You are on page 1of 34

Case Studies in Thermal Engineering 49 (2023) 103359

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering


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

Palladium-alloy membrane reactors for fuel reforming and


hydrogen production: Hydrogen Production Modeling
Mohamed A. Habib a, b, *, Md Azazul Haque a, b, Aadesh Harale c, Stephen Paglieri c,
Firas S. Alrashed c, Abduljabar Al-Sayoud a, Medhat A. Nemitallah a, b,
Shorab Hossain a, Ahmed Abuelyamen a, Esmail M.A. Mokheimer a, b,
Rached Ben-Mansour a
a
Mechanical Engineering Department and IRC- Hydrogen & Energy Storage, King Fahd University of Petroleum and Minerals, Dhahran, 31261,
Saudi Arabia
b
Researcher at K.A. CARE Energy Research & Innovation Center at Dhahran, Saudi Arabia
c
Carbon Management Division, R&D Center, Saudi Arabian Oil Company, Dhahran, 31311, Saudi Arabia

A R T I C L E I N F O A B S T R A C T

Handling Editor: Huihe Qiu Endeavors have recently been concentrated on minimizing the dependency on fossil fuels in order
Keywords:
to mitigate the ever-increasing problem of greenhouse gas (GHG) emissions. Hydrogen energy is
Palladium membranes regarded as an alternative to fossil fuels due to its cleaner emission attributes. Reforming of
Hydrogen separation modeling hydrocarbon fuels is amongst the most popular and widely used methods for hydrogen produc­
Membrane reactor tion. Hydrogen produced from reforming processes requires additional processes to separate from
Hydrogen purification the reformed gases. In some cases, further purification of hydrogen has to be carried out to use the
Fuel reforming modeling hydrogen in power generation applications. Metallic membranes, especially palladium (Pd)-based
Chemical kinetics of reforming reactions ones, have demonstrated sustainable hydrogen separation potential with around 99.99%
hydrogen purity. Comprehensive and critical research investigations must be performed to
optimize membrane-assisted reforming as well as to maximize the production of hydrogen. The
computational fluid dynamic (CFD) can be an excellent tool to analyze and visualize the flow/
reaction/permeation mechanisms at a lower cost in contrast with the experiments. In order to
provide the necessary background knowledge on membrane reactor modeling, this study reviews,
summarizes and analyses the kinetics of different fuel reforming processes, equations to deter­
mine hydrogen permeation, and lastly, various geometry and operating condition adopted in the
literature associated with membrane-reactor modeling works. It is indicated that hydrogen
permeation through Pd-membranes depends highly on the difference in hydrogen pressure. It is
found that hydrogen permeation can be improved by employing different pressure configuration,
introducing sweep flow on the permeate side of the membrane, reducing retentate side flow rate,
and increasing the temperature.

* Corresponding author. Mechanical Engineering Department and IRC- Hydrogen & Energy Storage, King Fahd University of Petroleum and Minerals, Dhahran,
31261, Saudi Arabia.
E-mail addresses: mahabib@kfupm.edu.sa (M.A. Habib), g201806100@kfupm.edu.sa (M.A. Haque), aadesh.harale@aramco.com (A. Harale), stephen.paglieri@
aramco.com (S. Paglieri), feras.alrashed@aramco.com (F.S. Alrashed), sayoudaq@kfupm.edu.sa (A. Al-Sayoud), medhatahmed@kfupm.edu.sa (M.A. Nemitallah),
g201805780@kfupm.edu.sa (S. Hossain), g201402540@kfupm.edu.sa (A. Abuelyamen), esmailm@kfupm.edu.sa (E.M.A. Mokheimer), rmansour@kfupm.edu.sa
(R. Ben-Mansour).

https://doi.org/10.1016/j.csite.2023.103359
Received 16 April 2023; Received in revised form 16 July 2023; Accepted 30 July 2023
Available online 2 August 2023
2214-157X/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

1. Introduction
Hydrogen is a perfect alternative source of clean energy in replacement of conventional fossil fuels, which result in massive
emissions of greenhouse gases causing global warming. Hydrogen has much better combustion characteristics in conventional com­
bustion systems and higher conversion efficiency in fuel cells [1]. The most widely used process to produce hydrogen is natural gas
(NG) steam reforming. This technology generates about 50% of the world ’s annual hydrogen supply. The key reaction taking place in
NG steam reforming is endothermic (CH4 + H2O ⇌ CO + 3H2) and is constrained by chemical equilibrium, thus achieving substantial
hydrogen production only with high temperature (850–900 ◦ C) activity [2,3]. As a result, a portion of feedstock of methane must be
burned in furnaces to sustain the heat duty for the reaction. This could be liable for lowering the overall efficiency of process, growing
emissions of greenhouse gas (GHG), and basing the cost of producing hydrogen on the cost of natural gas [4]. Besides methane, the
reforming or cracking several other hydrocarbon fuels such as methanol, ethanol, propane, toluene, glycerol-diesel, and naphtha, are
also employed to produce hydrogen [5–9].
Hydrogen separation from these processes can allow the enhancement in hydrogen production at lesser temperatures, because the
continuous selective elimination of hydrogen from the reaction system permits the composition of gas mixture to be held away from
equilibrium so that conditions of equilibrium are not reached and endothermic reactions can be conducted at lesser temperatures.
Furthermore, the lower operating temperature allows for the manufacture of the reforming tube with cheaper steel alloys in place of
the costly materials usually utilized to withstand the high operating temperatures of modern steam reforming plants [10–12].
The separation of hydrogen produced from reforming operation is carried out using several techniques such as cryogenic isolation,
pressure swing adsorption (PSA), and membranes [13]. Among these methods, the application of hydrogen-selective membranes is
gaining more attention due to its in-situ separation process, compact size and ultra-pure hydrogen extraction capability. Membrane
reactors are increasingly recognized as an efficient way for a wide variety of applications to replace traditional separation, processing,
and conversion technologies. By taking advantage of the advanced production of membrane materials, they can offer greater per­
formance, are highly adaptable and can have great economic potential. Depending upon the constituent materials, the
hydrogen-selective membranes are classified into four types: (a) polymer, (b) carbon, (c) metallic, and (d) ceramic membrane [14].
Among these, the metallic membranes, especially the palladium (pd)-based ones, demonstrated excellent diffusion properties for
mono-atomic hydrogen and the permeated hydrogen purity can reach up to 99.9999% [15]. Palladium membranes can also operate in
higher temperature, elevated pressure, can provide higher permeation fluxes, and have longer operational lifespan [16]. Several re­
view works have been performed in recent years outlining the different aspects of Pd-membrane hydrogen separation, including
membrane composition, type, performance of hydrogen permeation etc. Table 1 summarizes these surveys and outlines their key
findings.
Optimization of membrane integrated hydrocarbon reformer required significant research efforts. Experimental approach to this
optimization process will considerably raise the membrane development cost. Due to the limitations in experimental investigation, the
utilization of numerical modeling for maximizing hydrogen production and purification in membrane reactors is gaining much
attraction nowadays. The computational fluid dynamic (CFD) provides good visualization for the results and it allows to perform many
investigations at low cost compared to the experiments. The CFD modeling simulates the features of the gas flow as well as the
chemical processes by which the hydrogen is produced. There are various CFD codes (commercial and open-source) that can be utilized
to simulate the hydrogen production and purification, for instance, OpenFOAM [19], COMSOL Multi-physics [20,21], ANSYS Fluent
[22,23], and MATLAB [24]. To simulate the full process of hydrogen production and purification, the transport equations of conti­
nuity, momentum, energy, and species are required to be solved along with the chemical kinetics models for hydrogen generation and
hydrogen separation. In order to aid the R&D community with various aspects of membrane integrated hydrocarbon reformer, a
literature survey has been presented in this work on the kinetics of hydrocarbon reforming and the expressions for hydrogen
permeation in different membrane reactor configurations. The work presents critical review and analysis of modeling of chemical

Table 1
Overview of the key attributes of recent review works on hydrogen production using Pd-based membrane.

Researcher(s)/ Focused area Key findings


year

Habib et al. ⁃ Properties, production of Pd-based membranes and ⁃ Pd-membranes suffer from embrittlement, poisoning and
[15]/2021 reactor design phase transition which can be minimized by alloying Pd with
⁃ Density functional theory (DFT) computational other metals.
method to examine the permeation mechanism of Pd- ⁃ DFT can be useful tool in reducing the R&D cost of Pd-based
membranes membranes.
Sharma and ⁃ Ternary Pd-based membranes in gas-sensing ⁃ Ternary Pd-based membrane can be suitable for cases
Myung applications requiring high hydrogen selectivity with high tolerance to H2S
[17]/2020 ⁃ Aspects influencing permeability and solubility of poisoning.
hydrogen in ternary systems
Amiri et al. ⁃ Methane and liquid hydrocarbon use in hydrogen ⁃ Membrane integrated reforming reactors are relatively
[18]/2020 production inexpensive, energy efficient and compact compared to
⁃ Limitations of conventional reactor conventional ones.
⁃ Hydrogen production in different operating ⁃ Membrane reactors are limited to small/laboratory scale
parameters applications only and are often un-optimized.

2
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

kinetic reactions for hydrocarbons reforming as well as hydrogen purification/separation CFD modeling. The reaction mechanisms
with associated rates constants for reforming process of fuels in different catalysts are presented as well. Recommendations for
improvement of hydrogen permeation through Pd-membranes are provided.

2. Modeling of chemical kinetic reactions for hydrocarbons reforming


Hydrogen can be produced from various processes including hydrocarbon reforming hydrocarbon reforming, coal/biomass gasi­
fication, and electrolysis of water. Among these, hydrocarbon reforming contributes to almost 78% of the total hydrogen production
[11,25,26]. Depending upon the reactant type with hydrocarbon fuel, the reforming can be divided into three categories: (a) wet/­
steam reforming, (b) dry/CO2 reforming, and (b) auto-thermal/oxidation/partial oxidation reforming. Kinetics modeling of hydrogen
production primarily depends upon the type of reforming process considered [27]. The catalysts in these methods along with the fuel
type also determines the efficiency of the chemical process for hydrogen generation. Thus, in order to achieve and in-depth under­
standing of the reforming operation, this section presents the reactions and corresponding kinetic reaction rates for various fuels
including methane (CH4), propane (C3H8), toluene (C7H8), methanol (CH3OH), ethanol (C2H5OH), glycerol & diesel and lastly,
naphtha. Surface reactions associated with the catalyst in the reforming process is not included in this study.

2.1. Methane
Methane reforming is one of the most used methods to produce hydrogen from natural gas owing to its highest thermal efficiency.
Due to this reason along with the availability and handling of natural gas, this area of hydrogen generation is also highly researched.
There are three main ways in which methane-reforming operation can be conducted [28]: (a) steam/wet methane-reforming [29], (b)
CO2/dry methane-reforming [30] and (c) partial-oxidation of methane [31]. Moreover, the reforming operation can be done through
the combination of the three main processes as well [32]. Fig. 1 illustrates the techniques associated with methane reforming for
hydrogen generation. These processes greatly depend on the pressure, temperature and species distribution in the reformer, details of
which can be found in recent review works by Chen et al. [29] and Alli et al. [32]. Furthermore, auto-thermal reforming is another
reforming technique, which exothermic combustion of methane is used to supply the heat to endothermic reforming reactions [33].
Steam-methane reforming (SMR) or methane-steam reforming (MSR) is the widely used methane reforming technique and usually
involve three global chemical reactions, which are presented in Equations (1)–(3) [34], where the reaction 2 is commonly known as
water-shift gas reaction. The surface reactions along with the necessary values of pre-exponential factors (A), activation energies (Ea)
and temperature exponents (β) can be found in a detailed study by Maier et al. [35] for steam methane reforming over Ni-based
catalyst.
0
CH4 + H2 O → CO + 3H2 , ΔH298 = 206k Jmol− 1
(1)

0
CO + H2 O → CO2 + H2 , ΔH298 = − 41.1 kJmol− 1
(2)

0
CH4 + 2H2 O → CO2 + 4H2 , ΔH298 = 164.9 kJmol− 1
(3)

Xu and Forment [36] developed the reaction kinetics in Equations (4)–(6) for steam methane reforming over Ni-based catalyst,
which have been adopted in various numerical investigation [37–39]. Here r, k, P, K, and Keq correspond to the reaction rates, reaction
rate constants, partial pressure, rate constant for species adsorption and rate constants for equilibrium reactions, respectively. Most of
the CFD works of steam methane reforming [40–46] utilize the model by Xu and Forment [36].

Fig. 1. Methane-reforming processes for hydrogen generation.

3
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

( )
k1 PCO P3H
P2.5
PCH4 PH2 O − Keq,1
2

(4)
H2
r1 = ( )2
K P
1 + KCH4 PCH4 + KCO PCO + KH2 PH2 + HP2 OCH H2 O
4

( )
k2 PCO2 P4H
P 3.5 PCH4 P2H2 O − Keq,2
2

(5)
H2
r2 = ( )2
K P
1 + KCH4 PCH4 + KCO PCO + KH2 PH2 + HP2 OCH H2 O
4

( )
k3 PCO2 PH2
PCH4 PH2 O −
(6)
2
PH Keq,3
r3 = ( )2
K P
1 + KCH4 PCH4 + KCO PCO + KH2 PH2 + HP2 OCH H2 O
4

The expressions adopted for the reaction rate constant (k), equilibrium reaction rate constant (Keq), and species adsorption constant
rate (K) are given in Equations 8–10, while the values of the parameters used in these equations have been listed in Table 2.
( )
Ei
ki = Ai exp − (8)
R.T
( )
Ci
Keq,i = Bi exp − (9)
T
( )
Fj
Kj = Gj exp − (10)
R.T

In addition to the Xu and Forment [36] model, the model proposed by Numaguchi and Kikuchi [48] was also used in some numerical
simulation of steam methane reforming. Shigarov et al. [49] applied the model of Numaguchi and Kikuchi [48] on steam methane
reforming through 8%Ni/g-Al2O3 and 1%Rh/g-Al2O3 catalysts. Further details about the model and its kinetic constants can be found
on [48–50].
The values of pre-exponential factors and activation energies can be different depending on the nickel content in the catalyst. These
values along with the enthalpy change are summarized in Table 3 for different catalysts. The constants stated in Table 3 are based on
the catalyst reported on Xu and Forment [36] which consists of 15.8% nickel supported on a magnesium spinel (Ni-MgAl2O4) studied
by Jiwanuruk et al. [43], catalyst with 30% nickel content by Rodriguez et al. [45], Ni/MgAl2O4 and CaO by Lee et al. [51] and
commercial sulfide nickel catalyst (Ni-0309S) by Hoang et al. [52].
Dry reforming of methane (DRM) is another technique of hydrogen production from methane. In addition to the production of
hydrogen-rich fuel, DRM directly reacts with CO2 that aides in lowering CO2 emission [30]. On the other hand, dry reforming of
methane produces carbon during the reforming operation (Equations (14) and (15)), which deactivates the catalyst and lowers the
performance of catalytic conversion [53]. The conversion of methane and steam into a mixture of hydrogen, carbon monoxide, and
carbon dioxide is a combination of the above net reactions (Equation (1) -3) as well as the following reactions (Equations (11)–(17))
[35]. Although the chemical reaction mechanisms for methane reforming involves several reactions, most of these reactions can be
ignored without sacrificing the accuracy of the CFD model, except for the primary set of equations (Equations (1)–(3)).

Endothermic CO2 − reforming : CH4 + CO2 → 2CO + 2H2 , ΔHr0 = 247kJmol− 1


(11)

Table 2
Kinetic parameters for steam methane reforming reactions.

Rate constants for SMR reactions (ki ) [36]

i Reactions Ai (kmol bar5 kg−cat1 h− 1 ) Ei (Kj /mol)


15
1 CH4 + H2 O = CO + 3H2 14.225 × 10 240.1
2 CO + H2 O = CO2 + H2 1.02 × 1015 243.9
3 CH4 + 2H2 O = CO2 + 4H2 1.955 × 106 67.13
Rate constants for equilibrium reactions (Keq,i ) [47]
i Reactions Bi (dimension of Bi) Ci (Kj /mol)
1 CH4 + H2 O = CO + 3H2 1.198 × 1017 (kPa2) 26 830
2 CO + H2 O = CO2 + H2 2.117 × 1015 (kPa2) 22 430
3 CH4 + 2H2 O = CO2 + 4H2 1.767 × 10− 2 (− ) − 4400
Rate constants for species absorption (Kj ) [36]
j Species Gj (dimension of Gj) Fj (Kj /mol)
1 CH4 6.65 × 10− 4 (bar− 1) − 38.28
2 H2O 1.77 × 105 (− ) 88.68
3 CO 8.23 × 10− 4 (bar− 1) − 70.61
4 H2 6.12 × 10− 9 (bar− 1) − 82.90

4
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Table 3
Kinetic reaction constants for different catalysts in steam methane reforming.

Parameters 15.8% Nickel, Ni-MgAl2O4 [43] 30% Nickel [45] Ni/MgAl2O4 and CaO [51] Commercial sulfide nickel catalyst Ni-0309S [52]
15 14 12
k1 4.225 × 10 3.66 × 10 1.17 × 10 9.048 × 1011
k2 1.955 × 106 5.41 × 10− 3 5.42 × 102 5.43 × 105
k3 1.020 × 1015 8.83 × 1013 2.83 × 1011 2.14 × 109
Ea1 − 240.10 − 240.1 − 240.1 − 209.5
Ea2 67.13 67.1 67.13 70.2
Ea3 − 243.90 − 243.9 − 243.9 − 211.5
KCO 8.23 × 10− 5 8.23 × 10− 10 8.25 × 10− 5 8.11 × 10− 5
KH2 6.12 × 10− 9 6.12 × 10− 14 6.15 × 10− 9 7.05 × 10− 9
KCH4 6.65 × 10− 4 1.26 × 10− 6 6.65 × 10− 4 1.995 × 10− 3
KH2O 1.77 × 105 1.77 × 105 1.77 × 105 1.68 × 104
ΔH CO − 70.65 − 70.7 − 70.65 − 70.23
ΔH H2 − 82.90 − 82.9 − 82.9 − 82.55
ΔH CH4 − 38.28 − 27.3 − 38.28 − 36.65
ΔH H2O 88.68 88.7 88.68 85.77

Exothermic water − gas shift : CO + H2 O → CO2 + H2 , ΔHr0 = − 41.2kJmol− 1


(12)

Methanation (1) : CO + 3H2 → CH4 + H2 O, ΔHr0 = − 206kJmol− 1


(13a)

Methanation (2) : 2CO + H2 → CH4 + CO2 , ΔHr0 = − 247kJmol− 1


(13b)

Carbon formation by boundary reaction : 2CO → C + CO2 , ΔHr0 = − 172.4kJmol− 1


(14)

Carbon formation by methane cracking : CH4 → C + 2H2 , ΔHr0 = − 74.9kJmol− 1


(15)

Gasification of carbon by steam : C + H2 O → CO + H2 , ΔHr0 = 131.3kJmol− 1


(16)

Gasification of carbon by oxygen : C + O2 → CO2 , ΔHr0 = − 393.5kJmol− 1


(17)

Lee et al. [21,54] conducted a CFD simulation for dry reforming of methane over Rh/Al2O3 catalyst, where the reactions used were
taken from the investigation by Richardson and Paripatyadar [55]. The reforming and shift kinetics of dry reforming can be expressed
by the following reactions: (Equations (18) and (19)), where Reaction 19 is the opposite reaction of water-gas shift reaction.
CH4 + CO2 → 2CO + 2H2 (18)

H2 + CO2 → CO + H2 O (19)
The reaction rates for these dry reforming reactions are presented in Equations (20) and (21). Here the terms r, k, K and P refer to
reaction rate, rate constant, equilibrium constant and partial pressure, respectively. The expression to evaluate the rate constants are
provided in Equations (22) and (23), while Equations (24)–(27) contain the equilibrium constant expression.
[ ][ ]
KCO2 KCH4 PCO2 PCH4 (PCO PH2 )2
r1 = k1 1− (20)
1 + KCO2 PCO2 + KCH4 PCH4 K1 PCH4 PCO2

[ ]
PCO PH2 O
r2 = k2 PCO2 1 − (21)
K2 PCO2 PH2
( )
[ ] 1 [ ]
k1 = 1290 × 5000 × exp − 102065 Jmol− 1 mol s− 1 m− 3 (22)
RT
( )
[ ] 1 [ ]
k2 = 1.856 × 5000 × exp − 73105 Jmol− 1 mol s− 1 atm− 1 m− 3 (23)
RT
( )
[ ] 1 [ ]
K1 = e34.011 × exp − 258598 Jmol− 1 atm2 (24)
RT
( )
[ ] 1
K2 = 68.78 × exp − 37500.3 Jmol− 1 (25)
RT

5
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

( )
[ ] 1 [ ]
KCH4 = 2.63 × 10− 2 × exp 37641 Jmol− 1 atm− 1 (26)
RT
( )
[ ] 1 [ ]
KCO2 = 2.64 × 10− 2 × exp 40684 Jmol− 1 atm− 1 (27)
RT
Catalytic partial oxidation of methane usually contains additional oxidation reaction along with three-steam reforming reactions
[56]. The reaction pathways can be varied depending on the type of catalyst used and on the investigation approach. Typically,
majority of the research work include one additional reaction (CH4 + 2O2 →CO2 + 2H2 O) along with the commonly used steam
reforming reactions (Equations (1)–(3)) to investigate the partial oxidation. However, there are other configuration of equations as
well. A summary of the research works done on the kinetics of partial oxidation of methane is provided in Table 4. In addition to these,
a detailed kinetic mechanism for the partial oxidation of methane has been developed by Noagre [57] in 2008.

