted that the MSR process is favored at higher temperatures according to its
difference increases simultaneously. The enhanced heat transfer provides sufficient heat and has an
thermodynamic characteristics. Therefore, the enhanced endothermic reactions due to a higher tube-
overwhelming effect on the temperature distribution compared with the effect of heat absorption.
side temperature results in a higher methanol conversion, which achieves nearly 100% at the exit of
However, with the inlet temperature of thermal air lower than 673 K, the radial heat transfer is
reactor as shown in Figure 8b. When operating with higher flow rates of vapor fuel at 5 and 6 L/h,
diminished due to the reduce driving force of radial heat transfer. The endothermic process still
the tube-side temperatures experience a momentary increase and then decrease along the length of
plays the critical role compared with the effect of heat transfer. Therefore, after a sharp increase at
reactor. With significant heat consumption by reactions inside tubes, the radial heat transfer
the entrance of the reactor, the temperature subsequently decreases with a lower inlet temperature of
gradually becomes hard to compensate for the heat loss because the drive force of temperature
thermal air. Figure 9b represents the methanol conversion profile along the length of the reactor. With a
difference between the shell side and tube side tends to be smaller. In addition, with a lower operating
higher inlet temperature of thermal air, the methanol conversion increases due to the enhanced reaction
temperature in the catalyst bed and a higher W/F (ratio of catalyst weight to molar flow rate
rate. The methanol conversion achieves nearly 100% at the exit of reactor with the inlet temperature of
of methanol, kg s/mol), the methanol conversion with higher fuel flow rate is lower than other cases.
thermal air at 723 K.
700 700 100

Tube Side 3 L/h


650 650 4 L/h
5 L/h 80

Methanol Conversion (%)


6 L/h
Energies
600 2019, 12, x FOR PEER REVIEW
600 17 of 26
Temperature (K)

Temperature (K)

60

550 550
heat and has an overwhelming effect on the temperature40distribution compared with the effect of
3 L/h
heat 500
absorption. However, with 500 the inlet temperature of thermal air lower than 673 K, the 4 L/h
radial heat
20 5 L/h
transfer is diminished due to the reduce driving force of radial heat transfer. The endothermic 6 L/h
process
450 450 Shell Side
still plays the critical role compared with the effect of heat 0transfer. Therefore, after a sharp increase
0.0 0.1 0.2 0.3
at the entrance
0.0 0.1 0.2of 0.3
the0.4
reactor,
0.5 the
0.0 temperature
0.1 0.2 0.3 0.4 subsequently
0.5 decreases with a lower inlet0.4temperature
0.5

Reactor Length (m) Reactor Length (m) Reactor Length (m)


of thermal air. Figure 9b represents the methanol conversion profile along the length of the reactor.
With a higher inlet temperature (a) of thermal air, the methanol conversion (b) increases due to the
enhanced reaction rate. The methanol conversion achieves nearly 100% at the exit of reactor with the
Figure8.8.The
Theeffect
effectofofthethe massflowflow rateofofmethanol-and-water
methanol-and-watermixture
mixtureranging
rangingfromfrom33toto66L/h L/hon
on
inletFigure
temperature of thermalmass air at 723rate K.
(a)
(a)the
thetemperature
temperatureprofile
profileofofboth
bothtube
tubeside
sideand
andshell
shellside,
side,and
and(b)
(b)the
themethanol
methanolconversion
conversionalong
alongthe
the
length
lengthofofreactor.
reactor.
750 750 100
The effect of inlet temperature of the thermal air on the performance of reformer was
700 Tube Side 700 573 K
investigated, and the axial temperature profiles 623 K in both tube
80 side and shell side were illustrate in
Methanol Conversion (%)

650 650 673 K


Figure 9a. The methanol conversion along 723 theK length of the reactor at different inlet temperature of
Temperature (K)

Temperature (K)

60
the thermal
600 air is represented600in Figure 9b. The inlet temperature of the thermal air varies in the range
of 573−723
550
K with the inlet temperature 550
of vapor fuel at 433 K, the inlet flow rate of liquid fuel at 4.5
40
L/h, and the inlet flow rate of the thermal air at 6 g/s. As can be seen from the temperature 573 K profiles,
500 500
623 K
the increasing temperature of the thermal air raises both the 20 tube-side and the shell-side 673 temperature
K
450 450 723 K
along the reactor length. There isShell a sharp
Side and momentary increase of tube-side temperature near the
entrance
400 of tube. In this part,
400 the large driving force caused 0 by temperature difference leads to a
0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
significantReactor
heat Length
transfer
(m) from the shell Reactorside
Lengthto
(m)the tube side. After the tube-side temperature
Reactor Length (m) has been
increased, the increasing tendency slows down with the inlet temperature of thermal air at 723 K. It
(a) (b)
is known that higher operating temperature favors MSR process, and hence with a higher inlet
temperature
Figure9.9.of
Figure The
Thethermal
effectofair,
effect ofthe the
theinletreaction
inlet rate ofofof
temperature
temperature MSR is enhanced,
thermal
thermal air
airranging which
ranging fromleads
from 573
573tototo
723more
723 KKonabsorption
on (a) the of
(a)the
energy owing to
temperature
temperature the endothermic
profile
profile ofofboth
bothtube process.
tubeside
sideandandshellHowever,
shellside,
side,and the
and(b) driving
(b)the
themethanolforceconversion
methanol of radial along
conversion heat transfer
alongthe
thelength
lengthdue to
of reactor.
the temperature difference increases simultaneously. The enhanced heat transfer provides sufficient
of reactor.

