Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Journal of Cleaner Production 324 (2021) 129241

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Sustainable ammonia production from steam reforming of biomass-derived


glycerol in a heat-integrated intensified process: Modeling and
feasibility study
Mohammad Hasan Khademi a, *, Mohammad Lotfi-Varnoosfaderani b
a
Department of Chemical Engineering, College of Engineering, University of Isfahan, P.O. Box 81746-73441, Isfahan, Iran
b
Department of Chemical and Petroleum Engineering, Sharif University of Technology, P.O. Box 11155-9465, Tehran, Iran

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

Handling Editor: Kathleen Aviso Currently, ammonia, as a clean and sustainable energy carrier, is intensively synthesized from its elements during
the Haber-Bosch technology. This process requires a large amount of energy and emits numerous amounts of
Keywords: carbon dioxide, because hydrogen is dominantly produced from fossil fuels through reforming processes.
Sustainable ammonia production Biomass-derived glycerol steam reforming is an attractive alternative to traditional reforming for reducing the
Glycerol steam reforming
dependence on hydrocarbon resources and mitigating climate change. This research aims to intensify a heat-
Methane tri-reforming
integrated process for the co-production of ammonia and syngas from glycerol valorization. In this process,
Heat-integrated process
Synthesis gas glycerol reforming continuously provides hydrogen needed for ammonia synthesis, and the liquid glycerol is
simultaneously vaporized by heat generated from ammonia synthesis. Methane tri-reforming acts as a heat
source to drive glycerol reforming; at the same time, the effluent gas produced through glycerol reforming is
recycled to the tri-reforming side to reduce the greenhouse gas emissions. The role of different parameters on the
process performance is identified by a one-dimensional heterogeneous model. Numerical results show that by
adjusting the adequate operating conditions, glycerol and methane conversion >95%, nitrogen conversion
>25%, glycerol dryness fraction = 1.0, and syngas with hydrogen to carbon monoxide ratio above 2.0, suitable
for the Fischer-Tropsch and methanol synthesis processes, can be achieved. In addition, this heat-integrated
intensified process is promising in terms of energy saving, environmental pollution mitigation, feasibility and
effectiveness for industrial-scale application; however, experimental proof-of-concept is required to ensure the
safe operability of this process.

1. Introduction explored and investigated, but ammonia production rates are still far
from satisfactory (Soloveichik, 2019). To date, three unique approaches
Ammonia, as a potential hydrogen carrier and a carbon-free fuel, is to ammonia production via chemical looping technology have been re­
one of the essential chemicals in the petrochemical industry. It is mainly ported: (i) H2 chemical looping, a two-step process where metal nitrides
used to produce fertilizers like ammonium nitrate, urea, and ammonium are directly reduced to a metallic state in the presence of H2 and back to
phosphate, as well as many products such as explosives, fibers, plastics, a metal nitride under pure N2 (Shan et al., 2018); (ii) H2O chemical
and polymers. looping, a three-step process where a metal oxide precursor is reduced to
There are various pathways for ammonia production, such as Haber- its metallic form before undergoing nitridation and subsequent hydro­
Bosch process, electro-catalysis (Soloveichik, 2019), photo-catalysis lysis to yield ammonia and restore the original oxide (Sarafraz and
(Medford and Hatzell, 2017), and chemical looping (Weng et al., Christo, 2021); and (iii) alkali or alkaline-earth metal hydride chemical
2021) technologies. In the Haber-Bosch process, as the earliest tech­ looping, a two-step process where N2 is reduced directly by the metal
nology, nitrogen and hydrogen in the gas phase react over magnetic iron hydride to form a metal imide that then undergoes disproportionation
oxide catalyst at high temperatures and pressures. Recently, under H2 to produce NH3 and restore the metal hydride (Gao et al.,
electro-catalysis and photo-catalysis strategies have been actively 2018). In comparison with the conventional Haber-Bosch process, the

* Corresponding author.
E-mail address: m.khademi@eng.ui.ac.ir (M.H. Khademi).