2.2. Methanol
The steam reforming of methanol also encompasses three primary reactions. These reactions are the methanol steam reforming,
methanol decomposition, and water-gas shift in Equations (28)–(30), respectively [63]. The reactions rates for these reactions are over
Cu/ZnO/Al2O3 catalyst are provided in Equations (31)–(33). The elementary and surface reaction of methanol steam reforming over
Cu/ZnO/Al2O3 catalyst is discussed in details in an earlier study by Pepply et al. [64,65].
CH3 OH + H2 O ↔ CO2 + 3H2 (28)

CH3 OH ↔ CO + 2H2 (29)

CO + H2 O ↔ CO2 + H2 (30)
( )[ ]
PCH3 OH P2 PCO
k1 KCH ∗
3O
(1) P0.5
1 − Keq,1 PHCH2 OH2PH O CS1 T T
CS1a
(31)
H2 3 2
r1 = [ ( ) ( )]( )
∗ PCH3 OH ∗ 0.5 ∗ PH2 O 0.5 0.5
1 + KCH 3 O (1) P 0.5 + K HCOO (1) PCO2 PH2 + KOH (1) P 0.5 1 + K H (1a) PH2
H2 H2

( )[ ]
PCH3 OH P2 PCO
k12 KCH∗
3O
(2) P0.5
1 − Keq,2HP2 CH OH CS2
T T
CS2a
(32)
H 3
r2 = [ ( )2 ( )]( )
∗ PCH3 OH ∗ PH 2 O 0.5 0.5
1 + KCH 3 O (2) P 0.5 + K OH (2) P 0.5 1 + KH (2a) PH2
H2 H2

( )[ ]
PCO PH2 O P P T2
k3∗ KOH

(1) P0.5
1 − Keq,3HP2 COCOPH2 O CS1
(33)
H2 2
r3 = [ ( ) ( )]2
∗ PCH3 OH ∗ 0.5 ∗ PH2 O
1 + KCH 3O
(1) P 0.5 + K HCOO (1) PCO 2
P H 2
+ K OH (1) P0.5
H2 H2

In the auto-thermal reforming of methanol, Ghasemzadeh et al. [66] reported that the methanol oxidization reaction also takes place in
the reactor as shown in Equation (34). The reaction rate of this reaction can be evaluated using Equation (35). The kinetic rate re­
actions for methanol steam reforming can be found in a study by Uriz et al. [67] for Pd/ZnO, Heidarzadeh and Taghizadeh [68] for
Cu/ZnO/Al2O3 catalyst. Additionally, the methanol steam reforming over Ni-based catalyst is overviewed in a review work by Bepari
and Khila [69].
Methanol oxidation reaction : CH3 OH + 1.5O2 ↔ CO2 + 2H2 O (34)

Table 4
Summary of reaction kinetics for partial-oxidation of methane.

Researcher(s)/year Catalyst Number of Key reaction types


reactions

Hoang et al. [58]/2005 Ni catalyst 4 3 SMR reactions, 1 POM reaction


Barrio et al. [56]/2006 Ru/γ-Al2O3 4 3 SMR reactions, 1 POM reaction
Fernandes et al. Zirconium and platinum supported in alumina or nickel supported in 3 1 SRM, 1 DRM and 1 POM
[59]/2006 alumina reaction
Dobrego et al. [60]/2008 Inert porous medium 3 1 SMR, 2 POM reactions
Chen et al. [61]/2014 Ru catalyst 3 1 SRM, 1 DRM and 1 POM
reaction
Prukswan et al. Ni catalyst 18 Reduced POM reaction
[62]/2014 mechanism

6
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

P0.18
CH3 OH PO2
0.18
r4 = k4 (35)
P0.14
H2 O

2.3. Ethanol
Similar to the methane and methanol-reforming, the kinetic properties of ethanol reforming for different catalysts have been also
investigated in the literature, and as a result, different mechanisms have been proposed. In general, four reactions are considered to be
associated with ethanol steam reforming as showed in Equations (36)–(39) are the most representative reactions for the ethanol
reforming process [70,71].
C2 H5 OH ↔ CH4 + CO + H2 (36)

CO + H2 O ↔ CO2 + H2 (37)

CH4 + H2 O ↔ CO + 3H2 (38)

CH4 + 2H2 O ↔ CO2 + 4H2 (39)

Like other reforming, Ni-based catalysts are the most utilized ones in ethanol reforming operation [70–75]. Ni-catalysts with lower
loadings (eg. 6 wt% Ni in Ni/Al2O3 catalyst) are found more efficient in ethanol steam reforming for hydrogen production [76]. The
kinetic rates for the ethanol decomposition reaction (Equation (36)) are presented in Equation (40), where X and k denote the molar
fraction of gas component and reactions rate constant, respectively. The reaction rate constant (k1 ) can be determined using Equation
(41) [77–79], the value of which is 4.26 × 10− 9 mol.Pa− 1g−cat1s− 1 at 773 K reforming temperature. The remaining reactions (Equations
(37)–(39)) are identical to the methane steam reforming. Thus, the kinetic parameters proposed by Xu and Forment [36] can also be
applied for these reactions over Ni-based catalyst. Kinetic reaction rates of ethanol reforming reactions for different other catalyst can
be found on [80–85].
r1 = k1 PXC2 H5 OH (40)
( 5
) ( )
4.55 × 10− 2030
k1 = exp − (41)
T T

2.4. Propane
Rakib et al. [86,87] carried out Propane steam reforming in membrane reactor of fluidized bed for H2 separation. There are
different methods of producing hydrogen from propane, among which the steam reforming of propane is considered as more
economical one. In steam propane reforming process, at first the propane reacts with steam in an irreversible reaction in Equation (42).
Some portion of propane decomposes producing carbon and hydrogen (Equation (43)), which is often omitted from numerical
simulation [86]. The produced CO and H2 from the first reaction again react in a reversible form of methane and steam reactions. This
reaction is known as reverse steam methane reforming and followed by water-gas shift reaction and methane reforming reaction as
shown in Equations (44) and (45), respectively. Apart from these reactions, there are some unwanted side reactions that take place
inside the reactor, which are presented in Equations (46)–(49) [87].
0
C3 H8 + 3H2 O → 3CO + 7H2 , ΔH298 = 499kJmol− 1
(42)

C3 H8 → 3C + 4H2 (43)

0
CO + 3H2 ⇄CH4 + H2 O, ΔH298 = − 206kJmol− 1
(44)

0
CH4 + 2H2 O⇄CO2 + 4H2 , ΔH298 = 165kJmol− 1
(45)

CH4 ⇄C + 2H2 (46)

2CO⇄C + 2CO2 (47)

CO + H2 ⇄H2 O + C (48)

CO2 + 2H2 ⇄2H2 O + C (49)


Since reactions 42 and 44–46 are typically incorporated in propane steam reforming modeling, the kinetics rates of these reactions
are presented here. Rate expression for first reaction (Equation (42)) in industrial Ni-based catalyst is obtained from the investigation
by Ma [88], as shown in Equation (50). Here, the value of θ is 1 kPa-0.86 and k1 can be calculated using Equation (51) [88].
′ 0.93 ′ − 0.53
k1 PC3 H8 P H2 O
r1 = ′ 0.86
(50)
1 + θPH2

7
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

189.63×103
k1 = 2.1428 × 1014 × e− RT (51)
The remaining kinetic reaction rates for reactions (Equations (44)–(43)) can be determined by the Xu and Forment [36] model. The
rate expressions for these reactions are provided in Equations (52) and (53). The rate constant in Ni-based catalyst considered here
differs slightly from that in section 2.1, while the rate constants for species adsorption are same as values in Table 2. The reaction rate
constants and equilibrium constant are presented in Equations (5) and (55)–(57)8-59, respectively.
( )
PCO P0.5
H2 PCH4 PH2O
k2
− PH2.5

(52)
2
r2 = k2 ( )2
1 + KCO PCO + KH2 PH2 + KCH4 PCH4 + KH2O PH2O
PH2

( )
PCO PH2O PCO2

(53)
PH2 k3
r3 = k3 ( )2
1 + KCO PCO + KH2 PH2 + KCH4 PCH4 + KH2O
PH2
PH2O

( )
PCH4 P2H PCO2 P0.5
2O H
P3.5
− k2 k3
2

(54)
H 2
r4 = k4 ( )2
1 + KCO PCO + KH2 PH2 + KCH4 PCH4 + KH2O
PH2
PH2O

[ ]
240.1X103 kmol.bar0.5
k2 = 9.49 × 1015 × e− RT (55)
kgcat .h
[ ]
67.13X103 kmol
k3 = 4.39 × 106 × e− RT (56)
kgcat .h.bar
[ ]
243.9X103 kmol bar0.5
k4 = 2.29 × 1015 × e− RT (57)
kgcat .h
( )
− 26.83X103 +30.114
( )
(58)
T
K2 = e bar2
( )
4400− 4.036
(59)
T
K3 = e
Several numerical investigations have been performed on propane reforming process. Odiba et al. [89] studied catalytic oxidation
of propane in Au/Cr/γ-Al2O3 reformer bed in temperatures and velocities ranging from 563 to 663 K and 0.05–1 m/s, respectively. The
report said that the conversion of propane increases with temperature and decreases with flow inlet rise. About 99.99% conversion was
observed in-between 0.05-1 m/s. Zang et al. [90] investigated combined effect of methane partial oxidation and propane steam
reforming in Rh-based catalytic reactor. Studying the results within 373–1173 K operating temperature, they observed the combined
process to be more effective than propane steam reforming counterpart. Ghasemzadeh et al. [91] numerically examined the dehy­
drogenation process of propane in 673–873 K temperatures and 0.1–1 L/min flow rate. They studied the operation in both

Table 5
Summary of propane reforming numerical works.

Researcher(s)/year Reforming type Operating condition Catalyst/ Key findings


reformer

Odiba et al. Catalytic oxidation Temperature: 563–663 Au/Cr/ Propane conversion increased with temperature and
[89]/2013 K, inlet velocity: 0.05–1 γ-Al2O3 decreased with flow velocity. About 99.99% conversion
m/s was recorded between 0.05 and 1 m/s
Zhang et al. Partial oxidation of methane Temperature: 373–1173 Rh based Partial oxidation of methane and steam reforming of
[90]/2016 and steam reforming of K catalyst propane yielded higher hydrogen compared to partial
propane oxidation of methane only.
Ghasemzadeh et al. Dehydrogenation Temperature: 673–873 Pd-based Hydrogen yield increased by 41–49% in membrane
[91]/2017 K, flow rate: 0.1–1 L/min membrane reactor compared to conventional counterpart
Karadeniz et al. partial oxidation and steam Temperature: 313 and Rh/Al2O3 A thin oxidation zone emerges near gas-washcoat
[92]/2020 reforming of methane and 423 K Catalyst interface in partial oxidation reforming, followed by two
propane zones for steam and dry reforming respectively.
Barnoon et al. Steam reforming Temperature: 750–900 Packed bed Increase in temperature and velocity improved the
[93]/2021 K, inlet velocity: 0.5–1.5 hydrogen production to 92.2% from 77.5% at base
m/s configuration. Reaction rates also improved with the
increase in both these parameters.

8
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

conventional and Pd-membrane reactors. The study reported that hydrogen yield to be increased by 41–49% in membrane reactor
compared to conventional counterpart. Karadeniz et al. [92] investigated partial oxidation and steam reforming of methane and
propane in Rh/Al2O3 Catalyst at 313 and 423 K temperatures. They reported the emergence of a thin oxidation zone near gas-wash coat
interface in partial oxidation reforming, followed by two zones for steam and dry reforming respectively. Barnoon et al. [93] per­
formed a numerical analysis on propane steam reforming in 750–900K temperatures and 0.5–1.5 m/s inlet velocities in a packed bed
reformer. They reported an increase in temperature and velocity improved the hydrogen production to 92.2% from 77.5% at base
configuration. They also stated that the reaction rates improved with the increase in both these parameters. A summary of these works
is provided in Table 5.

2.5. Toluene
Steam-reforming of toluene can also be a potential source for hydrogen production. Numerous investigations have been carried out
to produce hydrogen from toluene [94–96]. At first, toluene reacts with steam in a forward-only reaction producing CO and H2. The
produced CO then reacts with the steam and turns into CO2 and H2. A portion of methane can be mixed with the toluene stream, in
which case the reforming process of methane also takes place. Equations (60)–(62) provide the reactions associated with the
steam-toluene reforming process [95].
0
C7 H8 + 7H2 O → 7CO + 11H2 , ΔH298K = 869.17 kJmol− 1
(60)

0
CH4 + H2 O⇄CO + 3H2 , ΔH298K = 206.12 kJmol− 1
(61)

0
CO + H2 O⇄CO2 + H2 , ΔH298K = − 41.17kJmol− 1
(62)

The kinetic reactions rates of the above reactions are provided in Equation (63)–(65), respectively. The reaction rate of endo­
thermic forward reaction of toluene-steam depends only on the concentration of toluene in the feed stream. While the other remaining
reactions, being reversible, can shift to backward directions, the kinetic rate expressions for which are presented in Equations (64) and
(65), respectively. Further details on the reaction rates can be found in the study by Oliveira and Da Silva [95].

R1 = 8.723 × 109 CC7 H8 ; (kgcat s)− 1


(63)
⎡ ⎤
( )
15 ⎢ CCO2 CH2 2 ⎥
R2 = 3.101 exp − ⎢CCH4 CH2 O −
⎣ ( )⎥ ; (kgcat s)− 1
(64)
Tg 32.9 ⎦
0.0265 exp Tg

⎡ ⎤
( )
138 ⎢ CCO2 CH2 ⎥
R2 = 250 exp − ⎢CCO CH2 O − ( )⎥ ⎦; (kgcat s)
− 1
(65)
Tg ⎣ 0.0265 exp 3966
Tg

Even though a number of studies have been performed on the toluene reforming, numerical/CFD works among these are fairly
limited. Zhao et al. [97] numerically investigated toluene-steam reforming packed bed reactor filled with Ni/cordierite catalyst. Steam
to carbon ratio was fixed at 2 while the temperatures were varied from 1023-1173 K. The reported that H2 content varies marginally
with temperature and observed 66 mol.% hydrogen remains after reforming. Moreover, approximately 94.1% conversion efficiency
was recorded at 1173 K. Oliveira and Silva [98] developed a numerical model to study the steam reforming of toluene in a fixed bet
reactor. The developed model was tested at 400 K temperature and 2.367 × 10− 2 m3/s gas flow rate. Under the testing conditions the
model provided reasonably accurate estimation on C7H8, CO, H2, CH4 and CO2 species. Oemar et al. [98] conducted a computational

Table 6
Summary of numerical investigations on toluene steam reforming.

Researcher(s)/ Operating condition Catalyst/reformer type Key findings


year

Zhao et al. Temperature: 1023–1173 K, S/ Ni/cordierite catalyst, packed H2 content varies marginally with temperature and 66 mol%
[97]/2010 C molar ratio: 2 bed hydrogen remains after reforming, 94.1% conversion efficiency
observed at 1173 K.
Oliveira and Temperature: 400 K, gas flow Fixed bed Developed model can provide a good estimation on C7H8, CO, H2,
Silva rate: 2.367 × 10− 2 m3/s CH4 and CO2 species.
[98]/2013
Oemar et al. Temperature: 100–1100 ◦ C, La0.8Sr0.2Ni0.8Fe0.2O3 catalyst, Adsorbed oxygen species is produced from the oxygen from the
[98]/2014 total flow rate: 375–580 ml/ fixed bed redox property of some metals such as Fe, water activation and/or
min, lattice oxygen species, while adsorbed aldehyde species produces
through reaction between adsorbed CH2 or C2H2 and adsorbed
oxygen.
Anjos et al. Temperature: 400 K Ni2Mg4Al1.8Ce0.2 and Ni/Olivine Hydrogen yield is 1.8 times higher with toluene than methane under
[94]/2020 catalysts, fixed bed identical operating condition.

9
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

study in La0.8Sr0.2Ni0.8Fe0.2O3 catalyst-filled fixed-bed reactor in 100–1100 ◦ C temperatures and 375–580 ml/min total flow rates. The
results show that adsorbed oxygen species is produced from the oxygen from the redox property of some metals such as Fe, water
activation and/or lattice oxygen species, while adsorbed aldehyde species produces through reaction between adsorbed CH2 or C2H2
and adsorbed oxygen. Anjos et al. [94] studied toluene steam reforming in a fixed bed reactor considering Ni2Mg4Al1.8Ce0.2 and
Ni/Olivine catalysts at 400 K operating temperature. They reported approximately 1.8 times higher Hydrogen yield with toluene than
methane. These studies are also summarized and presented in Table 6.

2.6. Glycerol and diesel


Glycerol, a common by-product of biodiesel production, is also gaining tractions for its potential in hydrogen production through
reforming. Diesel itself can be used in hydrogen production using reforming techniques. Similar to the methane, propane and other HC
fuels, hydrogen can be produced from glycerol/diesel in several ways. The most common processes, are steam reforming, partial
oxidation-auto-thermal reforming, aqueous phase reforming, supercritical water reforming [99,100], as depicted in Fig. 2. Here, steam
reforming refers to the catalytic reaction of steam with diesel/glycerol, aqueous phase reforming corresponds to the conversion of
hydrocarbon in water to hydrogen rich product gases [101], auto-thermal reaction utilized the combustion of supply heat, lastly, the
supercritical water reforming stands for reforming process in high moisture content at high pressure [102]. Since steam reforming is
more widely used hence this study focuses one steam reforming of glycerol/diesel fuel.
An extensive literature on the kinetics and mechanism of fuel such as glycerol and diesel has been reported based on the power-law
[103–109]. Various types of catalysts including transition metals (Ni, Co and Mo) and noble metals (Ru, Rh, Pd and Pt) supported on
different oxides can be used, thus, there are some variances on the constant values of kinetic rate reactions. The reaction mechanisms
include many feasible reactions, where the most important reactions are glycerol steam reforming reaction (Equation (66)) and
glycerol decompositions (Equation (67)) in addition to water gas shift reaction (Equation (68)) [110]. The rate of glycerol steam
reforming along with the formation rates of H2, CO2 and CO can be obtained by the partial pressures of the glycerol and steam over
Co-Ni/Al2O3 catalyst (Eq. (69)) [110,111]. The values of the kinetic parameters are listed in Table 7. A comprehensive report on
glycerol surface mechanisms and side reactions for Co-Ni/Al2O3 catalyst can be found in the experimental work by Cheng et al. [105].
C3 H8 O3 + 3H2 O → 7H2 + 3CO2 (66)

C3 H8 O3 → 4H2 + 3CO (67)

CO + H2 O ↔ CO2 + H2 (68)
( )
− Ea
RT
ri = Ae pβglycerol pγglycerol (69)

Additional reaction mechanisms in the modeling of glycerol steam reforming were also considered by some researchers, including
Saidi et al. [111], Shahirah et al. [112] and Yang et al. [113]. For instance, Saidi et al. [111] considered the CO methanation
mechanism (Equation (70)) with a kinetic reaction model (Equation (71)) extracted from the work of Shahirah et al. [112]. Where, the
term φ can be calculated using Equation (72).
CO + 3H2 → CH4 + H2 O (70)

Fig. 2. Glycerol/diesel reforming processes.