For
Forthe
theapplication
applicationofofHT-PEMFCs,
HT-PEMFCs,the thecomposition
compositionofofproduced
producedHH 2 -rich
2-richgas
gasrequires
requiresa ahigh
high
concentration of
concentration of HH 2 2, and low concentrations of both methanol and CO. Araya et al. [40]has
, and low concentrations of both methanol and CO. Araya et al. [40] hasanalyzed
analyzed
the
theeffect
effectofofmethanol
methanoland andwater
watervapor
vaporon onthe
thedegradation
degradationofofHT-PEMFC
HT-PEMFCby bypolarization
polarizationcurves
curvesand
and
impedance spectra. The work showed that the HT-PEMFC operated with 5% and
impedance spectra. The work showed that the HT-PEMFC operated with 5% and 8% concentration 8% concentration of
methanol in the
of methanol inanode gas had
the anode gas significant performance
had significant degradation.
performance The results
degradation. also showed
The results that a 3%
also showed that
ora lower methanol concentration in the anode gas feed showed a negligible impact on
3% or lower methanol concentration in the anode gas feed showed a negligible impact on the the performance
performance of HT-PEMFC, which indicated that the HT-PEMFC had a tolerance of methanol
concentration of up to about 3%. Therefore, 2% methanol concentration in the reformed gas could be
a relatively low content to ensure the performance of the HT-PEMFC. With the acid-doped PBI
electrolytes, the poisoning effect of CO has been studied by Li et al. [41,42], and found that the effect
is very temperature-dependent and can be sufficiently suppressed at elevated temperature. The
Energies 2020, 13, 610 18 of 25

of HT-PEMFC, which indicated that the HT-PEMFC had a tolerance of methanol concentration of up to
about 3%. Therefore, 2% methanol concentration in the reformed gas could be a relatively low content
to ensure the performance of the HT-PEMFC. With the acid-doped PBI electrolytes, the poisoning effect
of CO has been studied by Li et al. [41,42], and found that the effect is very temperature-dependent
and can be sufficiently suppressed at elevated temperature. The results showed that 3% CO can be
tolerated when operating at 200 ◦ C, and for 1% above 175 ◦ C, and for 0.5% above 150 ◦ C. Similar results
have been achieved in [43] that when the HT-PEM fuel cell operated at 180 ◦ C or above, the reformate
gas with a higher CO concentration of 2−5% can be accepted. In addition, elevated concentration of CO
has been reported to cause unexpected coke deposition over the catalyst of methanol steam reformer
and covers the copper sites, which will have a negative effect on the reformer performance [36].
Energies 2019, 12, x FOR PEER REVIEW 18 of 26
The developed mathematical model in this study allows analyzing the mole fraction profiles of
residual methanol and CO
concentration. as a byproduct
As mentioned in themethanol
above, a higher reactantconcentration
mixture when due togave
lowertheinletoperating
temperature parameters
of heating
of the reformer. It isgas and a higher
especially flow rate of
important tofuel is unfavorable
predict to the performance
the concentration of HT-PEM
of different fuel cell
components in the
stack. In this paper, concentration lower than 2% is preferred to ensure a considerably low content of
exit gas which is fed into the fuel cell stack. The contours of methanol concentration in the exit gas
methanol.
of reformer areThe indicated in Figure
mole fraction profiles 10a
of COwith
in thedifferent
exit gas of inlet temperatures
the reformer of thermal
were illustrated in Figureair10b.and
As different
flow rates of liquid with
expected, fuel.increasing
It can beinlet
seen that a higher
temperature temperature
of thermal of thermal
air, the methanol air leads
decomposition to a lower
reaction is mole
promoted and
fraction of methanol inthethewater-gas
exit gasshift
of reaction is diminished,
the methanol steam which leads to Simultaneously,
reformer. a sharply increasing trendthe increasing
of mole fraction of CO. In addition, increasing the flow rate of fuel mixture leads to a decrease of the
flow rate of fuel mixture causes an obvious increase of methanol concentration. As mentioned above,
CO mole fraction. Likewise, a higher CO concentration due to lower flow rate of fuel and higher
a higher methanol
temperature concentration due tobelower
of heating gas should inlettotemperature
controlled a tolerable level.of heatingthe
Generally, gasCOand a higher flow rate
concentration
of fuel is unfavorable
lower than 1% toisthe performance
acceptable. of toHT-PEM
Therefore, ensure that fuel
thecell stack.composition
reformate In this paper, concentration
from the reformer lower
is suitable for an HT-PEMFC, the operating parameters
than 2% is preferred to ensure a considerably low content of methanol. represented by the lines in the respective
contours in Figure 10a,b are recommended.

Methanol Concentration in Exit Gas (%)


17.00
16.00
15.00
720
Inlet Temperature of Thermal Air (K)

14.00
13.00
12.00
11.00
10.00
700 2.0 9.000
8.000
7.000
6.000
5.000
4.000
680 3.000
2.000
1.000
0.000

660

640

620

1 2 3 4 5 6 7 8 9
Flow Rate of Liquid Fuel (L/h)

(a)

CO Concentration in Exit Gas (%)


6.500
6.000
720
Inlet Temperature of Thermal Air (K)

5.500
5.000
4.500
4.000
700 3.500
3.000
2.500
2.000
1.500
680 1.000
0.5000
0.000

660
1.0
640

620

1 2 3 4 5 6 7 8 9
Flow Rate of Liquid Fuel (L/h)