https://doi.org/10.1016/j.jclepro.2021.129241
Received 12 July 2021; Received in revised form 30 September 2021; Accepted 2 October 2021
Available online 5 October 2021
0959-6526/© 2021 Elsevier Ltd. All rights reserved.
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

chemical looping ammonia synthesis technology offers the following worldwide production is increasing recently; moreover, it is anticipated
advantages: 1) functioning at ambient pressure, minimizing cost and to expand its global use continuously (U.S. Energy Information
safety concerns; 2) relying less on catalysts, reducing the costs of their Administration, International Energy Statistics, Biofuels Production).
production and recycling; and 3) using water as a hydrogen source, Biodiesel decreases CO2 emissions by more than 52% (Okoye et al.,
reducing energy consumption and associated CO2 emissions (Gálvez 2017); however, it is not competitive from the price viewpoint than
et al., 2007). Despite the advantages of chemical looping technology, fossil fuels. Glycerol (nearly 10%), as the main by-product, is produced
approximately 90% of global ammonia is still produced commercially by the conversion of animal fat or vegetable oils into biodiesel through
through the Haber-Bosch process, as the most common route (Dincer the catalytic trans-esterification process (Ma and Hanna, 1999). One
and Bicer, 2018), because there is no other economically viable for cost-effective way to reduce the price of biodiesel is to use glycerol to
synthesis of ammonia (Sarafraz and Christo, 2021). produce hydrogen (Silva et al., 2015). Steam reforming of glycerol is the
Although hydrogen can be produced by different processes such as most well-known and preferable technique for H2 production (Patch­
steam reforming of fossil fuels, partial oxidation of heavy oil, water aravorachot et al., 2019) because it can theoretically make 7 mol of
electrolysis, and biomass gasification, this is one of the principal vital hydrogen per mole of glycerol as follows:
challenges in the ammonia synthesis through Haber-Bosch process.
C3 H8 O3 (g) + 3H2 O(g) ↔ 7H2 (g) + 3CO2 (g) (1)
Steam reforming is the most common technology for producing
hydrogen from fossil fuels, most often natural gas. More than 72% of Recently, from the mathematical modeling and simulation perspec­
ammonia is emanated from methane steam reforming, which releases tives, some researchers such as Khodabandehloo et al. (2020) and Unlu
large amounts of CO2 to the atmosphere (about 2–3 tons of CO2 per each and Hilmioglu (2020) have focused on glycerol’s potential to convert
ton of ammonia) (Anderson et al., 2008), leading to global warming and hydrogen, and others like Saidi and Moradi (2020) and Macedo et al.
environmental pollution. The use of biomass-based sources (Gilbert (2019) used a hydrogen perm-selective membrane reactor instead of a
et al., 2014), water electrolysis (Xu et al., 2018), injection of produced conventional glycerol steam reformer to improve the glycerol conver­
CO2 into oil reservoirs for enhanced oil recovery (Bicer and Dincer, sion and hydrogen yield. However, the main challenge in the glycerol
2018), chemical looping combustion technology (Zhu et al., 2020), steam reforming process is the emissions of greenhouse gases, including
carbon capture and storage (CCS) methods (Global Status Report - CH4, CO2, and H2O into the atmosphere. One way to meet this challenge
Global CCS Institute), and improving the efficiency of the process (Bicer is to use the sorption-enhanced glycerol steam reforming process (Yang
et al., 2016) are promising alternatives to reduce CO2 emissions during et al., 2020), which involves the CO2 sorption along with the catalytic
ammonia production. reaction. Although this system reduces CO2 emissions, the major
From an environmental perspective, Gilbert et al. (2014) and Arora disadvantage of this process is the regeneration of CO2 sorbents in
et al. (2016) showed that using the biomass gasification approach for another reactor, which increases capital costs. An excellent solution to
ammonia production can reduce CO2 emissions by 54–68% compared to diminish the CO2 emissions in the glycerol steam reformer is to use the
the natural gas-based conventional ammonia synthesis process. Bicer exiting gas stream, containing carbon dioxide and methane as a feed­
et al. (2016) compared life cycle assessment of various ammonia pro­ stock of methane tri-reforming reaction.
duction routes from hydrogen provided by water electrolysis, in which The methane tri-reforming consists of dry reforming, steam reform­
electricity was supplied from municipal waste, biomass, nuclear, and ing, and partial oxidation of methane, that combination of these re­
hydropower. They concluded that municipal waste- and actions improves the reactor efficiency (Khademi and Alipour-Dehkordi,
hydropower-based water electrolysis has the lowest environmental as­ 2021). The advantages of this process are: 1) this is an autothermal re­
pects. Chisalita et al. (2020) conducted an environmental assessment of action, 2) carbon formation in this process is reduced due to the exis­
hydrogen production for ammonia synthesis through iron-based chem­ tence of steam and oxygen (Chein et al., 2017), 3) it does not require a
ical looping, water electrolysis, and methane steam reforming in com­ source of pure CO2 as a reactant, so CO2 produced by the glycerol steam
bination with CCS technology based on liquid-gas absorption by chilled reforming process can be fed to this reaction, and 4) the synthesis gas
ammonia and amines. The analysis indicated that the chemical looping with proper H2/CO ratio for the downstream customary units can be
hydrogen production and fossil sources-based electrolysis have the produced by adjusting the feed composition (Chein et al., 2017).
lowest and highest overall emissions to soil, water, and air, respectively. On the other hand, the heat required to drive the glycerol steam
From a cost viewpoint, Zhang et al. (2019) carried out a comparative reforming, as an endothermic reaction, can be provided by the methane
techno-economic analysis of the Haber-Bosch process integrated with tri-reforming. Using a heat integration strategy for these two reactions
methane steam reforming, water electrolysis, and biomass gasification. can significantly reduce energy consumption and greenhouse gas
They showed that biomass gasification and solid-oxide electrolysis have emissions. Numerous theoretical studies have been performed on the
the lowest and highest efficiencies, respectively, while water electrolysis thermal integration of methane tri-reforming or glycerol reforming with
is the worst method in terms of economics and is not economically other reactions. For example, two thermally coupled reactors were
feasible due to high electrical consumption. The coupling of CCS tech­ simulated, in which the methane tri-reforming reaction provided the
nology with fossil fuels-based ammonia production is more economical necessary heat for steam (Rahnama et al., 2014) and dry (Farniaei et al.,
than water electrolysis (Global Status Report - Global CCS Institute). 2014) reforming of methane. These processes were able to produce
However, the high energy intensity of solvent regeneration and the synthesis gas with different H2/CO ratios. Iliuta et al. (2012) suggested a
higher risk of fugitive emissions due to high solvent volatility are the heat-integrated process for aqueous-phase glycerol reforming and
main drawbacks of solvents utilization in liquid-gas absorption-based dimethyl ether synthesis reaction; in addition, Thirabunjongcharoen
CCS technology (Mikulčić et al., 2019). et al. (2020) offered a shell and tube heat-exchanger reactor to integrate
Even though methane steam reforming is currently the predominant methanol synthesis and aqueous-phase glycerol reforming reaction. In
hydrogen production route for ammonia synthesis and the least energy these two cases, aqueous-phase glycerol reforming produced biosyngas
intensive technology among other methods, ammonia production costs for dimethyl ether and methanol synthesis, which in return these syn­
can increase with natural gas prices in parallel (Bicer et al., 2016). In thesis reactions supplied sufficient energy for the endothermic reaction.
addition, fossil fuel depletion, global warming, and environmental Recently, Khazayialiabad et al. (2021) conducted a conceptual evalua­
problems are the main drawbacks of this hydrogen production route, tion of a thermally coupled reactor containing nitrobenzene hydroge­
which can be sufficient reasons to intensify and improve the ammonia nation and glycerol steam reforming for co-production of aniline and
synthesis process and to use renewable biofuels as a suitable alternative hydrogen.
to fossil fuels. Since the ammonia synthesis reaction is highly exothermic,
One of the most important renewable fuels is biodiesel, whose annual removing the heat of the reaction is another critical issue in this reactor