10
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Table 7
Kinetic parameters for components consumption and production [105,110].

Components Ea (kJ.mol− 1) A (mol.m− 2s− 1


kPa− (β+γ)
) γ β

C3H8O3 63.3 0.036 0.358 0.253


H2 67.28 0.468 0.274 0.253
CO2 64.06 0.074 0.403 0.281
CO 61.73 0.062 − 0.065 0.308

( )( )
k p3H2 pCO 1
r= 2.5
− pCH4 pH2 O (71)
p H2 K φ2

p H2 O
φ = 1 + KCO pCO = KH2 pH2 + KCH4 pCH4 + KH2 O (72)
p H2
Yang et al. [113] reported other reaction mechanism and stated that the main chemical reactions can be described by Equations
(73)–(75) in a sorption-enhanced glycerol-steam reforming. The reaction rates for these reactions can be evaluated using Equations
(76)–(78), respectively. Additional kinetic reaction should be included for the application of CaO as CO2 sorbent, as shown in Equation
(79), where κc and CCO2 ,eq refer to kinetic coefficient and equilibrium concentration of CO2 [113].

C3 H8 O3 + H2 O → CH4 + 2CO2 + 3H2 (73)

CH4 + H2 O ↔ CO + 3H2 (74)

CO + H2 O ↔ CO2 + H2 (75)
( )
74210
R1 = 1.838 × 105 × exp − CC3 H8 O3 CH2 O ρcat (76)
T
( )
26830
R2 = 1.198 × 1017 × exp − CCH4 CH2 O ρcat (77)
T
( )
4400
R3 = 0.01767 × exp − CCO CH2 O ρcat (78)
T
( )2.61
( )0.37 X
RCO2 = CCaO κc CCO2 − CCO2 ,eq 1− (79)
Xu

2.7. Naphtha
Catalytic reforming of naphtha is also another vital processes for hydrogen generation. In addition, naphtha reforming is also
important in refinery applications for the production of high-octane components. Similar to other hydrocarbon fuel reforming, the
performance in naphtha reforming also depends on the catalytic medium. Majority of the catalyst used in naphtha reforming are based
on platinum (Pt) based catalyst, which form with Pt, bi- and tri-metallic formation along with an acid function (Al2O3-Cl/Al2O3) [114].
Rahimpour [115] presented naphtha reforming in a fluidized-bed membrane reactor (FBMR) with a perm-selective Pd-Ag mem­
brane for H2 separation. The catalytic reforming process can be represented with four dominant idealized reactions, (a) dehydroge­
nation of naphthenes to aromatics (Equation (80)), (b) dehydrocyclization of paraffins to naphthenes (Equation (81)), (c)
hydrocracking of naphthenes to lower hydrocarbons (Equation (82)) and (d) hydrocracking of paraffins to lower hydrocarbons
(Equation (83)) [115,116].
( )
kj
Naphthenes(Cn ) ↔ Aromatics(Cn H2n− 6 ) + 3H2 ; ΔH1 = 71038.06 (80)
mol
( )
kj
Paraffins(Cn H2n+2 ) ↔ Naphthenes(Cn H2n ) + H2 ; ΔH2 = − 36953.33 (81)
kmol
( )
n kj
Naphthenes(Cn H2n ) + H2 →Lighter ends(C1 − C5 ); ΔH3 = − 51939.316 (82)
3 kmol
( )
n− 3 kj
Paraffins(Cn H2n+2 ) + H2 →Lighter ends(C1 − C5 ); ΔH4 = − 56597.546 (83)
3 kmol
Reaction rates for these reactions are expressed in Equations (84)–(87), while the rate constants for forward and equilibrium re­
actions are provided in Equations (88)–(90) and Equation (91) and (92), respectively. Where, κei equilibrium constant with unit MPa3
for reaction 79 and MPa− 1 for Reaction Eq. (80)). The terms κfi forward rate constants have units kmol./h.kgcat.MPa for reaction (Eq.

11
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Table 8
Reactions, equations for reaction rates, rate constants, and equilibrium constants for naphtha reforming by Iranshahi et al. [117].

Reaction type Reaction General reaction Reaction rate Rate constant equation; unit Equilibrium rate
No. equation constant equation;
unit
( ) ( )
Hydrogenation 1 ACHn ↔ An + 3H2 ( E B
r1n = κ1n PACHn − κ1n = exp a − ; (kmol. K1n = exp A − ;
RT T
) kg−cat1 h− 1 kPa− 1 ) 3
(kPa )
PAn P3H2
K1n
( ) ( )
Paraffin 2 ACHn + H2 ↔ NPn r2n = E B
( κ2n = exp a − ; (kmol. K2n = exp A − ;
dehydrocyclization to RT T
κ2n PACHn PH2 −
naphthene kg−cat1 h− 1 kPa− 2 ) (kPa− 1 )
)
PNP
K2n
( ) ( )
3 ACHn + H2 ↔ IPn r3n = E B
( κ3n = exp a − ; (kmol. K3n = exp A − ;
RT T
κ3n PACHn PH2 −
) kg−cat1 h− 1 kPa− 2 ) (kPa− 1 )
PIPn
K3n
( ( ) ( )
4 NPn ↔ ACPn + H2 E B
r4n = κ4n PNPn − κ4n = exp a − ; (kmol. K4n = exp A − ;
RT T
)
PACPn PH2 kg−cat1 h− 1 kPa− 1 ) (kPa)
K4n
( ( ) ( )
5 IPn ↔ ACPn + H2 E B
r5n = κ5n PIPn − κ5n = exp a − ; (kmol. K5n = exp A − ;
RT T
)
PACPn PH2 kg−cat1 h− 1 kPa− 1 ) (kPa)
K5n
( ) ( )
Paraffin 6 NPn ↔ An + 4H2 ( E B
r6n = κ6n PNPn − κ6n = exp a − ; (kmol. K6n = exp A − ;
dehydrocyclization to RT T
aromatic ) kg−cat1 h− 1 kPa− 1 ) (kPa4 )
PAn P4H2
K6n
( ) ( )
7 IPn ↔ An + 4H2 ( E B
r7n = κ7n PIPn − κ7n = exp a − ; (kmol. K7n = exp A − ;
RT T
) kg−cat1 h− 1 kPa− 1 ) 4
(kPa )
PAn P4H2
K7n
( ( ) ( )
Isomerization of 8 ACPn ↔ ACHn E B
r8n = κ8n PACPn − κ8n = exp a − ; (kmol. K8n = exp A − ;
naphthenes and RT T
)
paraffins PACHn kg−cat1 h− 1 kPa− 1 ) (− )
K8n
( ( ) ( )
9 NPn ↔ IPn E B
r9n = κ9n PNPn − κ9n = exp a − ; (kmol. K9n = exp A − ;
RT T
)
PIPn kg−cat1 h− 1 kPa− 1 ) (− )
K9n
( ( ) ( )
Isomerization of aromatics 10 R ↔ P, for n = 1, …, 6 E B
r10n = κ10n PR − κ10n = exp a − ; (kmol. K10n = exp A − ;
RT T
)
PP kg−cat1 h− 1 kPa− 1 ) (− )
K10n
( ( ) ( )
Transalkylation. 11 2R ↔ P + C, for n = 7, 8 E B
r11n = κ11n P2R − κ11n = exp a − ; (kmol. K11n = exp A − ;
RT T
)
PP PC kg−cat1 h− 1 kPa− 2 ) (− )
K11n
( ) ( )
Cracking of paraffins 12 P2 + H2 →2P1 , for n = 2; PNPn E
r12n = κ12n κ12n = exp a − ;
P3 + H2 →P1 + P2 , for n = 3; Pt RT
2 (kmol.kg−cat1 h− 1 )
NP4 + H2 → (P3 + P2 + P1 ),
3
for n = 4;
1
NP5 + H2 → (NP4 + P3 + P2 +
2
P1 ), for n = 5;
n− 3 1 ∑5
NPn + H2 → P,
i=1 i
3 2
for n = 6 − 9
( ) ( )
13 3 1 PIPn E
IP4 + H2 → (P3 + P2 + 3P1 ), r13n = κ13n κ13n = exp a − ;
2 2 Pt RT
for n = 4;
(kmol.kg−cat1 h− 1 )
1
IP5 + H2 → (IP4 + P3 + P2 +
3
3P1 ), for n = 5;
(continued on next page)

12
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Table 8 (continued )

Reaction type Reaction General reaction Reaction rate Rate constant equation; unit Equilibrium rate
No. equation constant equation;
unit

n− 3 n ∑5
IPn + H2 → P,
i=1 i
3 15
for n = 6 − 9
Cracking of naphthenes 14 n n ∑5 (
PAHCn
) (
E
)
ACHn + H2 → P
i=1 i r14n = κ14n κ14n = exp a − ;
3 15 Pt RT
(kmol.kg−cat1 h− 1 )
( ) ( )
15 n 1 PAHPn E
ACP5 + H2 → (NP4 + P3 + r15n = κ15n κ15n = exp a − ;
3 2 Pt RT
P2 + P1 ), for n = 5;
(kmol.kg−cat1 h− 1 )
n n ∑5
ACPn + H2 → P,
i=1 i
3 15
for n = 6 − 9
( )
Hydrodealkylations 16 An+1 + H2 →Ak + Cm H2m+2 r16n = κ16n PAn+1 P0.5 E
H2 κ16n = exp a − ;
RT
(kmol.kg−cat1 h− 1 kPa− 1.5
)

(79)) and kmol./h.kgcat.MPa2 for reaction (Eq. (80)), and kh.kgcat for reactions 82 and 83 [115].
( )
κf1 ( )
r1 = Ke1 Pn − Pa P3h (84)
Ke 1
( )
κf2 ( )
r2 = K e 2 Pn Ph − Pp (85)
Ke 2
( )
κf3
r3 = Pn (86)
Pt
( )
κf4
r4 = Pp (87)
Pt
( )
E1
κf1 = 9.87 exp 23.21 − a (88)
1.8T
( )
E2
κf2 = 9.87 exp 35.98 − a (89)
1.8T
( )
E3
κf3 = κf4 = exp 49.97 − a (90)
1.8T
( )
50784
κe1 = 1.04 × 10− 3 exp 46.15 − (91)
1.8T
( )
8000
κe2 = 9.87 exp − 7.12 + (92)
1.8T
A detailed kinetic study on naphtha reforming is performed by Iranshahi et al. [117], where kinetic rate constants are divided into
several groups, including (a) dehydrageneration of naphthenes (alkyl-cyclohexanes), (b) paraffin dehydrocyclization to naphthene, (c)
paraffin dehydrocyclization to aromatics, (d) isomerization of naphthenes and paraffins, (e) isomerization of aromatics, (f) trans­
alkylation, (g) Cracking of Paraffins, (h) cracking of naphthenes and (j) hydrodealkylations. The reactions, equations for reaction rates,
rate constants, and equilibrium constants are present in Table 8, whereas the values of these rate constants are provided in Tables 9–13.
Over the years several kinetic reaction models have been developed to study the catalytic naphtha reforming. Since naphtha is a
complex mixture of hydrocarbons with hundreds of species/components, thus, these reaction mechanisms vary in their number of
reactions and lumped components. These lumped models are proposed to reduce and categorize similar reactions, where a large
number of components are characterized by a smaller kinetic lump set [114]. A summary of these kinetics models from open literature
are presented in Table 14.

3. H2 purification/separation CFD modeling


Hydrogen rich gaseous mixture produced from hydrocarbon fuel reforming processes necessitates additional steps to separate the
hydrogen from reformed gas stream or to improve the purity of produced hydrogen. As discussed before in section 1, Pd-based
membrane can be an excellent alternative for in-situ hydrogen separation/purification operation. Membrane reactors can also

13
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Table 9
Heat of reaction and rate constants for hydrogenation, paraffin dehydrocyclization to naphthene, Paraffin dehydrocyclization to aromatic, and isomerization of
naphthenes and paraffins reactions in naphtha reforming reactions [117].

C6 E/R a A B ΔH (kJ/mol H2)

Hydrogenation: Reaction-1
C7 19.50 18.75 59.90 24.80 × 103 68.73
C8 19.50 20.70 60.23 25.08 × 103 208.47
for An = MX 19.50 17.89 60.37 23.27 × 103 64.50
for An = OX 19.50 19.15 60.32 23.49 × 103 65.10
for An = PX 19.50 18.66 60.13 23.36 × 103 64.74
for An = EB 19.50 18.71 60.40 24.78 × 103 68.70
C9+
19.50 20.38 61.05 21.33 × 103 66.05
Paraffin dehydrocyclization to naphthene: Reaction-2
C6 33.11 24.37 − 9.12 − 5.24 × 103 − 43.64
C7 33.11 29.10 − 8.95 − 3.95 × 103 − 32.85
C8 33.11 27.81 − 8.34 − 3.91 × 103 − 32.55
C+9 33.11 29.76 − 8.12 − 3.37 × 103 − 28.00
Paraffin dehydrocyclization to naphthene: Reaction-3
C6 33.11 26.36 − 14.35 − 6.25 × 103 − 52.00
C7 33.11 25.75 − 13.65 − 4.73 × 103 − 39.30
C8 33.11 29.00 − 13.17 − 5.94 × 103 − 49.40
C+9 33.11 29.76 − 12.39 − 5.89 × 103 − 49.00
Paraffin dehydrocyclization to naphthene: Reaction-4
C5 33.11 30.75 14.39 8.38 × 103 69.73
C6 33.11 31.94 14.77 7.31 × 103 60.74
C7 33.11 33.43 14.63 7.31 × 103 60.75
C8 33.11 31.31 15.98 7.43 × 103 61.75
C9+
33.11 32.96 15.21 7.09 × 103 59.00
Paraffin dehydrocyclization to naphthene: Reaction-5
C5 33.11 29.83 16.18 9.22 × 103 76.67
C6 33.11 30.87 16.03 8.50 × 103 70.60
C7 33.11 32.95 15.47 8.08 × 103 67.20
C8 33.11 34.19 16.37 9.45 × 103 78.60
C9+
33.11 32.96 16.02 9.62 × 103 80.00
Paraffin dehydrocyclization to aromatic: Reaction-6
C6 18.86 16.87 35.73 7.99 × 103 66.50
C7 18.86 17.17 35.21 7.60 × 103 63.20
C8
for An = MX 18.86 17.32 34.13 6.79 × 103 56.51
for An = OX 18.86 17.89 34.21 6.85 × 103 56.95
for An = PX 18.86 17.01 34.16 6.81 × 103 56.70
for An = EB 18.86 16.96 34.63 7.17 × 103 59.66
C+9 18.86 18.90 33.56 6.36 × 103 59.29
Paraffin dehydrocyclization to aromatic: Reaction-7
C6 18.86 16.44 36.01 8.20 × 103 68.18
C7 18.86 16.90 35.47 7.79 × 103 64.83
C8
for An = MX 18.86 16.27 34.81 7.30 × 103 60.73
for An = OX 18.86 16.78 34.87 7.35 × 103 61.17
for An = PX 18.86 16.04 34.83 7.32 × 103 60.90
for An = EB 18.86 15.93 35.23 7.68 × 103 63.88
C+9 18.86 17.66 34.81 7.3 × 103 60.77
Isomerization of naphthenes and paraffins: Reaction-8
C6 23.81 15.11 − 5.20 − 2.06 × 103 − 17.10
C7 23.81 24.73 − 6.63 − 3.36 × 103 − 27.90
C8 23.81 25.78 − 6.83 − 3.51 × 103 − 29.20
C+9 23.81 26.13 − 7.16 − 3.73 × 103 − 31.00
Isomerization of naphthenes and paraffins: Reaction-9
C4 26 25.08 0.80 − 1.11 × 103 − 9.23
C5 26 24.89 1.17 − 0.83 × 103 − 6.94
C6 26 24.70 0.69 − 1.19 × 103 − 9.86
C7 26 24.15 1.27 − 0.76 × 103 − 6.45
C8 26 25.54 0.04 − 1.68 × 103 − 13.95
C+9 26 24.48 − 1.09 − 2.53 × 103 − 21.00

achieve higher purity of hydrogen in a more compact size than the conventional PSA or cryogenic separation counterparts. Hydrogen
separation technique using membrane can operate at a wide operational pressure and temperature, where the operating temperature
can be as high as 900 ◦ C [14,132,133].
Palladium membranes are studied extensively for their enhanced hydrogen selectivity and permeability. The wide-spread appli­
cation of Pd membranes is limited by the considerably higher cost of palladium itself. Different approaches have been adopted to
reduce the cost of Pd membrane like the application of thin film membrane and alloying palladium with other relatively cheap

14
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Table 10
Heat of reaction and rate constants for reactions associated with isomerization of aromatics in naphtha reforming reactions [117].

R P E/R a A B ΔH (kJ/mol H2)

Isomerization of aromatics: Reaction-10


n=1 OX PX 21.24 25.21 12.13 − 1.26 × 103 − 10.50
n=2 OX MX 21.24 24.50 11.00 − 2.11 × 103 − 17.60
n=3 MX EB 21.24 24.53 15.81 1.50 × 103 12.60
n=4 OX EB 21.24 25.17 15.54 1.30 × 103 10.84
n=5 EB PX 21.24 25.84 11.91 − 1.43 × 103 − 11.89
n=6 MX PX 21.24 24.47 13.7 − 0.085 × 103 0.71

Table 11
Heat of reaction and rate constants for reactions associated with transalkylation.in naphtha reforming reactions [117].

R P C E/R a A B ΔH (kJ/mol H2)

Isomerization of aromatics: Reaction-11


n=7 T B PX 21.24 16.68 16.41 1.95 × 103 16.20
n=8 OX T A9+ 21.24 15.48 15.28 1.10 × 103 09.16

Table 12
Heat of reaction and rate constants for reactions associated with cracking of paraffins naphthenes in naphtha reforming reactions [117].

E/R a ΔH (kJ/mol H2)

Cracking of paraffins: Reaction-12


C2 34.61 37.85 − 65.20
C3 34.61 39.95 − 53.60
C4 34.61 40.15 − 49.50
C5 34.61 41.60 − 47.80
C6 34.61 42.08 − 47.26
C7 34.61 42.48 − 46.69
C8 34.61 40.53 − 46.61
C9+ 34.61 43.85 − 46.15
Cracking of paraffins: Reaction-13
C4 34.61 36.15 − 40.30
C5 34.61 41.57 − 45.30
C6 34.61 40.53 − 37.40
C7 34.61 41.57 − 41.85
C8 34.61 41.51 − 36.00
C9+ 34.61 42.93 − 35.65
Cracking of naphthenes: Reaction-14
C6 34.61 42.15 − 45.45
C7 34.61 44.70 − 40.76
C8 34.61 43.90 − 41.00
C9+ 34.61 44.15 − 40.10
Cracking of naphthenes: Reaction-15
C5 34.61 40.23 − 68.74
C6 34.61 41.55 − 54.00
C7 34.61 43.75 − 52.71
C8 34.61 43.65 − 51.95
C9+ 34.61 44.15 − 50.40

Table 13
Heat of reaction and rate constants for reactions associated with hydrodealkylations s in naphtha reforming reactions [117].

m k E/R a ΔH (kJ/mol H2)

Hydrodealkylations.: Reaction-16
C7 1 n 17.92 7.64 − 41.81
C8
for An+1 = PX 1 n 17.92 5.57 − 42.32
for An+1 = EB 2 n− 1 17.92 5.55 − 20.02
C+9
for Ak = OX 1 n 17.92 8.91 − 52.57
for Ak = T 2 n− 1 17.92 5.57 − 30.80

materials (e.g. Ag, Au, Ni, Rh, Ru etc.) [134,135]. These also improved the permeability and stability of the membrane. The properties
and preparation of the various hydrogen-selective Pd-based membranes are explored in authors recent review work [15].
Separation of hydrogen in membrane reactors is driven mainly by the pressure gradient. The partial pressure difference of the
hydrogen on both sides of membrane dictates the separation process through Pd-membrane. The permeation of hydrogen increased

15
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Table 14
Lumped components and number of corresponding reactions for naphtha-reforming developed over the year.