(b)
Figure 10. Profiles of (a) methanol concentration, and (b) CO concentration in the exit gas of reformer
Figure 10. Profiles of (a) methanol concentration, and (b) CO concentration in the exit gas of reformer
with different inlet flow rate of fuel mixture and inlet temperature of thermal air.
with different inlet flow rate of fuel mixture and inlet temperature of thermal air.
4.5. Influence of Geometric Parameters in the Multi-Tubular Packed-Bed Reactor
Energies 2020, 13, 610 19 of 25

The mole fraction profiles of CO in the exit gas of the reformer were illustrated in Figure 10b.
As expected, with increasing inlet temperature of thermal air, the methanol decomposition reaction is
promoted and the water-gas shift reaction is diminished, which leads to a sharply increasing trend of
mole fraction of CO. In addition, increasing the flow rate of fuel mixture leads to a decrease of the
CO mole fraction. Likewise, a higher CO concentration due to lower flow rate of fuel and higher
temperature of heating gas should be controlled to a tolerable level. Generally, the CO concentration
lower than 1% is acceptable. Therefore, to ensure that the reformate composition from the reformer is
suitable for an HT-PEMFC, the operating parameters represented by the lines in the respective contours
in Figure 10a,b are recommended.

4.5. Influence of Geometric Parameters in the Multi-Tubular Packed-Bed Reactor


It is widely accepted that the W/FCH3 OH associate with the dominating convective heat transfer
properties can strongly affect the performance of reformer. By keeping the overall weight of catalyst
W (kg) constant, which is substantially attained by setting the total volume of tubes invariant in the
reformer, the impact of geometric parameters can be investigated. With the fixed overall catalyst load,
the number of tubes will increase by a factor equal to the square of the factor by which the radius of
tubes decreases. For the packed-bed reformer in the current work, the geometric parameters used to
simulate the performance of the reformer are listed in Tables 6 and 7.

Table 6. Geometric parameters of the packed-bed reformer with variation in the number and diameter
of tubes.

Value
Parameters
Case 1 Case 2 Case 3 Case 4
Number of reactor tubes, Nt 144 36 16 9
Inner diameter of the tubular reactor, Di (m) 0.008 0.016 0.024 0.032
Outer diameter of the tubular reactor, Do (m) 0.010 0.018 0.026 0.034
Tube pitch, pt (m) 0.015 0.027 0.039 0.051
Number of baffle plates, Nb 4
Spacing between baffle plates, Pb (m) 0.12
Length of the reactor, L (m) 0.48
Area fraction of baffle plate that is window, fb
0.1955
(for 25% baffle plate)

Table 7. Geometric parameters of the packed-bed reformer with variation in the number and pitch
of baffles.

Value
Parameters
Case 1 Case 2 Case 3 Case 4
Number of baffle plates, Nb 3 4 6 8
Spacing between baffle plates, Pb (m) 0.16 0.12 0.08 0.06
Number of reactor tubes, Nt 16
Inner diameter of the tubular reactor, Di (m) 0.024
Outer diameter of the tubular reactor, Do (m) 0.026
Tube pitch, pt (m) 0.039
Length of the reactor, L (m) 0.48
Area fraction of baffle plate that is window, fb
0.1955
(for 25% baffle plate)

Axial temperature profiles in catalyst bed with the variations of tube number and tube diameter
in Table 6 are evaluated for the same mass of catalyst. The increase of tube diameter will lead to the
reduction of tube number in the reformer correspondingly, which boosts the overall heat exchange area
of reformer as well as the surface-to-volume ratio of a single cylindrical tube. As shown in Figure 11,
Energies 2020, 13, 610 20 of 25

tube-side temperatures have a sharp increase near the entrance of reformer owing to the driving force
of temperature difference between the internal and external fluid. After the initial increase, the reformer
Energies 2019, 12, x FOR PEER REVIEW 20 of 26
with less tubes and larger tube diameter (case 4) results to a roughly constant tube-side temperature
along the length of reactor. For the reformer with more tubes and smaller tube diameter (case 1),
interplay between the rates of heat transfer and endothermic reaction. The decrease of tube diameter
the tube-side temperature maintains the increasing trend until the maximum value, then progressively
promotes
Energiesthe
2019,heat
12, xtransfer,
FOR PEERwhichREVIEW favors the approach of catalyst bed to a higher temperature, hence 20 of 26
decreases along the length and reaches a final temperature at the exit of reformer even lower than the
increases the endothermic reaction rates. When the heat consumption of the endothermic reaction
other cases.between
interplay The temperature
the rates profiles in the and
catalyst bed are determined bydecrease
the interplay
of tubebetween
overcomes the heat supply fromof theheat transfer
external heatingendothermic reaction. The
source, the temperature will progressively diameter
drop.
the rates ofthe
promotes heat transfer
heat andwhich
transfer, endothermic
favors thereaction.
approach The
of decreasebed
catalyst of to
tube
a diameter
higher promoteshence
temperature, the
The variation of methanol conversion is directly related to the temperature profile in the catalyst
heat transfer, which favors the approach of catalyst bed to a higher temperature, hence increases the
bed.increases
As shownthe in endothermic
Figure 12a, the reaction
geometric rates. When theinheat
parameters caseconsumption
1 lead to higher of the endothermic
methanol reaction
conversion
endothermic
overcomes reaction rates. When the heat consumption of the endothermic reaction overcomes the
mainly becausethe heat
of its supply
short fromlength
pre-heat the external
and high heating source,
operating the temperature
temperature will progressively
in catalyst bed. However, drop.
heat supply
The from theofexternal
variation methanol heating source,isthe
conversion temperature
directly related will
to theprogressively
temperature drop.
profile in the catalyst
later the temperature drop limits the efficiency of MSR process and eventually results to an
bed.The
Asvariation
inconspicuously shown inofFigure
lower
methanol12a,conversion
methanol the geometric
conversion
is directly related tocase
at parameters
the outlet ofinthe
the temperature
1 lead toFigure
reformer.
profile
higher in the catalyst
methanol
12b shows the
bed.
conversion
CO
Asmainly
shownbecause
in Figure 12a, the
of itsinside geometric
short pre-heat parameters
length in case 1 lead to higher methanol conversion mainly
concentration profiles tubes along theand high of
length operating temperature
the reactor. in catalyst
Accordingly, the bed. However,
geometric
because of its short
later theoftemperature pre-heat length and high operating temperature in catalyst bed. However, latertothean
parameters more tubesdrop and limits
smaller thetube
efficiency of MSR
diameter processtoand
contribute eventually
a clear increase results
in CO
temperature
inconspicuously drop limits the efficiency of MSR process and eventually results to an inconspicuously
concentration due to lower methanol
the higher conversion
operating at the outlet of the reformer. Figure 12b shows the CO
temperature.
lower methanol profiles
concentration conversion insideat the outlet
tubes of the
along thereformer.
length ofFigure 12b shows
the reactor. the CO concentration
Accordingly, the geometric
profiles inside tubes along the length of the reactor. Accordingly,
parameters of more tubes and smaller tube diameter contribute to a clear increase the geometric parameters of more
in CO
tubes and smaller tube diameter
560 contribute
concentration due to the higher operating temperature. to a clear increase in CO concentration due to the higher
operating temperature. 540 Case 1
Case 2
Temperature of Tube Side (K)