2
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

Fig. 1. Flow diagram of the (a) conventional ammonia synthesis process and (b) heat-integrated intensified process.

Fig. 2. Conceptual diagram of three heat-integrated multi-tubular reactors for proposed process.

• According to the compressibility factor calculation based on the Peng-Robinson-Stryjek-Vera equation of state for ammonia synthesis (i.e., 1.0), glycerol steam
reforming (i.e., 0.982) and methane tri-reforming (i.e., 1.0), the behaviour of the gas can be considered ideal.
• Steady-state condition
• A uniform bed void fraction distribution is considered in both sides.
• The axial heat and mass diffusion can be disregarded compared to the bulk flow according to the large enough Peclet number (Pe > 100) (Weigand, 2004).
• Catalyst deactivation is ignored.
• The conduction resistance in the solid wall of tubes is ignored.
• The intra-particle mass and heat transfer restriction, defined as the effectiveness factor, is reported by De Groote and Froment (1996) and (Arab Aboosadi et al.,
2011) for methane tri-reforming. An empirical correlation for the effectiveness factor of catalyst size of 6–10 mm is developed by Dyson and Simon (1968) for
ammonia synthesis reaction as a function of conversion and temperature. This parameter is also assumed equal to one for glycerol steam reforming due to the small
diameter of the catalyst (i.e., 115 μm) and Thiele modulus <0.4.