Researcher(s)/year Number of reactions Number of lumped components

Smith [118]/1959 4 3
Jenkins and Stephens [119]/1980 78 31
Forment [120]/1987 81 28
Saxena et al. [121]/1994 40 22
Taskar and Riggs [122]/1997 36 35
Vathi and Chaudhuri [123]/1997 15 26
Ancheyta-Juárez and Villafuerte-Macías [124]/2000 71 24
Hu et al. [125]/2004 17 17
Hu et al. [126]/2004 51 21
Hou et al. [127]/2006 31 20
Hou et al. [128]/2006 17 18
Arani et al. [129]/2009 15 17
Honguin et al. [130]/2010 52 27
Wang et al. [131]/2012 86 38

with the increase in hydrogen partial pressure difference. The pressure gradient can be achieved through: (a) increasing pressure on
side of the membrane while maintaining atmospheric pressure on the other side, (b) maintaining atmospheric/elevated pressure on
one side and vacuum on the other side of the membrane and (c) introducing sweep flow in permeate side to reduce the partial pressure
and increase the pressure difference [136,137]. Typically argon, nitrogen and steam can be used as sweeping gas for hydrogen sep­
aration [138]. However, the application of steam as sweep gas is more widely used because of the higher heat capacity of steam,
availability in power plants and relatively simple hydrogen extraction from mixture. Permeated hydrogen in hydrogen-steam mixture
can be easily separated by the condensation of steam [139].
Beside the pressure gradient, temperature and resident time should also be regulated to maximize the hydrogen production [140].
Hydrogen flux through the membrane improves with both of these parameters. Furthermore, higher temperature is also necessary to
supply the necessary heat energy to activate the membrane for hydrogen permeation [15]. In order to present an overview on
hydrogen-selective membrane modeling, this section studies the geometries, governing equations and modeling parameters examined
in various CFD modeling works. Three parameters are usually used to express the performance of fuel-reforming and separation
process, which are fuel conversion, hydrogen yield and hydrogen removal/recovery rate as expressed in Equations (93)–(95),
respectively.
flow rate of fuel in reformer outlet
Fuel conversion rate [%] = × 100 (93)
flow rate of fuel in reformer inlet

flow rate o f H2 in products


H2 yield [%] = × 100 (94)
feed rate of fuel or maximum possible production of H2
[ ]
H2 permeation flow rate mol
H2 recovery rate [%] = [min ] × 100 (95)
Theoritical H2 production mol
min

3.1. Membrane geometries adopted in numerical modeling


The geometry of membrane reactor can also influence the performance in terms of hydrogen permeation. Several studies have been
conducted using Pd-based membranes on hydrogen separation and purification. Ghasemzadeh et al. [141] conducted a dehydroge­
nation analysis from ethanol steam reforming process with Pd-Ag membrane using the CFD model as shown in Fig. 3. The thickness,

Fig. 3. Layout of Pd-Ag membrane reformer examined by Ghasemzadeh et al. [141].

16
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

outer diameter and length of the Pd-Ag membrane are considered as 5 × 10− 5 m, 1 × 10− 2 m, and 7 × 10− 2 m, respectively. Sweeping
flow is incorporate in the annular section of the geometry to enhance the hydrogen permeation, while the feed is containing C6H12 and
H2 is supplied in the inner cylinder. Hydrogen separation using integration of Pd-based membrane also resulted in an improvement in
ethanol conversion. They also reported the potential of operating the membrane reactor in higher temperature and pressure than
conventional counterparts.
Ben-Mansour et al. [136] also utilized a simple two-dimensional geometry depicted in Fig. 4. Unlike the model by Ghasemzadeh
et al. [141] discussed above, the geometry of Ben-Mansour et al. [136] contained the feed gas mixture (CO, H2, CO2, and N2) in the
annular part and permeated hydrogen in the central cylinder. Aside from the temperature and sweep effect, the pressure effect in both
pressurization and vacuum method is investigated in their simulation. The results showed that at lower pressure in vacuum method,
the sweep flow can significantly increase (~25%) the hydrogen permeation. The application of high-pressure method is recommended
in the study to reduce the complicated arrangement requirement for hydrogen separation. Ghohe and Hormozi [142] used CFD in­
struments and examined the effect of geometric form on the flux of permeation of hydrogen through Pd membrane. The study reported
that the mean flux of a cylindrical shape is lower than that in conical shape counterpart.
Ghasemzadeh et al. [66] developed a CFD model to investigate the hydrogen separation in cylindrical dense Pd-Ag membrane. A
comparison of performance between auto-thermal methods and steam reforming of methanol in Cu/ZnO/Al2O3 catalyst was also
studied in the simulation. The reforming process takes place at the annular part of the reactor, while the inner membrane cylinder
contains the permeate zone. The results demonstrated a performance increase in auto-thermal reforming as compared to the wet/­
steam reforming counterpart. Moreover, the operating parameters like, pressure, temperature and sweep gas flow rate showed a
positive impact on hydrogen production and recovery.
The multi-tube membrane distribution model was investigated by Chen et al. [143]. Four membrane tubes are considered for their
study and the feed gas is supplied in crossflow configuration. They showed that the tandem arrangement of the tubes takes the lead to
reduce performance of each individual of the tubes at downstream, thus the system’s standard performance. An optimization study was
also carried out in their simulation work. The optimized configuration demonstrated a 12.2% rise in hydrogen recovery than the base
case. Lee and Lim [21] investigated the hydrogen production from dry-reforming of methane in multi-tube membrane reactor. The
reactor consists of two concentric cylindrical tubes, where the inner tube acts as heating tube and outer cylinder housed the reforming
zone as shown in Fig. 5. The annular section of the geometry contains the multi-tube arrangement, inside of which the sweep gas is
supplied to extract the permeated hydrogen. The results demonstrated that increasing the number of membrane tubes can significantly
improve the hydrogen flux subsequently enhancing the hydrogen recovery.
Abdi et al. [144] investigated the helical type Pd-membrane geometry and compared the results with that from straight one. Fig. 6
shows the adopted membrane configuration, where the feed gas (H2-N2 mixture) is supplied on the outer side of the membrane tube
and permeated hydrogen is in the inner side of the tube. The surface to volume radio in helical membrane is higher than the straight
membrane, which enhances hydrogen separation consequently raising the hydrogen recovery. However, for both the configurations,
the hydrogen recovery dropped with the increase in feed flow rate. The results also show that the helix pitch is more important in
improving hydrogen permeation than the coil diameter. Rakib et al. [86,87] examined the hydrogen permeation through plate-type
Pd-Ag membrane reactor for steam reforming of heptane, propane and methane, depicted in Fig. 7. The two-phase model geometry
consisted of six plate-type membranes inside a cylindrical reactor tube. At relatively lower temperature (~500 ◦ C), the membrane
reactor model demonstrated hydrogen yield comparable to that with the conventional counterpart at higher temperature (>750 ◦ C).
The geometries and corresponding reforming/separation configuration is summarized in Table 15.

Fig. 4. 2D axi-symmetric geometry investigated by Ben-Mansour et al. [136].

17
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Fig. 5. Multi-tube membrane reactor examined by Lee and Lim [21].

Fig. 6. Palladium-based hydrogen-separation modeling in straight and helical type geometry, investigated by Abdi et al. [144].

3.2. Governing equations for membrane-assisted H2 separation model


Since the flow velocity should be lower enough to allow the adequate residence time in the catalyst bed, the flow is often considered
as laminar in membrane reactor modeling [142,143,146]. In addition, the laminar flow condition is also considered for models with
separation-only approach [136,144]. Thus, considering laminar steady flow, the continuity and conservation equations are:

Continuity : ∇ • (ρ →
v ) = Sm (96)


Momentum equation : ∇.(ρ →
v →
v ) = − ∇p + ∇ • τ + ρ →
g +F (97)


Species transport equation : ∇ • (ρ →
v Y i ) = − ∇ • J j + Ri + S i (98)
( ∑ → )
Energy equation : ∇ • (→
v (ρE + P)) = ∇ • κeff ∇T − hj Jj + τeff • →
v + SE (99)

Moreover, Fick’s first law (Equation (63)) describes the steady-state transport system of H2 gases across the solid membrane [147,
148]. Henry’s law may be used for concentration as expressed in Equation (100). Sievert ’s law, known as renamed modified Henry’s
law, can be used for diatomic molecules as in Equation (101).
∂c
J= − D (100)
∂x

18
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Fig. 7. (a) Schematic of membrane integrated propane-steam reforming reactor [86], and (b) plate-type membrane panel [87] studied by Rakib et al. [86,87].

Table 15
Geometries adopted in H2 separation using membranes.

Researcher(s)/year Fuel/feed gas Reforming type Membrane Geometry type


material

Rakib et al. [86, C7H16, C3H8, CH4 and Steam reforming Pd/Ag 1D, two-phase model
87]/2011, 2010 H2O membrane
Ghasemzadeh et al. C6H12 and H2 No reforming, H2 separation only Pd-Ag 2d axisymmetric, annular cylindrical
[141]/2016
Ghasemzadeh et al. CH3OH/H2O and Auto-thermal and steam reforming Pd-Ag 2d axisymmetric, cylindrical-shaped and
[66]/2016 CH3OH/H2O/O2 with Cu/ZnO/Al2O3 catalyst conical-shaped membrane
Ghohe and Hormozi H2, CO2, and CO No reforming, H2 separation only Pd-based 2d axisymmetric, conical-shaped membranes
[142]/2019
Chen et al. [143]/2019 H2, CO2, and CO No reforming, H2 separation only Pd-based 3D stacked membrane in crossflow
configuration
Lee and Lim [21]/2020 CH4 and CO2 CO2 reforming with Rh/Al2O3 catalyst Pd-based 3D geometry, 9 membranes are distributed in
annular section of the reactor
Abdi et al. [144]/2020 H2 and N2 No reforming, H2 separation only Pd-based 3D geometry,
Straight and helical membrane tubes
Ben-Mansour et al. CO, H2, CO2, and N2 No reforming, H2 separation only Pd-Ag/α-Al2O3 2d axisymmetric, annular cylindrical
[136]/2020
Ben-Mansour et al. CH4 and H2O Steam reforming with Ni/γ-Al2O3 Pd-Ag/α-Al2O3 2d axisymmetric, annular cylindrical
[145]/2021 catalyst

c
S= (101)
Pgas

c
S= (102)
P0.5
gas

The species diffusion flux can therefore be written as follows over a membrane of thickness d as shown in Equation (103).
Furthermore, by plotting Q vs. t, permeability can be obtained from experimental permeation results.
∂Pn ΔPn φ ( n )
J = − DS ≈φ ≈ P − Pnp (103)
∂x Δx d f

Aφ ( n ) ( )
Q = JAt ≈ P − Pnp t = AK Pnf − Pnp t (104)
Δx f

19
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

∂Q ( ) φ
= AK Pnf − Pnp , where = (105)
∂t d
Permeance can be expressed according to the Arrhenius equation [149,150] as shown in Equations (106) and (107), where Ko is a
pre-exponential factor which is independent of the temperature (T), where K is the permeability, Ea is the energy of activation, and R is
the constant of the universal gas. This equation is right if n is independent of the T.
( )
− Ea
K = Ko exp (106)
RT

Ea
ln K = ln Ko − (107)
RT
Said et al. [44] investigated Pd-based membrane and for H2 separation following governing equation (Equation (108)). The H2 flux
in their study was considered according to the Sievert’s law. Here, PH2,M refers to hydrogen partial pressure in permeate zone, while
EH2 = 6.6 k/mol and AH2 = 0.4 mol/(m2 s bar0.5) [151].
( )
− EH2 ( 0.5 )
JH2 = AH2 exp PH2 − P0.5
H2 ,M (108)
Rg T

Ghasemzadeh et al. [110] studied Pd-Ag membrane with of 120 mm total length 120 mm and 10 mm outer diameter reinforced by
tubular support. The Sieverts’ law is used to calculate the hydrogen flux (JH2). The permeation flux of hydrogen in given in equation
(109), where EH2 = 10.58 kJ/mol and Pe0 = 0.54 mol/m2 s bar0.5 [152].
( )( )
− E
Peo • exp RTH2 P0.5
H2 , permeate
JH2 = (109)
d
Smith et al. [153] performed water gas shift reaction modeling in membrane reactor using CFD analysis. The permeation of H2
through the wall of membrane is developed using the permeation equation (110), where the trans-membrane partial pressure of
hydrogen adjoining to the membrane is dependent on it. Since the hydrogen permeation resistance through the support of alumina is
inconsequential and the palladium membrane thickness is 70–75 μm with infinite hydrogen selectivity, thus, the membrane can be
assumed of zero thickness.
( )
− 5833.5 ( )
QH2 = 2.95 × 10− 4 exp mol m m− 2 s− 1 Pa− 0.5 (110)
T
The constant hydrogen flux permeating through the membrane was assumed by Lee et al. [154]. However, it is evident that the flux
should not be constant, as the hydrogen concentration varies depending on the non-uniform temperature. Nonetheless, constant
hydrogen flux can be considered in order to simplify the model and concentrate on membrane position variance. The Sieverts rule for
hydrogen flux through the membrane is followed by the commonly used Pd-based membrane [154]. Where the thickness of the
membrane is d, the constant of membrane H2 permeability is Q, E is the obvious energy of activation of membrane, PH2 ,P and PH2 ,R are
H2’s partial pressure on the side of the permeate and retentate respectively, and permeance is K. Experimental analysis of the recent
advancement in membrane fabrication reports that with increased the efficiency, the partial pressure exponent can be as high as one.

Qe( RT ) ( √̅̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅̅̅ )


− E
( √̅̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅̅̅ )
JH2 = PH2 ,R − PH2 ,P = K PH2 ,R − PH2 ,P (111)
d

Fig. 8. Permeation of oxygen through an ion-transport membrane by Xu and Thompson [170].

20
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Soria et al. [155] examined Pd-Ag membrane for H2 separation application,where the separation factor (fH2) for hydrogen yield was
defined as
∑N
k=1 mH2 ,k
fH2 = ∑N (112)
k=1 mH 2 ,k
+ mout
H2 ,N+1

Fig. 9. (a) Example of a multilayered Pd-based membrane, (b) schematic of the multilayered membrane and (c) mass transfer mechanisms of all the elementary steps
involved in the permeation process [183].

21
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359


where k represents each sub-separator Nk=1 mH2 ,k symbolizes the sum of the molar flow rate of H2 in the stream of permeate of every
sup-separator while moutH2 ,N+1 refers to the outlet molar flow rate of hydrogen in the last sub-reformer.
Iulianelli et al. [156] studied Pd-Cu membrane supported by Al2O3 in temperatures between 300 and 470 ◦ C under 150–250 kPa
feed pressure. The standard perm-selectivity of H2 in their study was calculated considering the other feed gases can be shown in
Equation (113). Here, JH2 and Ji characterize the permeating flux of H2 and that of another separate gas, respectively. The hydrogen
permeating flux of H2 can be evaluated using Equation (114). Where PnH2 − retentate and PnH2 − permeate and d correspond to the hydrogen
partial pressures permeate and retentate side and membrane thickness, respectively. The term ’n’ refer to the pressure exponent and
signifies the dependence factor of the permeating flux of hydrogen on the partial pressure of H2. An Arrhenius-like equation can
demonstrate the dependency of the permeability of H2 on the temperature, as shown in Fig. 8, where PH2,0, T, EA and R are
pre-exponential factor, temperature, the apparent activation energy and the universal gas constant, respectively [157].
JH2
αH2 /i = , (i = N2 , CO2 , CH4 ) (113)
Ji
( )
PH2 PnH2 − retentate − PnH2 − permeate
JH2 = (114)
d
( )
EA
PH2 = PH2 ,0 exp (115)
RT
Hydrogen permeation and stability of 3.5–17 μm thick composite Pd77Ag23 membrane on porous alumina substrates was inves­
tigated by Barison et al. [158]. They recommended the first law of Fick to designate atomic diffusion of hydrogen as a function of
diffusion coefficient and concentration gradient through a homogeneous metal phase, and the law of Sievert that can be used under
specific circumstances to state the relationship between pressure and concentration, Equation (116) was obtained to demonstrate how
the flux of hydrogen (j) that varies with pressure and temperature.
a E
∅ ( √̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅̅̅ ) ∅0 eRT ( √̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅̅̅ )
j= Pfeed − Pperm = Pfeed − Pperm (116)
d L

Here Ea is the activation energy for the whole process, R is the gas constant, Pperm and Pfeed are the hydrogen pressure at the permeate
and at the feed sides, respectively, d is the thickness of the membrane, φ is the permeability, which is defined as the rate of gas
permeation per unit area, per driving force unit (pressure in this case), per membrane thickness unit, Ø0 is the pre-exponential factor
associated with metal permeability. For thick membranes that is self-sustained (generally >10 μm), this equation is valid. In particular,
a more common formula has been used to estimate the permeance (F/L) of thin palladium-based membranes, as shown in Equation
(117) [158].
∅( n )
j= Pfeed − Pnperm (117)
L
This formula reflects a general approach, but with some limitations, to take into account some phenomena that can go ahead to the
best fit of flux data vs (Pnfeed − Pnperm ) with an exponent n higher than 0.5. They include resistance to mass transfer in porous substrates,
surface resistance to H2 desorption and absorption, Knudsen or viscous flow through pores and leaks and film defects and diffusion of
hydrogen as molecular H2 along grain boundaries [159,160]. The term n was reported to reach 1 in thin Pd-based membranes
(thickness: 1–7 μm), and this was usually attributed as the rate-limiting step to surface phenomena [161]. Said et al. [42] used CFD
model for Solar molten salt heated membrane reformer for natural gas upgrading and hydrogen generation. For H2 separation by
membrane he used the following equation
( )
− E H2 ( n )
JH2 = AH2 exp PH2 − PnH2 ,M (118)
Rg T

The H2 permeation activation energy (EH2 = 11 kJ = mol) and membrane permeability (AH2 = 5.6 mol/m2s bar0.5). Temperature
range 600K–873K but decreasing the temperature below 773 K is not favorable, although reasonably high conversions of 0.5–0.7 and
H2 recovery of 1.8–2.6 are obtained. Ni/Al2O3 catalyst, 5 μm supported Pd film membrane is used. outlet pressures of the reforming
and membrane zone were set to 10 and 1 bar, respectively.
Basile et al. [162] investigated methane steam reforming in a Pd-Ag membrane reformer. Here, the total retentate pressure is kept
constant at 9.0 bar, while the total permeate pressure is varied between 5.0 and 9.0 bar. The reaction is conducted in 450 ◦ C, where n
varies within 0.5–1. The permeability of the membrane in their study was evaluated using Equation (119) below.
( )
PeH2 PnH2 ,retentate − PnH2 ,permeate
JH2 = (119)
d
Ghasemzadeh et al. [163] investigated H2 production by low pressure methanol steam reforming in a dense membrane reactor.
Pd-Ag membrane with a thickness of 50 μm was used in their study where the hydrogen flux is calculated with the expression presented
in Equation (120). The activation energy (Ea) was considered as 10 580 J/mol in relatively low pressure (1.5–2.5 bar) at 280 ◦ C.

22
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

( E )( n )
PH2 ,0 exp RTA
P H2 − retentate − PnH2 − permeate
JH2 = (120)
d
Boeltken et al. [164] studied ultra-compact micro-structured methane steam reformer with integrated Pd-membrane for in-situ
production of pure hydrogen. The H2 flux JH2 was measured and according to Sieverts’ law, the permeance JH2 of the membrane
was calculated using Equation (121). The expression for the H2 permeability of the 12.5 μm Pd membrane in their study was proposed
as shown in Equation (122) for 773 and 823 K temperature and retentate/reformate pressure of 6–12 bar.
( )
Q PnH2 − PnH2 ,M
JH2 = (121)
d
( ) /
− 1460.2 ( )
Q = 1.58 × 10− 7 exp mol msPa0.5 (122)
T

Table 16
Comparison of Governing equation of H2 flux by different research.