Case 3
520 560
Case 4

500 540 Case 1


Case 2
Temperature of Tube Side (K)

480 Case 3
520
Case 4
460 500

440 480

420 460
0.0 0.1 0.2 0.3 0.4 0.5

440 Reactor Length (m)

420
0.0 0.1 0.2 0.3 0.4 0.5
Figure 11. Temperature profiles of tube side along theLength
Reactor length(m)of reactor with different number and
diameter of tubes.
Figure 11. Temperature profiles of tube side along the length of reactor with different number and
diameter of tubes.
Figure 11. Temperature profiles of tube side along the length of reactor with different number and
diameter
100 of tubes. 0.6

0.5
80
Methanol Conversion (%)

CO Concentration (%)

100 0.4 0.6


60

0.3 0.5
80
Methanol Conversion (%)

40
CO Concentration (%)

0.2 0.4
Case 1 Case 1
60
Case 2 Case 2
20 0.3 Case 3
Case 3 0.1
Case 4 Case 4
40
0 0.0 0.2 Case 1
Case 1
0.0 0.1 0.2 0.3 0.4 Case 0.5
2 0.0 0.1 0.2 0.3 0.4 Case0.5
2
20 Case 3
Reactor Length (m) Case 3 0.1 Reactor Length (m)
Case 4 Case 4

0 0.0
0.0 0.1 (a) 0.2 0.3 0.4 0.5 0.0 0.1 (b) 0.2 0.3 0.4 0.5
Reactor Length (m) Reactor Length (m)
Figure 12.12.
Figure Profiles of of
Profiles (a)(a)
methanol conversion,
methanol (b)(b)
conversion, COCO
concentration along
concentration thethe
along length of of
length reactor with
reactor with
different number
different numberandanddiameter of of
diameter tubes.
tubes.
(a) (b)
The Figure
number 12. of tubesofand
Profiles welding conversion,
(a) methanol operations(b)of CO
the concentration
reformer, due to the
along decrease
the length in the with
of reactor tube
diameterdifferent
at fixednumber
catalystand diameter
load, of tubes.the investment and industrial costs. In addition, the CO
will increase
concentration of the outlet gas mixture is also higher in this case. Therefore, there is a tendency to
Thethe
investigate number of tubes
possibility of and welding
adopting operations
reactor of theareformer,
tubes with due to the
larger diameter decrease
[44]. However,in the
thetube
adoption of reformer with larger tube diameter may cause larger radial temperatures gradients and CO
diameter at fixed catalyst load, will increase the investment and industrial costs. In addition, the
Energies 2020, 13, 610 21 of 25