Researcher(s)/ H2 yield Activation ΔT ΔP n d, μm membrane


year. energy (Ea ), permeance to
kJ/mol hydrogen
6
Pomerantz and ∂c 13.0 – 1.2–1.7 2.8 × 10− mol/
J = − D
Ma/2011 ∂x m2 s Pa
[169]
Ben-Mansour ∂Pn ΔPn φ n – 1 1 –
J = − DS ≈φ ≈ (Pf − Pnp )
et al., /2020 ∂x Δx d
[136]
( )( )
Said et al., /2015 − EH2 6.6 0.5 AH2 = 0.4 mol/
JH2 = AH2 exp PnH2 − PnH2 ,M
[44] Rg T m2 s bar0.5
( )( )
Ghasemzadeh − EH2 10.5 0.5 50 AH2 = 0.54 mol/
AH2 • exp PnH2 ,permeate
et al., /2019 RT m2 s bar0.5
JH2 =
[110] d
( ) –
Byron et al., − 4 − 5833.5 – 70–75 –
QH2 = 2.95 × 10 exp
/2011 [153] T
Lee et al., /2018 − E – 0.5 – –
[154] Qe RT √̅̅̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅̅̅̅
JH2 = ( PH2 ,R − PH2 ,P ) =
√̅̅̅̅̅̅̅̅̅̅d̅ √̅̅̅̅̅̅̅̅̅̅̅
K( PH2 ,R − PH2 ,P )
∑N –
Soria et al., /2019 k=1 mH2 ,k
– – –
[155] fH2 = ∑N
k=1 mH2 ,k + mH2 ,N+1
out

6
Iulianelli et al., PH2 (PnH2 − retentate − PnH2 − permeate ) 18.5 300–470 ◦ C 50 kPa 0.52 20 1.8 × 10−
JH2 =
/2019 [156]
( )d
EA
PH2 = PH2 ,0 exp
RT
[158]Barison j =
φ n
(P − Pnperm )
– 0.8 3.5–17 –
et al., 2018 L feed
( )
[42]]Said et al., − EH2 11 600–873 K 100 KPa 0.5 5 AH2 = 5.6 mol/
JH2 = AH2 exp (PnH2 − PnH2 ,M )
2016 Rg T m2 s bar0.5
Basile et al., /2010 PeH2 • (PH2 ,retentate − PnH ,permeate )
n 450 C◦
4.0 bar 0.5–1 50
[162] JH2 = 2

d
Ghasemzadeh JH2 = 10.58 280 ◦ C 1.5–2.5 bar 0.5 50 2.69 × 10− 7
( )
et al., /2013 EA mol/s cm atm0.5
PH2 ,0 exp (PnH2 − retentate − PnH2 − permeate )
[163] RT
d
Boeltken et al., Q(PnH2 − PnH2 ,M ) 12.14 773 & 823 6–12 bar 0.5 12.5
JH2 =
/2014 [164] d K
Q = 1.58 ×
( )
− 1460.2
10− 7 exp mol/(msPa0.5 )
T
Baloyi et al., JH2 = PH2 ,0 (PnH2 − retentate − PnH2 − permeate ) 320–430 ◦ C 0.5 to
/2016 [165] 1
Caravella et al., JH2 = (1 − CPC)πMembrane DFBulk 300–500 ◦ C 400–1000 5–100
/2009 [166] (
πBulk
) kPa
CPC = 1 − Membrane
π
√⃒̅̅̅̅̅̅̅̅Bulk
DFBulk = Δ PH2 ⃒
7
Zhao et al., /2017 JH2 = QH2 (PnH2 − retentate − PnH2 − permeate ) 9.21–9.84 298–673 K 50–500 0.59 to 3.5 7.0 × 10−
[167] kPa 0.77
7
Gao et al., /2005 Q (PnH2 − retentate − PnH2 − permeate ) 15.4 480–753K 1.0 10 1.1 × 10−
JH2 =
[168] d

23
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

Baloyi et al. [165] analyzed The production of hydrogen through Pd-Ag (77 wt% Pd-23 wt% Ag) membrane in water gas shift
reactor. The hydrogen flux in their study is calculated using Equation (123) for 20 μm thick Pd-Ag membrane at different temperatures
(320, 380 and 430 ◦ C). Here, PH2,0 is hydrogen permeability in mol s− 1m− 1 Pa− 1. The “n’ value was found to be 0.523 at 320 ◦ C, 0.65 at
380 ◦ C and 0.62 at 430 ◦ C. the physical meaning of the “n” value, if n = 0.5, bulk diffusion of hydrogen is the rate limiting step, “0.5 <
n < 1” the surface contamination is the rate limiting step and n < 0.5 the thickness of the membrane plays a role.
( )
JH2 = PH2,0 PnH2 − retentate − PnH2 − permeate (123)

Caravella et al. [166] analyzed effect of concentration polarization on the hydrogen permeation through Pd-based membranes in
various membrane thickness (5, 15, 50, 100 μm) under different temperatures (300, 350, 400, 450, 500 ◦ C) mixture pressures
(400–1000 kPa). In their study, the hydrogen flux and concentration polarization coefficient (CPC) were defined according with
Equations (124)–(128), where πMembrane, πBulk and DFBulk are the membrane and bulk permeances and the permeation driving force of
bulk, respectively.

JH2 = (1 − CPC) × πMembrane × DF Bulk (124)


( )
J H2
CPC = 1 − (125)
πMembrane × DF Bulk
√⃒̅̅̅̅̅̅̅̅Bulk
DF Bulk = Δ PH2 ⃒ (126)

JH2
πBulk = (127)
DFBulk
( )
πBulk
CPC = 1 − (128)
π Membrane

Zhao et al. [167] investigated the low-temperature stability of body-centered cubic PdCu membranes. Hydrogen flux in their study
was evaluated using Equation (130), where n is the pressure exponent and increases with temperature (i.e., 0.59 at 673 K to 0.77 at
298 K).
( )
JH2 = QH2 PnH2 − retentate − PnH2 − permeate (129)

Gao et al. [168] studied the electroless plating synthesis, characterization and permeation properties of Pd–Cu membranes sup­
ported on ZrO2 modified porous stainless steel. The 10 μm thick Pd46Cu54/ZrO2-PSS membrane exhibited an infinite separation factor
for H2 over N2, with H2 permeance of 1.1 × 10− 7 mol/m2sPa at 753 K. The activation energies for hydrogen permeation and hydrogen
pressure exponents are respectively 14.5 kJ/mol and 0.6 for Pd84Cu16/ZrO2-PSS membrane and 15.4 kJ/mol and 1 for Pd46Cu54/Z­
rO2-PSS composite membrane. The H2 flux was calculated by Equation (131) presented below. A summary of the discussed permeation
flux expressions is presented in Table 16.
( )
Q PnH2 − retentate − PnH2 − permeate
JH2 = (130)
d

3.3. Hydrogen permeation modeling


Numerical modelling is an essential tool in addressing problems that are difficult are difficult to observe directly. Permeation
through selective membrane is one such problem where different aspects of permeation can be examined using numerical modeling.
Semi-permeable membranes are used to deal with separation or purification of oxygen, hydrogen, nitrogen, CO2, water etc. Oxygen
separation is typically done through ion-transport (ITM) membrane, which is investigated using numerical modeling by several re­
searchers in last decades. Xu and Thompson [170] developed an explicit model to investigate the permeation of oxygen through ion
conducting membrane. The study emphasized on the surface exchange kinetics on both sides of the membrane where resistance to the
oxygen permeation was quantitatively distinguished from the bulk diffusion resistance. The oxygen flux in the developed model was
correlated with the readily measurable variables like the operating temperature in the vicinity of the membrane thickness and the
gradient of oxygen pressure. Furthermore, the oxygen flux in the study of [170] was measured on a 3.99 mm thick disk in order to
obtain the parameters for the model like the surface exchange rate constants (kf and kr) and the oxygen vacancy bulk diffusion co­
efficient (DV). This measurement was carried out through changing the temperature and partial pressures of oxygen separately on both
sides of the membrane. The results showed the recombination of oxygen-ion acts as limiting factor the permeation at lower tem­
peratures (~750 ◦ C) whereas at high temperatures (~950 ◦ C), the bulk diffusion rate controls the oxygen permeation. This is because
of the activation energies for the oxygen vacancy diffusion along with the surface exchange rates, the values of which were estimated at
74 and 227, 241 kJ/mol, respectively.
Moreover, the bulk diffusion and the permeation resistance of surface exchange kinetics were evaluated quantitatively by intro­
ducing surface exchange coefficients at both high and low oxygen pressure sides of the ion-conducting membrane. The investigation
also showed a 5-step operation for the permeation of oxygen through the membrane from high pressure (side A) to low pressure side

24
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

(side B), as shown in Fig. 8. As illustrated in the Figure, gas-solid mass transfer of oxygen gas occurs in steps 1 and 5, surface exchange
between the molecular oxygen and oxygen vacancy takes place in steps 2 and 4, lastly, the oxygen vacancy bulk diffusion occurs in step
3. Besides the recombination of oxygen-ion and bulk diffusion rate, the gas-solid mass transfer sometimes partially controls the oxygen
permeation, which may occur in certain conditions including high ionic conductivity, thickness of the membrane and/or faster surface
exchange rates. The following Equations (131a) and (131b) was developed for the oxygen permeation in the study which involves bulk
diffusion and resistances of surface exchange.
( )
′′ − 0.5 ′ − 0.5
k r P O2 − P O2
JO2 = 1 2L
(131a)
kf k
′ +
Dv
+ k1ex′′
ex

ΔPO2
JO2 = (131b)
R′ex + Rdiff + R′′ex

Several assumptions were made in Xu and Thompson’s [170] oxygen permeation mode which includes: (a) 1D model, (b) mobile
charges encompasses entirely of electron hole and oxygen vacancy, (c) flux of oxygen vacancies controls the permeation of oxygen
across the membrane, subsequently increasing the electronic conductivity than the ionic conductivity in perovskites and decreasing
the ionic transference number to zero, (d) concentrations of electron holes are constant on both sides of the membrane, while the low of
mass action is considered for the remaining components, (e) forward reaction occurring on the feed side of the membrane is identical to
the reverse reaction on the other side (sweep side) of the membrane, (f) diffusion coefficient remains constant and is a function of
temperature only. Based on these assumptions, the expression for oxygen permeation (in Equations (131a) and (131b)) is further
modified into Equation (132). The coefficient for oxygen vacancy diffusion mostly depends on the temperature and the reactions rates
(both forward and backward). Therefore, the Arrhenius type equation can be employed to express the dependence of temperature for
these parameters, presented in Equations (133)–(135).
( )
kr 1
kf P′′ 0.5
− ′10.5
PO
(132)
O2
2
JO2 = 1 2L 1
′ 0.5 + Dv + k P′′ 0.5
kf P O2 f O2

( )
ED
kr = D0v × exp − (133)
RT
( )
Ef
kf = kf0 × exp − (134)
RT
( )
Er
kr = kr0 × exp − (135)
RT
Several other investigations have been carried out in recent years to comprehend the fundamental coupling between fuel con­
version and oxygen conversion. The works done by Kirchen et al. [171], and Hunt et al. [172,173] present the development of new
experimental ITM reactor along with the insight on optical and sampling processes of the reaction zone. Along with these a flow
configuration, modular geometry and measurements of the membrane temperature are also provided, all of which are necessary for a
detailed reactive CFD modeling. Initially, the experimental results showed an increase in oxygen flux when the sweep flow was
changed from inert to reaction gas. Like the previous works, the rise in oxygen flux was also due to the decreased partial pressure at the
surface of the membrane. On the other hand, a reduction in local oxygen concentration was observed not as low as that from mixed
reactor outlet stream. This indicates that application of reacting flow should be done carefully in order to infer the thermodynamic
state at the membrane surface. Local measurements of gas composition were done in their study which revealed that local gas
compositions were dissimilar to values obtained from the outlet. Kirchen et al. [171] also evaluated the forward and backward/reverse
reaction rates (i.e. kf and kr) as presented in Equations (136a) and (136b), respectively.
( )
Ea,f
kf = Af × exp − (136a)
RT
( )
Ea,r
kr = Ar × exp − (136b)
RT
Moreover, the bulk diffusion in oxygen permeation can be determined using the Wagner equation, as follows:
( )
− Dv ( ′′ ′ ) − Dv P′′O
Jd = − μVo − μVo = ln ′ 2 (137)
2L 2L PO2

A more complicated first-order flux model was tested in a study by Hunt et al. [174]. The developed model directly was contrasted
with the sensitivity of the lattice oxygen ion concentration, usually assumed in case of zero-order flux model. The variation obtained

25
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

from the proposed flux model parameter was marginal for LCF, which implied that the change in lattice oxygen ion concentration
assert minimal effect on the surface exchange reaction rates (both kf and kr). Nevertheless, the examined model is a valid method which
should be tested for other types of membrane. Taking the law of mass action into consideration, the dynamic equilibrium equations for
the surface exchange reactions is presented in Equation (138). This expression can be further reduced into Equation (139), where
1/k”ex and1/k’ex in represent the surface exchange resistances, while 2tm/DV corresponds to the bulk resistance [174].
( 0.5 )

3
D k k PO2 − P′′O2 0.5
5VM v f r
JO2 = ( ′ 0.5 )( ) [ (
′ 0.5
) ] (138)
2L kf PO2 + kr kf P′′O2 0.5 + kr + Dv kf PO2 + P′′O2 0.5 + 2kr

( − 0.5
)
0.5 ′
kr P′′O2 − − P O2
JO2 = 1
(139)
kf ′
kex
+ D2Lv + k1ex′′

(140)
′ ′ 0.5
kex = kf PO2

(141)
′′ 0.5
kex = kf P′′O2
( ′ )
(142)
′ ′
JO2 = kex Cv − Cveq

( )
′′
JO2 = kex Cv′′ − Cveq
′′
(143)

The oxygen permeation flux obtained using Equation (132), presented by Xu and Thomson [170], was found less accurate in a work
by Behrouzifar et al. [175]. The study recommended the introduction of three correlation terms into the expression in order to improve
the accuracy. The first correlation term considers the deviation from the elementary reactions while the other two correlations take
care of the influence of feed and sweep gas flow rates. Based on that Behrouzifar et al. [175] developed a mathematical model to
examine the permeation of oxygen through BSCF membranes. The developed model was reported to have some distinguishing features
over the other oxygen permeation model. The model being an explicit function of all operational parameters, considered both surface
exchange kinetics and bulk diffusion for permeation, and capable of evaluating oxygen vacancy bulk diffusion coefficient, permeation
resistances and surface exchange rates. Both forward as well as backward reactions were considered non-elementary. The reaction
rates Feed and sweep flow rates are included in the model using the modified partial pressure definition, while the flowrates, properties
and hydrodynamics of the fluids were taken into consideration by means of Reynolds number. The dimensionless characteristic
thickness of the membrane was defined and evaluated as a criterion for calculating limiting permeation step. Finally, the integration of
the three correlations resulted in acceptable agreement between the model and corresponding experiment.
( )
kr 1 1
kf ′′ 0.5 − ′ 0.5
P O PO
(144)
2 2
JO2 = 1 2L 1
′ 0.5 + Dv + k P′′ 0.5
kf P O2 f O2

Unlike the oxygen permeation, the mechanism of hydrogen permeation through metallic (e.g., Pd) membrane involves dissociative
adsorption at the high-pressure feed side and recombination through desorption at the lower pressure side. Diffusion also takes place
during hydrogen permeation operation, which is regarded as a rate-limiting step for comparatively thicker membrane, which was
common in earlier permeation studies. The hydrogen permeation follows the Seivert’s law where flux is inversely proportional to the
thickness of the membrane and has a square root dependence on the partial pressure of hydrogen. Even though thinner membranes are
favorable to the hydrogen permeation, as the thickness reduces and approaches the ‘critical thickness’, the solid state diffusion be­
comes rapid to the point where it impacts the other rate processes subsequently limiting the permeation rate and deviate from the
Seivert’s law. This deviation can be attribute to several factors such as the surface processes [176–178], surface poisoning [179,180],
and grain boundaries [181].
Numerous investigations have been carried out in recent years on the fundamental surface science of Pd-H system due to the
importance of palladium (Pd) as a catalytic and hydrogen storage material. However, the usability of these findings has not been tested
for the dense membrane. A detailed rate model for the permeation of hydrogen in palladium was presented in a work by Ward and Dao
[182], where the model description and related concept of the rate processed necessary in permeation were outlined. The kinetic
model detailed in prior published work was extended in their investigation through explicit modeling of the kinetic steps and pro­
ducing acceptable results for all constants and rate parameters. The modified model proposed by Ward and Dao [182] considered the
hydrogen permeation in different stages in series. The stages include: (a) molecular transport of hydrogen from the bulk gas stream to
the gas layer on the membrane surface adjacent to the high-pressure side, (b) dissociative adsorption on the high gas surface, (c)
transition of the atomic hydrogen into the bulk metal, (d) atomic diffusion through the bulk metal, (e) transition from the bulk metal to
the surface on the low partial pressure side, (f) recombining desorption from the surface, and (g) gas transport away from the surface to
the bulk gas. Intrinsic forward and reverse rate are considered for all the stages in the model. If one of the steps becomes slower that the
other steps, it can bottleneck the overall rate of permeation. Furthermore, the rate expressions for each of the steps must be known in
order to develop a rigorous hydrogen permeation model.

26
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

The model developed by Ward and Dao [182] incorporated reasonable rate parameters, which were calculated based on the surface
science and also from the literature. The results show that the absence of diffusion and external mass transfer resistance are the
rate-defining process for clean Pd under moderately high temperature (≥573 K) even for the thickness of the membrane approaching 1
μm. Moreover, desorption becomes the controlling parameter for hydrogen permeation in low-temperature operation, while the
adsorption bottlenecks the permeation process in very low partial pressure of hydrogen at upstream or in case of surface contamination
which leads to considerable decrease in stacking on the membrane. External gas-phase mass transfer efficiency is vital for the apparent
Sievert’s law behavior and subsequently the permeation flux for palladium membrane with thickness ≤10 μm in lower partial pressure
of hydrogen or whit porous support. A central difference approximation to the derivative about the average pressure in the mass
transfer region technique was utilized to estimate the decrease in driving force at upstream, using the following expression.
(P1 − P1s ) JRT
P0.5 0.5
1 − P1 = = (145)
2P−1 0.5 4h1 P−1 0.5
Alternatively, a similar equation for the downstream surface can be formulated for the comparative decrease in Sievert’s law
driving force (because of the external mass transfer), which is provided in Equation (146), where hi is much greater than the term
JRT
4P− 0.5 (P0.5 − P0.5 )
.
1 1 2

( )
P0.5 − P0.5 JRT 1 1
1s 2s
= 1 − ( 0.5 ) − (146)
P0.5
1
0.5
− P2 0.5
4 P1 − P2 − 0.5
h 1 P1 h2 P2− 0.5

The characteristics of both dense and porous type hydrogen selective metallic membranes were reviewed and outlines in an
investigation by Al-Mufachi et al. [133]. The study compared both type of membrane and reported that the dense membranes are
capable of delivering higher hydrogen flux and better selectivity than the porous counterparts. Among the different metal membranes,
palladium-based ones are gaining more research focus in recent times due to their inherent ability to provide better hydrogen
permeability with higher purity along with respectable mechanical properties. However, the use of pure palladium membrane in­
creases the cost of the membrane making it not feasible for many use cases. The material cost can be minimized by alloying the Pd with
comparatively inexpensive elements like copper (Cu). Besides reducing the cost, alloying can make the membrane durable, resistant to
H2S poisoning and improves the permeability of hydrogen. In addition, reducing the membrane thickness of these Pd-based mem­
branes can result in further improvements in hydrogen permeability.

∂P1/2 1/2
ΔPH2
(147)
H2
J = DS ≅ DS
∂x Δx
A general expression for estimating the hydrogen flux through a membrane is provided in Equation (148), while the formula to
evaluate the φ is given in Equation (149).
( )
∅ Pn1 − Pn2
J= (148)
x
( )
− Eφ
φ = φ0 × exp (149)
RT
A model incorporating several elementary steps of permeation (as provided in Table 17 and illustrated in Fig. 9) operation for
hydrogen transport in a supported Pd-based membrane was developed by Caravella et al. [183]. The model considered self-supported
membrane and aimed to improve the study by Ward and Dao [182]. External mass transfer in the multicomponent gaseous phases
based on the Stefan–Maxwell equations was included in the model for both sides of the membrane. Poiseuille, Knudsen and ordinary
diffusion were considered in the transport of the multicomponent gas mixture in a multi-layered porous support. Irreversible ther­
modynamic theory was employed for the diffusion in the Pd-alloy layer where the chemical potential of hydrogen was regarded as a
driving force for diffusion. Other parameters like desorption, adsorption, transition from Pd-based surface to Pd-based bulk and
vice-versa were defined by the expression provided by Ward and Dao [182]. The asymmetric support considered in the model con­
stitutes of five layers. The permeation steps are along with their respective influences are separated in the developed model. The

Table 17
Elementary steps for hydrogen permeation through supported Pd-alloy membranes [183].

Location where the elementary steps are considered Elementary steps

Feed side 1. Fluid phase mass transfer of the film near membrane interface.
2. Decomposition of hydrogen into atoms the membrane surface, i.e., adsorption
Inside Pd-alloy membrane layer 3. Transition of hydrogen atom from Pd-alloy surface to Pd-alloy bulk.
4. Transport of hydrogen atom within the Pd-alloy bulk
5. Transition of hydrogen atom from Pd-alloy bulk to Pd-alloy surface on the permeate side.
Permeate side 6. Recombining the hydrogen atoms from Pd-alloy surface, i.e., desorption
7. Fluid phase mass transfer through multi-layered porous support
8. Fluid phase mass transfer of the film near membrane interface.

27
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

distribution of hydrogen partial pressure in the Pd-alloy, support and adjacent gaseous films were evaluated in terms of temperature,
and membrane thickness. The hydrogen flux along with the impact of each of the permeation step were also examined in their study.
Upon comparing the model with that from Ward and Dao [182], it was found that hydrogen permeation increases with the temperature
while the previous model indicated a non-monotone type correlation between the hydrogen flux and temperature. Thus, the model
developed by Caravella et al. [183] can be a suitable tool to enhance the flux prediction (In Equation (150)) and optimize the per­
formance of actual hydrogen separation system.
⎡ ⎛ ⎞⎤
( ) − 2
Nb DH ⎣ 1 − ξ1 WHH Nb YV H ( 2 )
JH = ln + ⎝ + 2 ⎠⎦
ξ1 − ξ1 (150)
δpd− layer 1 − ξ2 RT 3RT

Here, ξ is the concentration of atomic hydrogen, WHH is the energy for interaction between two atomic hydrogens in the Pd bulk (J/

mol), Nb is the concentration of octahedral sites in Pd bulk (molPd/m3) and V H is the partial molar volume of hydrogen in Pd bulk (m3/
mol).
Surface adsorption, bulk diffusion, solution into bulk metal, dissolution and desorption; all these parameters play important roles in
enhancing the permeability of a Pd-based thin-film membrane. Thus, a complete understanding of these rate-controlling parameters is
imperative for achieving optimized permeation from thin-film type Pd-based membranes. Hydrogen permeation through a
Pd0.77Ag0.23 alloy was examined by Bhargav et al. [184] by extending the diffusion models of pure Pd membrane from the literature
with a micro-kinetic model for surface and bulk reactions between H2 and the alloy. The develop model was assessed critically with the
measurements from the literature. They study reported that the effect of surface processes becomes more vital with the decrease in the
thickness of the film (<10 μm). The micro-kinetic model developed in the study involves the adsorption, diffusion, dissolution and
desorption of hydrogen with the Pd0.77Ag0.23 alloy membrane. The solubility of hydrogen in the alloy membrane was identified as one
of the critical steps for the prediction of hydrogen flux.
To simulate a permeation process through the membrane, the discontinuity of the flow must be taken into consideration. This
discontinuity can be dealt with introducing the source and sink terms in the model. However, a recent study by Abdi et al. [144]
demonstrated a way around the use of source/sink terms and using which multi-layered membrane with different diffusion mechanism
can be simulation maintain the desired accuracy. In addition, the thickness of the membrane is considered as a solution domain and
geometry does not need to be separated adjacent to the geometry. Their developed 3D model uses user-defined function (UDF) and
user-defined scalars (UDS) for the hydrogen separation. Two different geometries, straight and helical, were tested in the model, where
the helical geometry demonstrated 13% increase in hydrogen recovery and 20% increase in average flux over the straight counterpart.
Rossi et al. [185] prepared a mathematical model to investigate the hydrogen separation from syngas, which can also be utilized in
examining the parameters that characterize the mass transfer phenomena like permeability, limiting gas and surface coverage.
Furthermore, the study showed that the model can be used in evaluating separation performance even coupled with water-gas shift
reaction kinetics. Several parameters including the adsorption/desorption of hydrogen, permeation of hydrogen into Pd-bulk and
porous layer and the kinetics of different species competitive desorption on Pd surface.
Lee et al. [186] investigated the validity of permeability trends from previous published works for Pd/BCC composite membranes.
The results showed that the surface reactions significantly affect the permeation process of Pd/BCC composite membranes. They
proposed a new model which incorporates the surface reactions at catalytic layer and bulk diffusion through the metallic layer. The
developed model with this consideration demonstrated a good agreement with the experimental permeation behavior of the Pd/BCC
composite membranes. In addition, the study also established the correlation between the diffusivity coefficients of the BCC metals and
the temperature. Sievert’s las was utilized in their calculation assuming diffusion through the bulk metal layer as a rate-defining step.
The expression for hydrogen flux from their model is as follows, where PE represents the permeability. Moreover, the permeability,
here, is the product of solubility and diffusivity.
Since the study by Lee et al. [186] was focused on investigating the influence of the catalytic layer on the dense metallic mem­
branes, therefore, an in-depth examination was carried out on the permeation characteristics of Pd catalytic layers over commercial
tantalum (Ta) tubes, i.e. Pd/Ta composite membrane. A time-lag analysis was also performed based on the temperature and pulse
pressure changes, the results of which indicates that the association/disassociation related surface reaction of hydrogen on the cat­
alytic layer can adversely impact the permeation through the studies composite membrane. The permeability of the composite
membrane [187–189] on the other hand, was explained by preparing a new model that encompasses the surface reactions along with
diffusion of atomic hydrogen through the metal lattice structure. They reported that the catalytic layer effect can significantly change
impact the permeation process and should be included in investigation the separation process. The diffusivity for the composite
membranes was also expressed in the study as a function of temperature for Ta.

4. Conclusion
Reaction kinetics of fuel reforming for hydrogen production and hydrogen separation through Pd-membrane are extensively
reviewed in this study. Reforming different hydrocarbon fuels including methane, methanol, ethanol, propane, toluene, glycerol,
diesel and naphtha is discussed in this review. Reaction mechanisms with associated rates constants for reforming process of these fuels
in different catalysts are presented as well. The main findings are described in the following:
• The hydrogen produced from fuel reforming can be separated from integrating Pd-based membranes.
• Hydrogen permeation through Pd-membranes depends highly on the difference in hydrogen pressure.

28
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

• can be improved by employing different pressure configuration, introducing sweep flow on the permeate side of the membrane,
reducing retentate side flow rate, and increasing the temperature. The geometry of the membrane also influences the hydrogen
permeation.
• The most studied geometry in membrane modeling is the configuration with two concentrate cylinders. Several other geometries
are found in the literature which can improve the permeation process.
• For instance, the conical membrane shape can allow higher hydrogen permeation compared to the regulate cylindrical shape.
Furthermore, helical geometry of the membrane also demonstrated better performance than straight membrane tube in terms of
hydrogen flux through the membrane.
• The expressions for hydrogen permeability and flux through the membrane are also summarized in the study.

Author statement
The paper was revised as per the reviewers’ comments. The modifications are shown in the manuscript in “Green” color. A
summary of the modification is sown in the “response” file:

Declaration of competing interest


I declare the paper is not published or submitted to any other journals.

Data availability

No data was used for the research described in the article.

Acknowledgment
The authors would like to acknowledge the support provided by Saudi Aramco and King Fahd University of Petroleum and Minerals
(KFUPM) in carrying out this review and analysis through Project No. ME2489 and Project No. DUP22102. The support provided by
the King Abdullah City for Atomic and Renewable Energy (K.A. CARE) and Projects DF201009 and INHE2102 to two of the authors is
also acknowledged.

Nomenclature

GHG Greenhouse gas


CFD Computational fluid dynamics
NG Natural gas
DFT Density functional theory
R&D Reynolds number and similar abbreviations do not use italics
PSA Pressure swing adsorption
JH2 Hydrogen permeation flux

References
[1] M.A. Nemitallah, H.A. Haque, M. Hussain, A. Abdelhafez, M.A. Habib, Stratified and hydrogen combustion techniques for higher turndown and lower
emissions in gas turbines, J. Energy Resour. Technol. (2021) 1–42, https://doi.org/10.1115/1.4052541.
[2] A. Basile, A. Iulianelli, J. Tong, Single-stage hydrogen production and separation from fossil fuels using micro- and macromembrane reactors, Compend.
Hydrog. Energy (2015) 445–468, https://doi.org/10.1016/B978-1-78242-361-4.00015-7. Elsevier.
[3] V. Twigg M, Catalyst Handbook, second ed., Routledge, 1989 https://doi.org/10.1201/9781315138862.
[4] Rinkesh, What is hydrogen energy? Conserv futur energy n.d. https://www.conserve-energy-future.com/advantages_disadvantages_hydrogenenergy.php.
[5] V. Subramani, A. Basile, T.N. Veziroğlu, Compendium of Hydrogen Energy, Elsevier, 2015, https://doi.org/10.1016/C2014-0-02671-8.
[6] Y. Naimi, A. Antar, Hydrogen Generation by Water Electrolysis, 2018, https://doi.org/10.5772/intechopen.76814. Adv. Hydrog. Gener. Technol., InTech.
[7] G. Franchi, M. Capocelli, M. De Falco, V. Piemonte, D. Barba, Hydrogen production via steam reforming: a critical analysis of MR and RMM technologies,
Membranes 10 (2020), https://doi.org/10.3390/membranes10010010.
[8] A. Kokka, A. Katsoni, I.V. Yentekakis, P. Panagiotopoulou, Hydrogen production via steam reforming of propane over supported metal catalysts, Int. J.
Hydrogen Energy 45 (2020) 14849–14866, https://doi.org/10.1016/j.ijhydene.2020.03.194.
[9] P.P. Silva, R.A. Ferreira, J.F. Nunes, J.A. Sousa, L.L. Romanielo, F.B. Noronha, et al., Production of hydrogen from the steam and oxidative reforming of LPG:
thermodynamic and experimental study, Braz. J. Chem. Eng. 32 (2015) 647–662, https://doi.org/10.1590/0104-6632.20150323s00003441.
[10] M. Ozturk, I. Dincer, A comprehensive review on power-to-gas with hydrogen options for cleaner applications, Int. J. Hydrogen Energy 46 (2021)
31511–31522, https://doi.org/10.1016/j.ijhydene.2021.07.066.
[11] S. Liguori, K. Kian, N. Buggy, B.H. Anzelmo, J. Wilcox, Opportunities and challenges of low-carbon hydrogen via metallic membranes, Prog. Energy Combust.
Sci. 80 (2020), 100851, https://doi.org/10.1016/j.pecs.2020.100851.
[12] F. Alrashed, U. Zahid, Comparative analysis of conventional steam methane reforming and PdAu membrane reactor for the hydrogen production, Comput.
Chem. Eng. 154 (2021), 107497, https://doi.org/10.1016/j.compchemeng.2021.107497.
[13] V. Martin-Gil, M.Z. Ahmad, R. Castro-Muñoz, V. Fila, Economic framework of membrane technologies for natural gas applications, Separ. Purif. Rev. 48 (2019)
298–324, https://doi.org/10.1080/15422119.2018.1532911.
[14] S. Adhikari, S. Fernando, Hydrogen membrane separation techniques, Ind. Eng. Chem. Res. 45 (2006) 875–881, https://doi.org/10.1021/ie050644l.

29
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

[15] M.A. Habib, A. Harale, S. Paglieri, F.S. Alrashed, A. Al-Sayoud, M.V. Rao, et al., Palladium-alloy membrane reactors for fuel reforming and hydrogen
production: a review, Energy Fuel. 35 (2021) 5558–5593, https://doi.org/10.1021/acs.energyfuels.0c04352.
[16] I. Hijazi, Y. Zhang, R. Fuller, A simple palladium hydride embedded atom method potential for hydrogen energy applications, J. Energy Resour. Technol. 141
(2019), https://doi.org/10.1115/1.4042405.
[17] B. Sharma, L. Myung, Pd-based ternary alloys used for gas sensing applications: a review, Int. J. Hydrogen Energy 44 (2019) 30499–30510, https://doi.org/
10.1016/j.ijhydene.2019.09.170.
[18] T.Y. Amiri, K. Ghasemzageh, A. Iulianelli, Membrane reactors for sustainable hydrogen production through steam reforming of hydrocarbons: a review, Chem
Eng Process - Process Intensif 157 (2020), 108148, https://doi.org/10.1016/j.cep.2020.108148.
[19] M.S. Salehi, M. Askarishahi, H.R. Godini, R. Schomäcker, G. Wozny, CFD simulation of oxidative coupling of methane in fluidized-bed reactors: a detailed
analysis of flow-reaction characteristics and operating conditions, Ind. Eng. Chem. Res. 55 (2016) 1149–1163, https://doi.org/10.1021/acs.iecr.5b02433.
[20] K. Ghasemzadeh, M. Khosravi, S.M. Sadati Tilebon, A. Aghaeinejad-Meybodi, A. Basile, Theoretical evaluation of Pd–Ag membrane reactor performance
during biomass steam gasification for hydrogen production using CFD method, Int. J. Hydrogen Energy 43 (2018) 11719–11730, https://doi.org/10.1016/j.
ijhydene.2018.04.221.
[21] S. Lee, H. Lim, The effect of changing the number of membranes in methane carbon dioxide reforming: a CFD study, J. Ind. Eng. Chem. 87 (2020) 110–119,
https://doi.org/10.1016/j.jiec.2020.03.020.
[22] Y.S. Sanusi, E.M.A. Mokheimer, A numerical investigation of hydrogen production in an integrated membrane reformer-combustor, Energy Proc. 142 (2017)
1077–1082, https://doi.org/10.1016/j.egypro.2017.12.359.
[23] T. Nakajima, T. Kume, Y. Ikeda, M. Shiraki, H. Kurokawa, T. Iseki, et al., Effect of concentration polarization on hydrogen production performance of ceramic-
supported Pd membrane module, Int. J. Hydrogen Energy 40 (2015) 11451–11456, https://doi.org/10.1016/j.ijhydene.2015.03.088.
[24] J.R. Fernandez, J.C. Abanades, R. Murillo, Modeling of sorption enhanced steam methane reforming in an adiabatic fixed bed reactor, Chem. Eng. Sci. 84
(2012) 1–11, https://doi.org/10.1016/j.ces.2012.07.039.
[25] A. Konieczny, K. Mondal, T. Wiltowski, P. Dydo, Catalyst development for thermocatalytic decomposition of methane to hydrogen, Int. J. Hydrogen Energy 33
(2008) 264–272, https://doi.org/10.1016/j.ijhydene.2007.07.054.
[26] E. Kosa, C. Ezgi, Numerical modeling of hydrogen-rich gas production from gasoline autothermal reforming in a plug flow reactor for electric vehicles,
J. Energy Resour. Technol. 143 (2021), https://doi.org/10.1115/1.4052247.
[27] R. Scenna, A.K. Gupta, The influence of the distributed reaction regime on fuel reforming conditions, J. Energy Resour. Technol. 140 (2018) 1–7, https://doi.
org/10.1115/1.4040404.
[28] Y. Yan, H. Yan, L. Li, L. Zhang, Z. Yang, Thermodynamic analysis on reaction characteristics of the coupling steam, CO2 and O2 reforming of methane,
J. Energy Resour. Technol. 140 (2018), https://doi.org/10.1115/1.4040074.
[29] L. Chen, Z. Qi, S. Zhang, J. Su, G.A. Somorjai, Catalytic hydrogen production from methane: a review on recent progress and prospect, Catalysts 10 (2020),
https://doi.org/10.3390/catal10080858.
[30] J.M. Lavoie, Review on dry reforming of methane, a potentially more environmentally-friendly approach to the increasing natural gas exploitation, Front.
Chem. 2 (2014), https://doi.org/10.3389/fchem.2014.00081.
[31] W. Chen, CO2 conversion for syngas production in methane catalytic partial oxidation, Biochem. Pharmacol. 5 (2014) 1–9, https://doi.org/10.1016/j.
jcou.2013.11.001.
[32] R.D. Alli, P.A.L. de Souza, M. Mohamedali, L.D. Virla, N. Mahinpey, Tri-reforming of methane for syngas production using Ni catalysts: current status and
future outlook, Catal. Today (2022), https://doi.org/10.1016/j.cattod.2022.02.006.
[33] M. Luneau, E. Gianotti, N. Guilhaume, E. Landrivon, F.C. Meunier, C. Mirodatos, et al., Experiments and modeling of methane autothermal reforming over
structured Ni–Rh-based Si-SiC foam catalysts, Ind. Eng. Chem. Res. 56 (2017) 13165–13174, https://doi.org/10.1021/acs.iecr.7b01559.
[34] L. Roses, G. Manzolini, S. Campanari, CFD simulation of Pd-based membrane reformer when thermally coupled within a fuel cell micro-CHP system, Int. J.
Hydrogen Energy 35 (2010) 12668–12679, https://doi.org/10.1016/j.ijhydene.2010.07.080.
[35] L. Maier, B. Schädel, K.H. Delgado, S. Tischer, O. Deutschmann, Steam reforming of methane over nickel: development of a multi-step surface reaction
mechanism, Top. Catal. 54 (2011) 845–858, https://doi.org/10.1007/s11244-011-9702-1.
[36] J. Xu, G.F. Froment, Methane steam reforming, methanation and water-gas shift: I. Intrinsic kinetics, AIChE J. 35 (1989) 88–96, https://doi.org/10.1002/
aic.690350109.
[37] M. Farniaei, M. Abbasi, H. Rahnama, M.R. Rahimpour, A. Shariati, Syngas production in a novel methane dry reformer by utilizing of tri-reforming process for
energy supplying: modeling and simulation, J. Nat. Gas Sci. Eng. 20 (2014) 132–146, https://doi.org/10.1016/j.jngse.2014.06.010.
[38] K.S. Patel, A.K. Sunol, Modeling and simulation of methane steam reforming in a thermally coupled membrane reactor, Int. J. Hydrogen Energy 32 (2007)
2344–2358, https://doi.org/10.1016/j.ijhydene.2007.03.004.
[39] E.M.A. Mokheimer, M. Ibrar Hussain, S. Ahmed, M.A. Habib, A.A. Al-Qutub, On the modeling of steam methane reforming, J. Energy Resour. Technol. 137
(2015), https://doi.org/10.1115/1.4027962.
[40] M. Sheintuch, Pure hydrogen production in a membrane reformer: demonstration, macro-scale and atomic scale modeling, Chem. Eng. J. 278 (2015) 363–373,
https://doi.org/10.1016/j.cej.2014.11.100.
[41] C. Herce, C. Cortés, S. Stendardo, Numerical simulation of a bubbling fluidized bed reactor for sorption-enhanced steam methane reforming under industrially
relevant conditions: effect of sorbent (dolomite and CaO-Ca12Al14O33) and operational parameters, Fuel Process. Technol. 186 (2019) 137–148, https://doi.
org/10.1016/j.fuproc.2019.01.003.
[42] S.A.M. Said, D.S.A. Simakov, M. Waseeuddin, Y. Román-Leshkov, Solar molten salt heated membrane reformer for natural gas upgrading and hydrogen
generation: a CFD model, Sol. Energy 124 (2016) 163–176, https://doi.org/10.1016/j.solener.2015.11.038.
[43] T. Jiwanuruk, S. Putivisutisak, P. Ponpesh, P. Bumroongsakulsawat, T. Tagawa, H. Yamada, et al., Effect of flow arrangement on micro membrane reforming
for H2 production from methane, Chem. Eng. J. 293 (2016) 319–326, https://doi.org/10.1016/j.cej.2016.02.075.
[44] S.A.M. Said, D.S.A. Simakov, E.M.A. Mokheimer, M.A. Habib, S. Ahmed, M. Waseeuddin, et al., Computational fluid dynamics study of hydrogen generation by
low temperature methane reforming in a membrane reactor, Int. J. Hydrogen Energy 40 (2015) 3158–3169, https://doi.org/10.1016/j.ijhydene.2015.01.024.
[45] M.L. Rodríguez, M.N. Pedernera, D.O. Borio, Two dimensional modeling of a membrane reactor for ATR of methane, Catal. Today 193 (2012) 137–144,
https://doi.org/10.1016/j.cattod.2012.04.010.
[46] S.H. Chan, D.L. Hoang, O.L. Ding, Transient performance of an autothermal reformer - a 2-D modeling approach, Int. J. Heat Mass Tran. 48 (2005) 4205–4214,
https://doi.org/10.1016/j.ijheatmasstransfer.2005.02.042.
[47] Y. Wang, F. Yoshiba, M. Kawase, T. Watanabe, Performance and effective kinetic models of methane steam reforming over Ni/YSZ anode of planar SOFC, Int.
J. Hydrogen Energy 34 (2009) 3885–3893, https://doi.org/10.1016/j.ijhydene.2009.02.073.
[48] T. Numaguchi, K. Kikuchi, Intrinsic kinetics and design simulation in a complex reaction network: steam-methane reforming, Chem. Eng. Sci. 43 (1988)
2295–2301, https://doi.org/10.1016/0009-2509(88)87118-5.
[49] A. Shigarov, V. Kirillov, I. Landgraf, Computational study of Pd-membrane CH4 steam reformer with fixed catalyst bed: searching for a way to increase
membrane efficiency, Int. J. Hydrogen Energy 39 (2014) 20072–20093, https://doi.org/10.1016/j.ijhydene.2014.10.018.
[50] M. Mbodji, J.M. Commenge, L. Falk, D. Di Marco, F. Rossignol, L. Prost, et al., Steam methane reforming reaction process intensification by using a
millistructured reactor: experimental setup and model validation for global kinetic reaction rate estimation, Chem. Eng. J. 207–208 (2012) 871–884, https://
doi.org/10.1016/j.cej.2012.07.117.
[51] H. Lee, A. Kim, B. Lee, H. Lim, Comparative numerical analysis for an efficient hydrogen production via a steam methane reforming with a packed-bed reactor,
a membrane reactor, and a sorption-enhanced membrane reactor, Energy Convers. Manag. 213 (2020), 112839, https://doi.org/10.1016/j.
enconman.2020.112839.