The number of tubes and welding operations of the reformer, due to the decrease in the tube
diameter at xfixed
Energies 2019, 12, catalyst
FOR PEER load, will increase the investment and industrial costs. In addition,
REVIEW 21 the
of 26CO
concentration of the outlet gas mixture is also higher in this case. Therefore, there is a tendency to
cold-spot inside
investigate thethe catalystof
possibility bed [22], which
adopting means
reactor tubesthat
withthe heat supplied
a larger diameter from external thermal
[44]. However, air
the adoption
cannot effectively
of reformer withcompensate
larger tubethe heat consumed
diameter may cause bylarger
the reforming process, andgradients
radial temperatures this resultsandincold-spot
lower
methanol
inside the conversion.
catalyst bed Furthermore,
[22], whichfor tubes
means with
that the smaller diameter,
heat supplied fromespecially for ratiosair𝐷/𝑑𝑝
external thermal cannot
smaller than compensate
effectively 6, the impactthe ofheat
the porosity
consumed change
by thenear the tube
reforming wallsand
process, cannot
this be neglected
results in lower then [45].
methanol
Therefore,
conversion. the choice of geometric
Furthermore, parameters
for tubes should
with smaller be a trade-off
diameter, between
especially reformer
for ratios D/dpperformance
smaller than 6,
andthemanufacturing costs. change near the tube walls cannot be neglected then [45]. Therefore, the choice
impact of the porosity
of One of theparameters
geometric methods toshouldimprove be the heat transfer
a trade-off between properties
reformerof the reformer
performance with
and fixed flow rate
manufacturing costs.
of thermal air is to increase the external heat transfer coefficient by installing
One of the methods to improve the heat transfer properties of the reformer with fixed flow rate of baffle plates by
increasing
thermal air turbulence
is to increasein shell-side
the external fluid. The axial
heat transfer profiles by
coefficient ofinstalling
methanolbaffle conversion
plates byand CO
increasing
concentration
turbulence in inside tubesfluid.
shell-side are shown
The axial in profiles
Figure 13. With the conversion
of methanol constant length and CO of concentration
tubes, the added inside
number
tubes areof baffle
shownplates is associated
in Figure 13. With the with the reduction
constant length ofoftubes,
bafflethe
spacing. For the reformer
added number with is
of baffle plates
smaller bafflewith
associated spacing and moreof
the reduction baffle
baffleplates (caseFor
spacing. 4),the
a higher methanol
reformer conversion
with smaller baffleisspacing
attainedandwith
more
an baffle
increase in CO concentration at reformer outlet. This change can be explained
plates (case 4), a higher methanol conversion is attained with an increase in CO concentration by the increased
operating
at reformertemperature
outlet. Thisin catalyst
change bed can due to the effective
be explained by theradial heat operating
increased transfer oftemperature
reformer. However,
in catalyst
thebed
larger
duenumber of baffles
to the effective in the
radial shell-side
heat transferofofreformer
reformer.makes the fabrication
However, the larger more
number expensive.
of baffles in the
shell-side of reformer makes the fabrication more expensive.

100 0.6

0.5
80
Methanol Conversion (%)

O Concentration (%)

0.4
60

0.3

40 Case 1
Case 2 0.2 Case 1
Case 3 Case 2
20 Case 4 Case 3
0.1
Case 4

0 0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
Reactor Length (m) Reactor Length (m)

(a) (b)
Figure 13.13.
Figure Profiles of (a)
Profiles methanol
of (a) methanolconversion, (b)(b)
conversion, COCO
concentration along
concentration thethe
along length of reactor
length with
of reactor with
different number
different numberandandpitch of of
pitch baffles.
baffles.

5. Conclusions
5. Conclusions
Simulation
Simulation based
based onona acomprehensive
comprehensivepseudo-homogenous
pseudo-homogenous model model of of aa conventional
conventionalpacked-bed
packed-
reformer has been developed to investigate the effects of operating
bed reformer has been developed to investigate the effects of operating conditionsconditions and geometric parameters
and geometric
on temperature
parameters distribution,
on temperature methanol
distribution, conversion
methanol and CO
conversion andconcentration
CO concentration alongalong
the axis of the
the axis
multi-tubular methanol steam reformer. The model took into account
of the multi-tubular methanol steam reformer. The model took into account the main chemicalthe main chemical reactions,
and theand
reactions, masstheand
massheat
andtransfer phenomena
heat transfer in both
phenomena intube
bothside
tubeand
sideshell side of
and shell the
side ofreformer. In the
the reformer.
In radial direction,
the radial the the
direction, overall heat
overall transfer
heat coefficient
transfer coefficientincluding
includingthethe
convective
convective heat
heattransfer
transfernear
nearthe
internal and external surfaces of the tube wall, and the conductive heat transfer
the internal and external surfaces of the tube wall, and the conductive heat transfer through the tubethrough the tube wall
has been considered. The Ergun equation was used to calculate the pressure drop
wall has been considered. The Ergun equation was used to calculate the pressure drop inside catalyst inside catalyst bed.
A dynamic
bed. A dynamic model
model ofofthe
themulti-tubular
multi-tubularmethanol
methanolsteamsteamreformer
reformerwaswasdeveloped
developedin in the
the platform
platform of
ASPEN. Good agreement of methanol conversion was achieved between
of ASPEN. Good agreement of methanol conversion was achieved between the data of the MATLAB the data of the MATLAB
model
model andand
thethe ASPEN
ASPEN model.
model. TheThe performance
performance of aofcounter-current
a counter-current reactor
reactor and and a co-current
a co-current reactor
reactor
has been compared. Results showed that the reactor in the form of a co-current heat exchanger had had
has been compared. Results showed that the reactor in the form of a co-current heat exchanger a
a lower
lower CO concentration
CO concentration and better
and better heat transfer
heat transfer efficiency.
efficiency. The intraparticle
The intraparticle diffusion
diffusion limitation
limitation for
MSR reaction was considered by taking into account the effectiveness factor along the reactor. Axial
temperature profiles of both tube side and shell side with different flow rate of liquid fuel and
different inlet temperature of thermal air were represented in this paper. The developed
mathematical model also allowed analyzing the mole fractions along the length of reformer of
Energies 2020, 13, 610 22 of 25

for MSR reaction was considered by taking into account the effectiveness factor along the reactor.
Axial temperature profiles of both tube side and shell side with different flow rate of liquid fuel and
different inlet temperature of thermal air were represented in this paper. The developed mathematical
model also allowed analyzing the mole fractions along the length of reformer of residual methanol
and generated CO in the gas mixture. The results revealed that lower flow rate of liquid fuel and
higher inlet temperature of thermal air led to a better methanol conversion, but also a higher CO
concentration in outlet gas mixture. Because the generate hydrogen through steam reforming is
intended for use in HT-PEMFCs, the composition of the H2 -rich gas requires a concentration of
methanol lower than 2% and CO lower than 1%. Therefore, the operating conditions should be limited
to a certain region, where a suitable composition of reformed gas for feeding into an HT-PEM fuel
cell stack can be achieve. With fixed catalyst load, the increase in the number of tubes and decrease
in the tube diameter improved the methanol conversion, while also increasing the CO concentration.
In addition, the reformer with more baffle plates in the shell side achieved better methanol conversion
at the cost of higher CO concentration. Therefore, the choice of geometric parameters should be a
trade-off between the manufacturing costs and the reformer performance.