30
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

[52] D.L. Hoang, S.H. Chan, O.L. Ding, Kinetic and modelling study of methane steam reforming over sulfide nickel catalyst on a gamma alumina support, Chem.
Eng. J. 112 (2005) 1–11, https://doi.org/10.1016/j.cej.2005.06.004.
[53] M. Hassan Amin, A mini-review on CO2 reforming of methane, Prog Petrochemical Sci 2 (2018), https://doi.org/10.31031/PPS.2018.02.000532.
[54] B. Lee, S.W. Yun, S. Kim, J. Heo, Y.T. Kim, S. Lee, et al., CO 2 reforming of methane for H 2 production in a membrane reactor as CO 2 utilization:
computational fluid dynamics studies with a reactor geometry, Int. J. Hydrogen Energy 44 (2019) 2298–2311, https://doi.org/10.1016/j.
ijhydene.2018.09.184.
[55] J.T. Richardson, S.A. Paripatyadar, Carbon dioxide reforming of methane with supported rhodium, Appl. Catal. 61 (1990) 293–309, https://doi.org/10.1016/
S0166-9834(00)82152-1.
[56] V. Barrio, G. Schaub, M. Rohde, S. Rabe, F. Vogel, J. Cambra, et al., Reactor modeling to simulate catalytic partial oxidation and steam reforming of methane.
Comparison of temperature profiles and strategies for hot spot minimization, Int. J. Hydrogen Energy 32 (2007) 1421–1428, https://doi.org/10.1016/j.
ijhydene.2006.10.022.
[57] D.D. Nogare, Modeling Catalytic Methane Partial Oxidation, Università degli Studi di Padova, 2008.
[58] D.L. Hoang, S.H. Chan, O.L. Ding, Kinetic modelling of partial oxidation of methane in an oxygen permeable membrane reactor, Chem. Eng. Res. Des. 83
(2005) 177–186, https://doi.org/10.1205/cherd.04151.
[59] F.A.N. Fernandes, C.P. Souza, J.F. Sousa, Modeling of partial oxidation of methane in a membrane reactor, Eng. Térmica 5 (2006) 40–45.
[60] K.V. Dobrego, N.N. Gnesdilov, S.H. Lee, H.K. Choi, Overall chemical kinetics model for partial oxidation of methane in inert porous media, Chem. Eng. J. J 144
(2008) 79–87, https://doi.org/10.1016/j.cej.2008.05.027.
[61] W.H. Chen, CO2 conversion for syngas production in methane catalytic partial oxidation, J. CO2 Util. 5 (2014) 1–9, https://doi.org/10.1016/j.
jcou.2013.11.001.
[62] S. Pruksawan, B. Kitiyanan, R.M. Ziff, Partial oxidation of methane on a nickel catalyst, Kinetic Monte-Carlo Simulation Study 1–15 (2011).
[63] K. Ghasemzadeh, J.N. Harasi, T.Y. Amiri, A. Basile, A. Iulianelli, Methanol steam reforming for hydrogen generation: a comparative modeling study between
silica and Pd-based membrane reactors by CFD method, Fuel Process. Technol. 199 (2020), 106273, https://doi.org/10.1016/j.fuproc.2019.106273.
[64] B.A. Peppley, J.C. Amphlett, L.M. Kearns, R.F. Mann, Methanol-steam reforming on Cu/Zno/Al2O3. Part 1: the reaction network, Appl. Catal. Gen. 179 (1999)
21–29, https://doi.org/10.1016/S0926-860X(98)00298-1.
[65] B.A. Peppley, J.C. Amphlett, L.M. Kearns, R.F. Mann, Methanol-steam reforming on Cu/ZnO/Al2O3 catalysts. Part 2. A comprehensive kinetic model, Appl.
Catal. Gen. 179 (1999) 31–49, https://doi.org/10.1016/S0926-860X(98)00299-3.
[66] K. Ghasemzadeh, E. Andalib, A. Basile, Evaluation of dense Pd-Ag membrane reactor performance during methanol steam reforming in comparison with
autothermal reforming using CFD analysis, Int. J. Hydrogen Energy 41 (2016) 8745–8754, https://doi.org/10.1016/j.ijhydene.2015.11.139.
[67] I. Uriz, G. Arzamendi, P.M. Diéguez, F.J. Echave, O. Sanz, M. Montes, et al., CFD analysis of the effects of the flow distribution and heat losses on the steam
reforming of methanol in catalytic (Pd/ZnO) microreactors, Chem. Eng. J. 238 (2014) 37–44, https://doi.org/10.1016/j.cej.2013.05.097.
[68] M. Heidarzadeh, M. Taghizadeh, Methanol steam reforming in a spiral-shaped microchannel reactor over Cu/ZnO/Al2O3Catalyst: a computational fluid
dynamics simulation study, Int. J. Chem. React. Eng. 15 (2017), https://doi.org/10.1515/ijcre-2016-0205.
[69] S. Bepari, D. Kuila, Steam reforming of methanol, ethanol and glycerol over nickel-based catalysts-A review, Int. J. Hydrogen Energy 45 (2020) 18090–18113,
https://doi.org/10.1016/j.ijhydene.2019.08.003.
[70] R. Ma, B. Castro-Dominguez, A.G. Dixon, Y.H. Ma, CFD study of heat and mass transfer in ethanol steam reforming in a catalytic membrane reactor, Int. J.
Hydrogen Energy 43 (2018) 7662–7674, https://doi.org/10.1016/j.ijhydene.2017.08.173.
[71] X. Yang, S. Wang, B. Li, Y. He, H. Liu, Performance of ethanol steam reforming in a membrane-assisted packed bed reactor using multiscale modelling, Fuel
274 (2020), 117829, https://doi.org/10.1016/j.fuel.2020.117829.
[72] W.H. Chen, C.V. Lu, K.Q. Tran, Y.L. Lin, S.R. Naqvi, A new design of catalytic tube reactor for hydrogen production from ethanol steam reforming, Fuel 281
(2020), 118746, https://doi.org/10.1016/j.fuel.2020.118746.
[73] A. Davidy, CFD simulation of ethanol steam reforming system for hydrogen production, ChemEngineering 2 (2018) 34, https://doi.org/10.3390/
chemengineering2030034.
[74] R. Ma, B. Castro-Dominguez, I.P. Mardilovich, A.G. Dixon, Y.H. Ma, Experimental and simulation studies of the production of renewable hydrogen through
ethanol steam reforming in a large-scale catalytic membrane reactor, Chem. Eng. J. 303 (2016) 302–313, https://doi.org/10.1016/j.cej.2016.06.021.
[75] M. Patel, T.K. Jindal, K.K. Pant, Kinetic study of steam reforming of ethanol on Ni-based ceria-zirconia catalyst, Ind. Eng. Chem. Res. 52 (2013) 15763–15771,
https://doi.org/10.1021/ie401570s.
[76] A. Bshish, Z. Yaakob, A. Ebshish, F.H. Alhasan, Hydrogen production via ethanol steam reforming over Ni/Al2O3 catalysts: effect of Ni loading, J. Energy
Resour. Technol. 136 (2014), https://doi.org/10.1115/1.4024915.
[77] J. Sun, X.P. Qiu, F. Wu, W.T. Zhu, H2 from steam reforming of ethanol at low temperature over Ni/Y2O3 and Ni/La2O3 catalysts for fuel-cell application, Int.
J. Hydrogen Energy 30 (2005) 437–445, https://doi.org/10.1016/j.ijhydene.2004.11.005.
[78] G.M. Zanin, F.F. Moraes, Modeling hydrogen production from ethanol steam reforming in a membrane microreactor, Braz. J. Chem. Eng. 30 (2009) 729–739.
[79] M. De-Souza, G.M. Zanin, F.F. Moraes, Parametric study of hydrogen production from ethanol steam reforming in a membrane microreactor, Braz. J. Chem.
Eng. 30 (2013) 355–367.
[80] A. Cifuentes, R. Torres, J. Llorca, Modelling of the ethanol steam reforming over Rh-Pd/CeO2 catalytic wall reactors, Int. J. Hydrogen Energy 45 (2020)
26265–26273, https://doi.org/10.1016/j.ijhydene.2019.11.034.
[81] N.R. Peela, D. Kunzru, Steam reforming of ethanol in a microchannel reactor: kinetic study and reactor simulation, Ind. Eng. Chem. Res. 50 (2011)
12881–12894, https://doi.org/10.1021/ie200084b.
[82] M. Saidi, A. Jahangiri, Theoretical study of hydrogen production by ethanol steam reforming: technical evaluation and performance analysis of catalytic
membrane reactor, Int. J. Hydrogen Energy 43 (2018) 15306–15320, https://doi.org/10.1016/j.ijhydene.2018.06.110.
[83] Y.M. Bruschi, E. López, N.S. Schbib, M.N. Pedernera, D.O. Borio, Theoretical study of the ethanol steam reforming in a parallel channel reactor, Int. J.
Hydrogen Energy 37 (2012) 14887–14894, https://doi.org/10.1016/j.ijhydene.2012.01.175.
[84] E.M. Izurieta, D.O. Borio, M.N. Pedernera, E. López, Parallel plates reactor simulation: ethanol steam reforming thermally coupled with ethanol combustion,
Int. J. Hydrogen Energy 42 (2017) 18794–18804, https://doi.org/10.1016/j.ijhydene.2017.06.134.
[85] F. Gallucci, M. De Falco, S. Tosti, L. Marrelli, A. Basile, Ethanol steam reforming in a dense Pd-Ag membrane reactor: a modelling work. Comparison with the
traditional system, Int. J. Hydrogen Energy 33 (2008) 644–651, https://doi.org/10.1016/j.ijhydene.2007.10.039.
[86] M.A. Rakib, J.R. Grace, C.J. Lim, S.S.E.H. Elnashaie, Modeling of a fluidized bed membrane reactor for hydrogen production by steam reforming of
hydrocarbons, Ind. Eng. Chem. Res. 50 (2011) 3110–3129, https://doi.org/10.1021/ie100954a.
[87] M.A. Rakib, J.R. Grace, C.J. Lim, S.S.E.H. Elnashaie, B. Ghiasi, Steam reforming of propane in a fluidized bed membrane reactor for hydrogen production,
Renew. Energy 35 (2010) 6276–6290, https://doi.org/10.1016/j.ijhydene.2010.03.136.
[88] L. Ma, Hydrogen Production from Steam Reforming of Light Hydrocarbons in an Autothermic System, University of New South Wales, 1995.
[89] S. Odiba, M. Olea, S. Hodgson, A. Adgar, Computational Fluid Dynamics for Microreactors Used in Catalytic Oxidation of Propane, Proc. 2013 COMSOL Conf.,
Boston, MA, USA, 2013.
[90] D. Zhang, Z. Shao, J.P. Xin, A.J. Li, Modeling of high-temperature reforming reactor based on interactive thermal control with a tail gas combustor, Adv.
Manuf. 4 (2016) 47–54, https://doi.org/10.1007/s40436-015-0118-1.
[91] K. Ghasemzadeh, M.M. Alinejad, M. Ghahremani, R. Zeynali, A. Pourgholi, Theoritical study of palladium membrane reactor performance during propane
dehydrogenation using CFD method, Indones J Chem 17 (2017) 113–118, https://doi.org/10.22146/ijc.23625.
[92] H. Karadeniz, C. Karakaya, S. Tischer, O. Deutschmann, Numerical simulation of methane and propane reforming over a porous Rh/Al2O3 catalyst in
stagnation-flows: impact of internal and external mass transfer limitations on species profiles, Catalysts 10 (2020) 1–25, https://doi.org/10.3390/
catal10080915.

31
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

[93] P. Barnoon, D. Toghraie, B. Mehmandoust, M.A. Fazilati, S.A. Eftekhari, Comprehensive study on hydrogen production via propane steam reforming inside a
reactor, Energy Rep. 7 (2021) 929–941, https://doi.org/10.1016/j.egyr.2021.02.001.
[94] E.B. dos Anjos, A.M. da Silva Filho, J.D. da Silva, Numerical simulation of the steam reforming of toluene in a fixed-bed catalytic reactor to produce hydrogen,
J Brazilian Soc Mech Sci Eng 42 (2020) 114, https://doi.org/10.1007/s40430-020-2195-8.
[95] C.C.B. Oliveira, J.D. Da Silva, Mathematical modelling of the steam reforming of toluene for fuel gas production in a fixed bed catalytic reactor, Chem Eng
Trans 35 (2013) 307–312, https://doi.org/10.3303/CET1335051.
[96] D. Swierczynski, C. Courson, A. Kiennemann, Study of steam reforming of toluene used as model compound of tar produced by biomass gasification, Chem Eng
Process Process Intensif 47 (2008) 508–513, https://doi.org/10.1016/j.cep.2007.01.012.
[97] B. Zhao, X. Zhang, L. Chen, R. Qu, G. Meng, X. Yi, et al., Steam reforming of toluene as model compound of biomass pyrolysis tar for hydrogen, Biomass
Bioenergy 34 (2010) 140–144, https://doi.org/10.1016/j.biombioe.2009.10.011.
[98] U. Oemar, A. Ming Li, K. Hidajat, S. Kawi, Mechanism and kinetic modeling for steam reforming of toluene on La0.8Sr0.2Ni0.8Fe0.2O3 catalyst, AIChE J. 60
(2014) 4190–4198, https://doi.org/10.1002/aic.14573.
[99] C. Aline, H. José, R. Andrade, F. Alves, R. Sequinel, V. Rossato, et al., Overview of glycerol reforming for hydrogen production, Renew. Sustain. Energy Rev. 58
(2016) 259–266, https://doi.org/10.1016/j.rser.2015.12.279.
[100] M. Maximini, T. Huck, R. Wruck, K. Lucka, M. Maximini, T. Huck, et al., Investigations on steam reforming catalysts for diesel fuel investigations on steam
reforming catalysts for diesel fuel, in: D. Stolten, T. Grube (Eds.), 18th World Hydrog. Energy Conf, 2010. Essen, Germany.
[101] G. Zoppi, G. Pipitone, R. Pirone, S. Bensaid, Aqueous phase reforming process for the valorization of wastewater streams: application to different industrial
scenarios, Catal. Today 387 (2022) 224–236, https://doi.org/10.1016/j.cattod.2021.06.002.
[102] E. Markočič, B. Kramberger, J.G. van Bennekom, H. Jan Heeres, J. Vos, Z. Knez, Glycerol reforming in supercritical water; a short review, Renew. Sustain.
Energy Rev. 23 (2013) 40–48, https://doi.org/10.1016/j.rser.2013.02.046.
[103] C.K. Cheng, S.Y. Foo, A.A. Adesina, H2-rich synthesis gas production over Co/Al2O 3 catalyst via glycerol steam reforming, Catal. Commun. 12 (2010)
292–298, https://doi.org/10.1016/j.catcom.2010.09.018.
[104] R. Sundari, P.D. Vaidya, Reaction kinetics of glycerol steam reforming using a Ru/Al 2O 3 catalyst, Energy Fuel. 26 (2012) 4195–4204, https://doi.org/
10.1021/ef300658n.
[105] C.K. Cheng, S.Y. Foo, A.A. Adesina, Glycerol steam reforming over bimetallic Co-Ni/Al2O3, Ind. Eng. Chem. Res. 49 (2010) 10804–10817, https://doi.org/
10.1021/ie100462t.
[106] C.K. Cheng, S.Y. Foo, A.A. Adesina, Steam reforming of glycerol over Ni/Al2O3 catalyst, Catal. Today 178 (2011) 25–33, https://doi.org/10.1016/j.
cattod.2011.07.011.
[107] C.D. Dave, K.K. Pant, Renewable hydrogen generation by steam reforming of glycerol over zirconia promoted ceria supported catalyst, Renew. Energy 36
(2011) 3195–3202, https://doi.org/10.1016/j.renene.2011.03.013.
[108] C. Wang, B. Dou, H. Chen, Y. Song, Y. Xu, X. Du, et al., Hydrogen production from steam reforming of glycerol by Ni-Mg-Al based catalysts in a fixed-bed
reactor, Chem. Eng. J. 220 (2013) 133–142, https://doi.org/10.1016/j.cej.2013.01.050.
[109] K.K. Pant, R. Jain, S. Jain, Renewable hydrogen production by steam reforming of glycerol over Ni/CeO2 catalyst prepared by precipitation deposition method,
Kor. J. Chem. Eng. 28 (2011) 1859–1866, https://doi.org/10.1007/s11814-011-0059-8.
[110] K. Ghasemzadeh, M. Ghahremani, T.Y. Amiri, A. Basile, Performance evaluation of Pd–Ag membrane reactor in glycerol steam reforming process: development
of the CFD model, Int. J. Hydrogen Energy 44 (2019) 1000–1009, https://doi.org/10.1016/j.ijhydene.2018.11.086.
[111] M. Saidi, P. Moradi, Conversion of biodiesel synthesis waste to hydrogen in membrane reactor: theoretical study of glycerol steam reforming, Int. J. Hydrogen
Energy 45 (2020) 8715–8726, https://doi.org/10.1016/j.ijhydene.2020.01.064.
[112] M.N. Shahirah, S. Abdullah, J. Gimbun, Y.H. Ng, C.K. Cheng, A study on the kinetics of syngas production from glycerol over alumina-supported
samarium–nickel catalyst, Int. J. Hydrogen Energy 41 (2016) 10568–10577, https://doi.org/10.1016/j.ijhydene.2016.04.193.
[113] X. Yang, S. Wang, S. Liu, Y. He, CFD analysis on sorption-enhanced glycerol reforming in a membrane-assisted fluidized bed reactor, Int. J. Hydrogen Energy
44 (2019) 21424–21433, https://doi.org/10.1016/j.ijhydene.2019.06.088.
[114] M.R. Rahimpour, M. Jafari, D. Iranshahi, Progress in catalytic naphtha reforming process: a review, Appl. Energy 109 (2013) 79–93, https://doi.org/10.1016/
j.apenergy.2013.03.080.
[115] M.R. Rahimpour, Enhancement of hydrogen production in a novel fluidized-bed membrane reactor for naphtha reforming, Int. J. Hydrogen Energy 34 (2009)
2235–2251, https://doi.org/10.1016/j.ijhydene.2008.10.098.
[116] A. Khosravanipour Mostafazadeh, M.R. Rahimpour, A membrane catalytic bed concept for naphtha reforming in the presence of catalyst deactivation, Chem
Eng Process Process Intensif 48 (2009) 683–694, https://doi.org/10.1016/j.cep.2008.08.003.
[117] D. Iranshahi, M. Karimi, S. Amiri, M. Jafari, R. Rafiei, M.R. Rahimpour, Modeling of naphtha reforming unit applying detailed description of kinetic in
continuous catalytic regeneration process, Chem. Eng. Res. Des. 92 (2014) 1704–1727, https://doi.org/10.1016/j.cherd.2013.12.012.
[118] R. Smith, Kinetic analysis of naphtha reforming with platinum catalyst, Chem. Eng. Prog. 55 (1959) 75–80.
[119] J.H. Jenkins, T.W. Stephens, Kinetics of cat reforming, Hydrocarb. Process. 60 (1980) 163–167.
[120] G.F. Froment, The kinetics of complex catalytic reactions, Chem. Eng. Sci. 42 (1987) 1073–1087, https://doi.org/10.1016/0009-2509(87)80057-X.
[121] A.K. Saxena, G. Das, H.B. Goyal, V.K. Kapoor, Simulation and optimisation package for semi-regenerative catalytic reformer, Hydrocarb Technol (1994) 71–83.
[122] U. Taskar, J.B. Riggs, Modeling and optimization of a semiregenerative catalytic naphtha reformer, AIChE J. 43 (1997) 740–753, https://doi.org/10.1002/
aic.690430319.
[123] G.P. Vathi, K.K. Chaudhuri, Modelling and simulation of commercial catalytic naphtha reformers, Can. J. Chem. Eng. 75 (1997) 930–937, https://doi.org/
10.1002/cjce.5450750513.
[124] J. Ancheyta-Juárez, E. Villafuerte-Macías, Kinetic modeling of naphtha catalytic reforming reactions, Energy Fuel. 14 (2000) 1032–1037, https://doi.org/
10.1021/ef0000274.
[125] H.U. Yongyou, X. Weihua, S. Hongye, C. Jian, A dynamic model for naphtha catalytic reformers, Proc. 2004 IEEE Int. Conf. Control Appl. (2004) 159–164,
https://doi.org/10.1109/CCA.2004.1387204. IEEE; n.d.
[126] S. Hu, X.X. Zhu, Molecular modeling and optimization for catalytic reforming, Chem. Eng. Commun. 191 (2004) 500–512, https://doi.org/10.1080/
00986440390255933.
[127] W. Hou, H. Su, Y. Hu, J. Chu, Lumped kinetics model and its on-line application to commercial catalytic naphtha reforming process, Huagong Xuebao/Journal
Chem Ind Eng 57 (2006) 1605–1611.
[128] W. Hou, H. Su, Y. Hu, J. Chu, Modeling, simulation and optimization of a whole industrial catalytic naphtha reforming process on aspen plus platform, Chin. J.
Chem. Eng. 14 (2006) 584–591, https://doi.org/10.1016/S1004-9541(06)60119-5.
[129] H.M. Arani, M. Shirvani, K. Safdarian, E. Dorostkar, Lumping procedure for a kinetic model of catalytic naphtha reforming, Braz. J. Chem. Eng. 26 (2009)
723–732, https://doi.org/10.1590/S0104-66322009000400011.
[130] Z. Hongjun, S. Mingliang, W. Huixin, L. Zeji, J. Hongbo, Modeling and simulation of moving bed reactor for catalytic naphtha reforming, Petrol. Sci. Technol.
28 (2010) 667–676, https://doi.org/10.1080/10916460902804598.
[131] L. Wang, Q. Zhang, C. Liang, A 38-lumped kinetic model for reforming reaction and its application in continuous catalytic reforming, Chin. J. Chem. Eng. 63
(2012) 1076–1082, https://doi.org/10.3969/j.issn.0438-1157.2012.04.013.
[132] J.M. Leimert, J. Karl, Nickel membranes for in-situ hydrogen separation in high-temperature fluidized bed gasification processes, Int. J. Hydrogen Energy 41
(2016) 9355–9366, https://doi.org/10.1016/j.ijhydene.2016.04.073.
[133] N.A. Al-Mufachi, N.V. Rees, R. Steinberger-Wilkens, Hydrogen selective membranes: a review of palladium-based dense metal membranes, Renew. Sustain.
Energy Rev. 47 (2015) 540–551, https://doi.org/10.1016/j.rser.2015.03.026.

32
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

[134] A. Santucci, S. Tosti, A. Basile, Alternatives to Palladium in Membranes for Hydrogen Separation: Nickel, Niobium and Vanadium Alloys, Ceramic Supports for
Metal Alloys and Porous Glass Membranes, 2013, pp. 183–217, https://doi.org/10.1533/9780857097330.1.183. Handb. Membr. React., Elsevier.
[135] M.R. Rahimpour, F. Samimi, A. Babapoor, T. Tohidian, S. Mohebi, Palladium membranes applications in reaction systems for hydrogen separation and
purification: a review, Chem Eng Process Process Intensif 121 (2017) 24–49, https://doi.org/10.1016/j.cep.2017.07.021.
[136] R. Ben-Mansour, A. Abuelyamen, M.A. Habib, CFD modeling of hydrogen separation through Pd-based membrane, Int. J. Hydrogen Energy 45 (2020)
23006–23019, https://doi.org/10.1016/j.ijhydene.2020.06.141.
[137] Ö. Yücel, M.A. Hastaoglu, Comprehensive study of steam reforming of methane in membrane reactors, J. Energy Resour. Technol. 138 (2016), https://doi.org/
10.1115/1.4032733.
[138] R.Y. Chein, Y.C. Chen, J.N. Chung, Sweep gas flow effect on membrane reactor performance for hydrogen production from high-temperature water-gas shift
reaction, J. Membr. Sci. 475 (2015) 193–203, https://doi.org/10.1016/j.memsci.2014.09.046.
[139] J.S. Oklany, K. Hou, R. Hughes, A simulative comparison of dense and microporous membrane reactors for the steam reforming of methane, Appl. Catal. Gen.
170 (1998) 13–22, https://doi.org/10.1016/S0926-860X(98)00027-1.
[140] A.B. Leyko, A.K. Gupta, Temperature and pressure effects on hydrogen separation from syngas, J. Energy Resour. Technol. 135 (2013), https://doi.org/
10.1115/1.4024028.
[141] K. Ghasemzadeh, R. Zeynali, F. Ahmadnejad, A.A. Babalou, A. Basile, Investigation of palladium membrane reactor performance during ethanol steam
reforming using CFD method, Chem. Prod. Process Model. 11 (2016) 51–55, https://doi.org/10.1515/cppm-2015-0056.
[142] F.M. Ghohe, F. Hormozi, A numerical investigation on H2 separation by a conical palladium membrane, Int. J. Hydrogen Energy 44 (2019) 10653–10665,
https://doi.org/10.1016/j.ijhydene.2019.02.149.
[143] W.H. Chen, P.C. Kuo, Y.L. Lin, Evolutionary computation for maximizing CO2 and H2 separation in multiple-tube palladium-membrane systems, Appl. Energy
235 (2019) 299–310, https://doi.org/10.1016/j.apenergy.2018.10.106.
[144] H. Abdi, N. Pourmahmoud, J. Soltan, A novel CFD simulation of H2 separation by Pd-based helical and straight membrane tubes, Kor. J. Chem. Eng. 37 (2020)
2041–2053, https://doi.org/10.1007/s11814-020-0657-4.
[145] R. Ben-mansour, M.A. Haque, A. Harale, S.N. Paglieri, F.S. Alrashed, M. Raghib, et al., Comprehensive parametric investigation of methane reforming and
hydrogen separation using a CFD model, Energy Convers. Manag. 249 (2021), 114838, https://doi.org/10.1016/j.enconman.2021.114838.
[146] Y.S. Sanusi, E.M.A. Mokheimer, Performance analysis of a membrane-based reformer-combustor reactor for hydrogen generation, Int. J. Energy Res. 43 (2019)
189–203, https://doi.org/10.1002/er.4250.
[147] M. Vadrucci, F. Borgognoni, A. Moriani, A. Santucci, S. Tosti, Hydrogen permeation through Pd-Ag membranes: surface effects and Sieverts’ law, Int. J.
Hydrogen Energy 38 (2013) 4144–4152, https://doi.org/10.1016/j.ijhydene.2013.01.091.
[148] A.I. Livshits, The hydrogen transport through the metal alloy membranes with a spatial variation of the alloy composition: potential diffusion and enhanced
permeation, Int. J. Hydrogen Energy 42 (2017) 13111–13119, https://doi.org/10.1016/j.ijhydene.2017.04.016.
[149] H. Kreuder, T. Boeltken, M. Cholewa, J. Meier, P. Pfeifer, R. Dittmeyer, Heat storage by the dehydrogenation of methylcyclohexane – experimental studies for
the design of a microstructured membrane reactor, Int. J. Hydrogen Energy 41 (2016) 12082–12092, https://doi.org/10.1016/j.ijhydene.2016.05.140.
[150] G. Bruni, C. Rizzello, A. Santucci, D. Alique, M. Incelli, S. Tosti, On the energy efficiency of hydrogen production processes via steam reforming using
membrane reactors, Int. J. Hydrogen Energy 44 (2019) 988–999, https://doi.org/10.1016/j.ijhydene.2018.11.095.
[151] D.S.A. Simakov, M. Sheintuch, Model-based optimization of hydrogen generation by methane steam reforming in autothermal packed-bed membrane
reformer, AIChE J. 57 (2011) 525–541, https://doi.org/10.1002/aic.12265.
[152] A. Iulianelli, T. Longo, S. Liguori, A. Basile, Production of hydrogen via glycerol steam reforming in a Pd-Ag membrane reactor over Co-Al2O3 catalyst, Asia
Pac. J. Chem. Eng. (2010) 138–145, https://doi.org/10.1002/apj.365.
[153] R.J. Byron Smith, L. Muruganandam, S. Murthy Shekhar, CFD analysis of water gas shift membrane reactor, Chem. Eng. Res. Des. 89 (2011) 2448–2456,
https://doi.org/10.1016/j.cherd.2011.02.031.
[154] B. Lee, S.W. Yun, S. Kim, J. Heo, Y.T. Kim, S. Lee, et al., CO2 reforming of methane for H2 production in a membrane reactor as CO 2 utilization: computational
fluid dynamics studies with a reactor geometry, Int. J. Hydrogen Energy 4 (2019) 2298–2311, https://doi.org/10.1016/j.ijhydene.2018.09.184.
[155] M.A. Soria, D. Barros, L.M. Madeira, Hydrogen production through steam reforming of bio-oils derived from biomass pyrolysis: thermodynamic analysis
including in situ CO2 and/or H2 separation, Fuel 244 (2019) 184–195, https://doi.org/10.1016/j.fuel.2019.01.156.
[156] A. Iulianelli, K. Ghasemzadeh, M. Marelli, C. Evangelisti, A supported Pd-Cu/Al2O3 membrane from solvated metal atoms for hydrogen separation/
purification, Fuel Process. Technol. 195 (2019), https://doi.org/10.1016/j.fuproc.2019.106141.
[157] A. Caravella, F. Scura, G. Barbieri, E. Drioli, Sieverts law empirical exponent for PD-based membranes: critical analysis in pure H2 permeation, J. Phys. Chem.
B 114 (2010) 6033–6047, https://doi.org/10.1021/jp1006582.
[158] S. Barison, S. Fasolin, S. Boldrini, A. Ferrario, M. Romano, F. Montagner, et al., PdAg/alumina membranes prepared by high power impulse magnetron
sputtering for hydrogen separation, Int. J. Hydrogen Energy 43 (2018) 7982–7989, https://doi.org/10.1016/j.ijhydene.2018.03.065.
[159] J.W. Phair, R. Donelson, Developments and design of novel (non-palladium-based) metal membranes for hydrogen separation, Ind. Eng. Chem. Res. 45 (2006)
5657–5674, https://doi.org/10.1021/ie051333d.
[160] B.D. Morreale, M.V. Ciocco, R.M. Enick, B.L. Morsi, B.H. Howard, A.V. Cugini, et al., The permeability of hydrogen in bulk palladium at elevated temperatures
and pressures, J. Membr. Sci. 212 (2003) 87–97, https://doi.org/10.1016/S0376-7388(02)00456-8.
[161] T. Maneerung, K. Hidajat, S. Kawi, Ultra-thin (<1μm) internally-coated Pd-Ag alloy hollow fiber membrane with superior thermal stability and durability for
high temperature H2 separation, J. Membr. Sci. 452 (2014) 127–142, https://doi.org/10.1016/j.memsci.2013.10.040.
[162] A. Basile, S. Campanari, G. Manzolini, A. Iulianelli, T. Longo, S. Liguori, et al., Methane steam reforming in a Pd-Ag membrane reformer: an experimental study
on reaction pressure influence at middle temperature, Int. J. Hydrogen Energy 36 (2011) 1531–1539, https://doi.org/10.1016/j.ijhydene.2010.10.101.
[163] K. Ghasemzadeh, S. Liguori, P. Morrone, A. Iulianelli, V. Piemonte, A.A. Babaluo, et al., H2 production by low pressure methanol steam reforming in a dense
Pd-Ag membrane reactor in co-current flow configuration: experimental and modeling analysis, Int. J. Hydrogen Energy 38 (2013) 16685–16697, https://doi.
org/10.1016/j.ijhydene.2013.06.001.
[164] T. Boeltken, A. Wunsch, T. Gietzelt, P. Pfeifer, R. Dittmeyer, Ultra-compact microstructured methane steam reformer with integrated Palladium membrane for
on-site production of pure hydrogen: experimental demonstration, Int. J. Hydrogen Energy 39 (2014) 18058–18068, https://doi.org/10.1016/j.
ijhydene.2014.06.091.
[165] L.N. Baloyi, B.C. North, H.W. Langmi, B.J. Bladergroen, T.V. Ojumu, The production of hydrogen through the use of a 77 wt% Pd 23 wt% Ag membrane water
gas shift reactor, S. Afr. J. Chem. Eng. 22 (2016) 44–54, https://doi.org/10.1016/j.sajce.2016.11.001.
[166] A. Caravella, G. Barbieri, E. Drioli, Effect of the concentration polarization on the hydrogen permeation through pd-based membranes, Chem Eng Trans 17
(2009) 1681–1686, https://doi.org/10.3303/CET0917281.
[167] C. Zhao, A. Goldbach, H. Xu, Low-temperature stability of body-centered cubic PdCu membranes, J. Membr. Sci. 542 (2017) 60–67, https://doi.org/10.1016/j.
memsci.2017.07.049.
[168] H. Gao, J.Y.S. Lin, Y. Li, B. Zhang, Electroless plating synthesis, characterization and permeation properties of Pd-Cu membranes supported on ZrO2 modified
porous stainless steel, J. Membr. Sci. 265 (2005) 142–152, https://doi.org/10.1016/j.memsci.2005.04.050.
[169] N. Pomerantz, Y.H. Ma, Novel method for producing high H2 permeability Pd membranes with a thin layer of the sulfur tolerant Pd/Cu fcc phase, J. Membr.
Sci. 370 (2011) 97–108, https://doi.org/10.1016/j.memsci.2010.12.045.
[170] S.J. Xu, W.J. Thomson, Oxygen permeation rates through ion-conducting perovskite membranes, Chem. Eng. Sci. 54 (1999) 3839–3850, https://doi.org/
10.1016/S0009-2509(99)00015-9.
[171] P. Kirchen, D.J. Apo, A. Hunt, A.F. Ghoniem, A novel ion transport membrane reactor for fundamental investigations of oxygen permeation and oxy-
combustion under reactive flow conditions, Proc. Combust. Inst. 34 (2013) 3463–3470, https://doi.org/10.1016/j.proci.2012.07.076.

33
M.A. Habib et al. Case Studies in Thermal Engineering 49 (2023) 103359

[172] A. Hunt, G. Dimitrakopoulos, P. Kirchen, A.F. Ghoniem, Measuring the oxygen profile and permeation flux across an ion transport (La0.9Ca0.1FeO3 -δ)
membrane and the development and validation of a multistep surface exchange model, J. Membr. Sci. 468 (2014) 62–72, https://doi.org/10.1016/j.
memsci.2014.05.043.
[173] A. Hunt, G. Dimitrakopoulos, A.F. Ghoniem, Surface oxygen vacancy and oxygen permeation flux limits of perovskite ion transport membranes, J. Membr. Sci.
489 (2015) 248–257, https://doi.org/10.1016/j.memsci.2015.03.095.
[174] A. Hunt, G. Dimitrakopoulos, A.F. Ghoniem, Surface oxygen vacancy and oxygen permeation flux limits of perovskite ion transport membranes, J. Membr. Sci.
489 (2015) 248–257, https://doi.org/10.1016/j.memsci.2015.03.095.
[175] A. Behrouzifar, A.A. Asadi, T. Mohammadi, A. Pak, Experimental investigation and mathematical modeling of oxygen permeation through dense
Ba0.5Sr0.5Co0.8Fe0.2O3− (BSCF) perovskite-type ceramic membranes, Ceram. Int. 38 (2012) 4797–4811, https://doi.org/10.1016/j.ceramint.2012.02.068.
[176] J.P. Collins, J.D. Way, Preparation and characterization of a composite palladium-ceramic membrane, Ind. Eng. Chem. Res. 32 (1993) 3006–3013, https://doi.
org/10.1021/ie00024a008.
[177] V. Jayaraman, Y.S. Lin, Synthesis and hydrogen permeation properties of ultrathin palladium-silver alloy membranes, J. Membr. Sci. 104 (1995) 251–262,
https://doi.org/10.1016/0376-7388(95)00040-J.
[178] R.C. Hurlbert, J.O. Konecny, Diffusion of hydrogen through palladium, J. Chem. Phys. 34 (1961) 655–658, https://doi.org/10.1063/1.1701003.
[179] I. Ali-Khan, K.J. Dietz, F.G. Waelbroeck, P. Wienhold, The rate of hydrogen release out of clean metallic surfaces, J. Nucl. Mater. 76–77 (1978) 337–343,
https://doi.org/10.1016/0022-3115(78)90167-8.
[180] A.B. Antoniazzi, A.A. Haasz, P.C. Stangeby, The effect of adsorbed carbon and sulphur on hydrogen permeation through palladium, J. Nucl. Mater. 162–164
(1989) 1065–1070, https://doi.org/10.1016/0022-3115(89)90410-8.
[181] S. Yan, H. Maeda, K. Kusakabe, S. Morooka, Thin palladium membrane formed in support pores by metal-organic chemical vapor deposition method and
application to hydrogen separation, Ind. Eng. Chem. Res. 33 (1994) 616–622, https://doi.org/10.1021/ie00027a019.
[182] T.L. Ward, T. Dao, Model of hydrogen permeation behavior in palladium membranes, J. Membr. Sci. 153 (1999) 211–231, https://doi.org/10.1016/S0376-
7388(98)00256-7.
[183] A. Caravella, G. Barbieri, E. Drioli, Modelling and simulation of hydrogen permeation through supported Pd-alloy membranes with a multicomponent
approach, Chem. Eng. Sci. 63 (2008) 2149–2160, https://doi.org/10.1016/j.ces.2008.01.009.
[184] A. Bhargav, G.S. Jackson, Thermokinetic modeling and parameter estimation for hydrogen permeation through Pd0.77Ag0.23 membranes, Int. J. Hydrogen
Energy 34 (2009) 5164–5173, https://doi.org/10.1016/j.ijhydene.2009.04.010.
[185] A. Rossi, G. Lamonaca, A. Santucci, S. Tosti, Validation of a dynamic model of hydrogen permeation through Pd-based membranes, Int. J. Greenh. Gas Control
5 (2011) 521–530, https://doi.org/10.1016/j.ijggc.2010.11.009.
[186] C.H. Lee, Y.S. Jo, Y. Park, H. Jeong, Y. Kim, H. Sohn, et al., Unconventional hydrogen permeation behavior of Pd/BCC composite membranes and significance
of surface reaction kinetics, J. Membr. Sci. 595 (2020), 117506, https://doi.org/10.1016/j.memsci.2019.117506.
[187] A. Abuelyamen, R. Ben-Mansour, M.A. Habib, V.R. Manga, A. Harale, S. Paglieri, A. Alsayoud, Poisonous effect of carbon bearing species on adsorption of
hydrogen on Pd-membrane surfaces, in: International Journal of Hydrogen Energy, 2023 (Press).
[188] C. Liu, X. Zhang, J. Zhai, X. Li, X. Guo, G. He, Research progress and prospects on hydrogen separation membranes, Clean Energy 7 (Issue 1) (February 2023)
217–241, https://doi.org/10.1093/ce/zkad014.
[189] M. El-Shafie, Hydrogen separation using palladium-based membranes: assessment of H2 separation in a catalytic plasma membrane reactor, Int. J. Energy Res.
46 (3) (2022) 3572–3587, https://doi.org/10.1002/er.7406.

34

You might also like