Author Contributions: Conceptualization, J.Z., S.S.A. and S.L.S.; data curation, J.Z.; formal analysis, J.Z. and
S.S.A.; funding acquisition, S.K.K.; investigation, J.Z. and X.C.; methodology, J.Z., S.S.A. and S.L.S.; project
administration, S.S.A. and S.K.K.; resources, S.K.K.; software, J.Z., S.L.S. and X.C.; supervision, S.S.A. and S.K.K.;
validation, J.Z., S.S.A. and X.C.; visualization, J.Z.; writing, original draft, J.Z.; writing, review and editing, J.Z.,
S.S.A., X.C., S.L.S. and S.K.K. All authors have read and agreed to the published version of the manuscript.
Funding: The research leading to these results has received funding from the Chinese scholarship council
(CSC) and from the Danish Energy Technology Development and Demonstration Program (EUDP) through the
Commercial Breakthrough of Advanced fuel cells (COBRA Drive) project, grant number 64018-0118.
Conflicts of Interest: The authors declare no conflict of interest.

Nomenclature
 
a ratio of the heat transfer area inside the reactor to the reactor volume m−1
 
Ac area of cross section of catalyst bed m2
 
Ai internal surface area of tube m2
 
Ao external surface area of tube m2
 
Cpi specific heat of gas component i Jmol−1 K−1
 
CTS total surface concentration of site i mol m−2
i  
Cis surface concentration of component i mol m−3
CWP Weisz-Prater parameter
Dp diameter of the catalyst particle in the reformer (m)
Ds diameter of the reformer shell (m)
Di inner diameter of tube (m)  
Dij diffusivity for a binary mixture of i and j m2 s−1
 
Di,K Knudsen diffusivity of component i m2 s−1
 
Di,eff effective diffusivity m2 s−1
Do outer diameter of tube (m)
fb area fraction of baffle plate that is window (for 25% baffle plate, fb = 0.1955)
Ej activation energy for rate constant of reaction j kJmol−1
 
Fi molar flow rate of component i mol s−1
 
G superficial mass velocity m s−1
Ge mass velocity of shell-side fluid  
Gb mass velocity through the baffle window kg m−2 s−1
 
Gp mass velocity for crossflow perpendicular to the tubes kg m−2 s−1
 
h individual heat transfer coefficient Wm−2 K−1 or Js−1 m−2 K−1
 
k thermal conductivity Wm−1 K−1
Energies 2020, 13, 610 23 of 25

 
kj rate constant of reaction j ( j = R, D or W) m2 s−1 mol−1
 
Ki∗ adsorption coefficient of specie i bar−0.5
eq
Kj equilibrium constant of reaction j ( j = R, D or W)
L length of the packed-bed
 reactor
 (m)
M molecular weight kg mol−1
Nt number of tubes
Nb number of baffles
pi partial pressure of component i (bar)
P pressure of the catalyst bed (bar)
Pb baffle spacing (m)
Pr Prandtl number
pt tube pitch (m)
rate of formation of component i mol s−1 (kg of catalyst)−1 , the number of moles of i
 
ri
reacting per unit time per unit mass  of catalyst
rj rate of reaction j ( j = R, D or W) mol s−1 m−2 , the number of moles of reaction per time
per unit surface
 area 
R gas constant Jmol−1 K−1
Re Reynolds number  
Sb area available for shell-side fluid flow through the baffle window m2
 
Sc surface area per unit mass of fresh catalyst m2 kg−1
Sp interstitial area available
  for crossflow perpendicular to the bank of tubes at the widest
point in the shell m2
 
U overall heat transfer coefficient Wm−2 K−1 or Js−1 m−2 K−1
 
Vc critical volume cm3 mol−1
W weight of catalyst (kg)
xi mole fraction of specie i
xw thickness of the tube wall (m)
µr dimensionless dipole moment (D)
ηi viscosity of fluid (Pa s)
η effectiveness
 factor

ρ density kg m−3
φ void fraction
φ1 Thiel modulus for a first-order reaction
ω acentric factor
κ special correctionfor highly
 polar substances
∆H enthalpy change Jmol−1
 
∆Si entropy change of adsorption for species i Jmol−1 K−1
∆T temperature difference (K)
Subscripts
m gas mixture
c catalyst
t tube side of the reformer
s shell side of the reformer
wi inner side of tube wall
wo outer side of tube wall
w tube wall

References
1. Edlund, D. Methanol Fuel Cell Systems: Advancing towards Commercialization; PAN: Stanford, CA, USA, 2016;
ISBN 9814303143.
2. Iulianelli, A.; Ribeirinha, P.; Mendes, A.; Basile, A. Methanol steam reforming for hydrogen generation via
conventional and membrane reactors: A review. Renew. Sustain. Energy Rev. 2014, 29, 355–368. [CrossRef]
Energies 2020, 13, 610 24 of 25

3. Karim, A.; Bravo, J.; Gorm, D.; Conant, T.; Datye, A. Comparison of wall-coated and packed-bed reactors for
steam reforming of methanol. Catal. Today 2005, 110, 86–91. [CrossRef]
4. Dalena, F.; Senatore, A.; Basile, M.; Knani, S.; Basile, A.; Iulianelli, A. Advances in methanol production
and utilization, with particular emphasis toward hydrogen generation via membrane reactor technology.
Membranes 2018, 8, 98. [CrossRef] [PubMed]
5. Sá, S.; Silva, H.; Brandão, L.; Sousa, J.M.; Mendes, A. Catalysts for methanol steam reforming-A review.
Appl. Catal. B Environ. 2010, 99, 43–57. [CrossRef]
6. Araya, S.S.; Zhou, F.; Liso, V.; Sahlin, S.L.; Vang, J.R.; Thomas, S.; Gao, X.; Jeppesen, C.; Kær, S.K.
A comprehensive review of PBI-based high temperature PEM fuel cells. Int. J. Hydrogen Energy 2016, 41,
21310–21344. [CrossRef]
7. Zhou, F.; Andreasen, S.J.; Kær, S.K.; Park, J.O. Experimental investigation of carbon monoxide poisoning effect
on a PBI/H3PO4 high temperature polymer electrolyte membrane fuel cell: Influence of anode humidification
and carbon dioxide. Int. J. Hydrogen Energy 2015, 40, 14932–14941. [CrossRef]
8. Papavasiliou, J.; Avgouropoulos, G.; Ioannides, T. Steady-state isotopic transient kinetic analysis of steam
reforming of methanol over Cu-based catalysts. Appl. Catal. B Environ. 2009, 88, 490–496. [CrossRef]
9. Jiang, C.J.; Trimm, D.L.; Wainwright, M.S.; Cant, N.W. Kinetic study of steam reforming of methanol over
copper-based catalysts. Appl. Catal. A Gen. 1993, 93, 245–255. [CrossRef]
10. Peppley, B.A.; Amphlett, J.C.; Kearns, L.M.; Mann, R.F. Methanol-steam reforming on Cu/ZnO/Al2 O3
catalysts. Part 2. A comprehensive kinetic model. Appl. Catal. A Gen. 1999, 179, 31–49. [CrossRef]
11. Agrell, J.; Birgersson, H.; Boutonnet, M. Steam reforming of methanol over a Cu/ZnO/Al2 O3 catalyst:
A kinetic analysis and strategies for suppression of CO formation. J. Power Sources 2002, 106, 249–257.
[CrossRef]
12. Sá, S.; Sousa, J.M.; Mendes, A. Steam reforming of methanol over a CuO/ZnO/Al2 O3 catalyst, part I: Kinetic
modelling. Chem. Eng. Sci. 2011, 66, 4913–4921. [CrossRef]
13. Herdem, M.S.; Mundhwa, M.; Farhad, S.; Hamdullahpur, F. Multiphysics Modeling and Heat Distribution
Study in a Catalytic Microchannel Methanol Steam Reformer. Energy Fuels 2018, 32, 7220–7234. [CrossRef]
14. Lee, M.T.; Greif, R.; Grigoropoulos, C.P.; Park, H.G.; Hsu, F.K. Transport in packed-bed and wall-coated
steam-methanol reformers. J. Power Sources 2007, 166, 194–201. [CrossRef]
15. Bravo, J.; Karim, A.; Conant, T.; Lopez, G.P.; Datye, A. Wall coating of a CuO/ZnO/Al2O3 methanol steam
reforming catalyst for micro-channel reformers. Chem. Eng. J. 2004, 101, 113–121. [CrossRef]
16. Cui, X.; Kær, S.K. Two-dimensional thermal analysis of radial heat transfer of monoliths in small-scale steam
methane reforming. Int. J. Hydrogen Energy 2018, 43, 11952–11968. [CrossRef]
17. Yoshida, K.; Tanaka, S.; Hiraki, H.; Esashi, M. A micro fuel reformer integrated with a combustor and a
microchannel evaporator. J. Micromech. Microeng. 2006, 16, S191. [CrossRef]
18. Chein, R.Y.; Chen, Y.C.; Lin, Y.S.; Chung, J.N. Experimental study on the hydrogen production of integrated
methanol-steam reforming reactors for PEM fuel cells. Int. J. Therm. Sci. 2011, 50, 1253–1262. [CrossRef]
19. Kim, T. Micro methanol reformer combined with a catalytic combustor for a PEM fuel cell. Int. J. Hydrogen
Energy 2009, 34, 6790–6798. [CrossRef]
20. Davieau, D.D.; Erickson, P.A. The effect of geometry on reactor performance in the steam-reformation process.
Int. J. Hydrogen Energy 2007, 32, 1192–1200. [CrossRef]
21. Vadlamudi, V.K.; Palanki, S. Modeling and analysis of miniaturized methanol reformer for fuel cell powered
mobile applications. Int. J. Hydrogen Energy 2011, 36, 3364–3370. [CrossRef]
22. Ma, H.; Zhou, M.; Ying, W.; Fang, D. Two-dimensional modeling of a plant-scale fixed-bed reactor for
hydrogen production from methanol steam reforming. Int. J. Hydrogen Energy 2016, 41, 16932–16943.
[CrossRef]
23. Mears, D.E. Diagnostic criteria for heat transport limitations in fixed bed reactors. J. Catal. 1971, 20, 127–131.
[CrossRef]
24. Karim, A.; Bravo, J.; Datye, A. Nonisothermality in packed bed reactors for steam reforming of methanol.
Appl. Catal. A Gen. 2005, 282, 101–109. [CrossRef]
25. Vidal Vázquez, F.; Simell, P.; Pennanen, J.; Lehtonen, J. Reactor design and catalysts testing for hydrogen
production by methanol steam reforming for fuel cells applications. Int. J. Hydrogen Energy 2016, 41, 924–935.
[CrossRef]
Energies 2020, 13, 610 25 of 25

26. Montebelli, A.; Visconti, C.G.; Groppi, G.; Tronconi, E.; Ferreira, C.; Kohler, S. Enabling small-scale methanol
synthesis reactors through the adoption of highly conductive structured catalysts. Catal. Today 2013, 215,
176–185. [CrossRef]
27. Lee, J.K.; Ko, J.B.; Kim, D.H. Methanol steam reforming over Cu/ZnO/Al2 O3 catalyst: Kinetics and
effectiveness factor. Appl. Catal. A Gen. 2004, 278, 25–35. [CrossRef]
28. Tesser, R.; Di Serio, M.; Santacesaria, E. Methanol steam reforming: A comparison of different kinetics in the
simulation of a packed bed reactor. Chem. Eng. J. 2009, 154, 69–75. [CrossRef]
29. Wan, Y.; Zhou, Z.; Cheng, Z. Hydrogen production from steam reforming of methanol over CuO/ZnO/Al2 O3
catalysts: Catalytic performance and kinetic modeling. Chin. J. Chem. Eng. 2016, 24, 1186–1194. [CrossRef]
30. Yaws, C.L. Chemical Properties Handbook; McGraw-Hill: New York, NY, USA, 1999; ISBN 0070734011.
31. Scott Fogler, H. Elements of Chemical Reaction Engineering; Prentice-Hall International London: London, UK,
1987; Volume 42, ISBN 8120322347.
32. Chung, T.H.; Ajlan, M.; Lee, L.L.; Starling, K.E. Generalized multiparameter correlation for nonpolar and
polar fluid transport properties. Ind. Eng. Chem. Res. 2005, 27, 671–679. [CrossRef]
33. Chung, T.H.; Lee, L.L.; Starting, K.E. Applications of Kinetic Gas Theories and Multiparameter Correlation
for Prediction of Dilute Gas Viscosity and Thermal Conductivity. Ind. Eng. Chem. Fundam. 1984, 23, 8–13.
[CrossRef]
34. Poling, B.E.; Prausnitz, J.M.; O’connell, J.P. The Properties of Gases and Liquids; Mcgraw-Hill: New York, NY,
USA, 2001; Volume 5.
35. Taylor, R.; Krishna, R. Multicomponent Mass Transfer; John Wiley & Sons: Hoboken, NJ, USA, 1993; Volume 2,
ISBN 0471574171.
36. Agarwal, V.; Patel, S.; Pant, K.K. H2 production by steam reforming of methanol over Cu/ZnO/Al2O3catalysts:
Transient deactivation kinetics modeling. Appl. Catal. A Gen. 2005, 279, 155–164. [CrossRef]
37. McCabe, W.L.; Smith, J.C.; Harriott, P. Unit Operations of Chemical Engineering, 6th ed.; McGraw-Hill: New
York, NY, USA, 2001; Volume 1130.
38. Mochizuki, H.; Yokoi, T.; Imai, H.; Watanabe, R.; Namba, S.; Kondo, J.N.; Tatsumi, T. Facile control of
crystallite size of ZSM-5 catalyst for cracking of hexane. Microporous Mesoporous Mater. 2011, 145, 165–171.
[CrossRef]
39. Winterberg, M.; Tsotsas, E. Impact of tube-to-particle-diameter ratio on pressure drop in packed beds.
AIChE J. 2000, 46, 1084–1088. [CrossRef]
40. Araya, S.S.; Grigoras, I.F.; Zhou, F.; Andreasen, S.J.; Kær, S.K. Performance and endurance of a high
temperature PEM fuel cell operated on methanol reformate. Int. J. Hydrogen Energy 2014, 39, 18343–18350.
[CrossRef]
41. Li, Q.; He, R.; Gao, J.A.; Jensen, J.O.; Bjerrum, N.J. The CO poisoning effect in PEMFCs operational at
temperatures up to 200 ◦ C. J. Electrochem. Soc. 2003, 150, 1599–1605. [CrossRef]
42. Li, Q.; Oluf, J.; Savinell, R.F.; Bjerrum, N.J. High temperature proton exchange membranes based on
polybenzimidazoles for fuel cells. Prog. Polym. Sci. 2009, 34, 449–477. [CrossRef]
43. Das, S.K.; Reis, A.; Berry, K.J. Experimental evaluation of CO poisoning on the performance of a high
temperature proton exchange membrane fuel cell. J. Power Sources 2009, 193, 691–698. [CrossRef]
44. Montebelli, A.; Visconti, C.G.; Groppi, G.; Tronconi, E.; Kohler, S. Optimization of compact multitubular
fixed-bed reactors for the methanol synthesis loaded with highly conductive structured catalysts. Chem. Eng.
J. 2014, 255, 257–265. [CrossRef]
45. Arab, S.; Commenge, J.M.; Portha, J.F.; Falk, L. Methanol synthesis from CO2 and H2 in multi-tubular
fixed-bed reactor and multi-tubular reactor filled with monoliths. Chem. Eng. Res. Des. 2014, 92, 2598–2608.
[CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like