to enhance the degree of conversion and prevent catalyst deactivation. technology in terms of low-temperature level and ammonia production.
There are three cooling technologies to control the temperature of the Recently, Iranshahi and co-authors suggested recuperative coupling
reactor, namely direct cooling (Carvalho et al., 2014), quench cooling reactors as a direct cooling strategy for ammonia synthesis. They pre­
(Sadeghi and Kavianiboroujeni, 2008), and indirect cooling (Dashti sented different configurations for thermally coupled reactors, in which
et al., 2006). A comparison between these cooling methods was per­ ammonia synthesis reaction was cooled by cyclohexane dehydrogena­
formed by Khademi and Sabbaghi (2017) and Azarhoosh et al. (2014). tion (Ghani et al., 2019), 2-butanol dehydrogenation (Ghani and Iran­
They showed that direct cooling was the most efficient cooling shahi, 2019), naphtha reforming (Shakeri et al., 2019), and methyl

3
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

formate (Nikzad and Iranshahi, 2021). 3. Methodology


In this work, a novel heat-integrated intensified process is proposed
for the sustainable ammonia production from steam reforming of 3.1. Mathematical model
biomass-derived glycerol to improve the NH3 synthesis process in order
to save energy and significantly reduce CO2 emissions, as the most Temperature and concentration changes along the reactor axis are
important cause of greenhouse effect. This process is also able to pro­ specified by a 1D heterogeneous model. A differential element is
duce synthesis gas, as a valuable product, through methane tri- considered along each reactor to derive the mass and energy balance
reforming reaction. In this regard, the effect of various parameters on equations for each side. The governing equations are simplified by the
the performance of the process, including ammonia production, dryness following assumptions:
fraction, methane and glycerol conversion, and H2/CO ratio in the
synthesis gas was evaluated using a heterogeneous one-dimensional 3.1.1. Mass balance for evaporation-side
model. Due to the existence of a two-phase flow in the evaporation-side of
reactor I and II, a mass balance equation for this side is expressed as:
2. Process description dx Ntube πdt TP ( b )
F0C3 H8 O3 + U1− 2 T2 − T1 = 0 for ​ x < 1 (2)
dz λ
2.1. Conventional ammonia synthesis process
at z = 0 x=0 (3)
A flow diagram of the conventional ammonia synthesis process is
shown in Fig. 1(a). Firstly, natural gas is desulfurized and then mixed where T2b = 625 K is the boiling point of glycerol, and dryness fraction,
with steam before entering the reformer. The carbon monoxide gener­ x, is defined as the ratio of the mass of vapor to the total mass of the
ated in the steam reformer is treated with water in a water-gas shift mixture.
converter to produce additional hydrogen. The steam changes to the
liquid phase in a condenser, and the discharged stream runs to the 3.1.2. Energy balance for evaporation-side
pressure swing adsorption (PSA). A pure hydrogen stream exits the PSA The two-phase flow can change to single-phase (vapor) due to the
unit, and the waste gases as well as methane makeup gas are combusted operating conditions, so to calculate the glycerol vapor temperature, the
with air to supply sufficient heat for the reformer. Finally, the hydrogen- sensible heat of the vapor must be taken into account for the single-
rich gas stream reacts with nitrogen to produce ammonia. phase. In this regard, an energy balance equation is given as:
( )
d Cp T2
F0C3 H8 O3 + Ntube πdt UTP
1− 2 (T2 − T1 ) = 0 for ​ x = 1 (4)
2.2. Heat-integrated intensified process dz

Fig. 1(b) indicates a flow diagram of the suggested process for the co- at z = z0 T2 = Tb2 (5)
production of ammonia and syngas. Glycerol in the liquid phase is first z0 is the position where the two-phase flow regime changes to a
evaporated in two equal-size heat-integrated multi-tubular reactors in single-phase flow. The two-phase heat transfer coefficient in the
series. After mixing the steam and glycerol vapor, if required, this stream evaporation-side was reported by Kandlikar (1990) as follows:
must be passed through a heat-exchanger to reach an adequate degree of
temperature for glycerol steam reforming reaction, and then entered hTP
= C1 CoC2 (25Frl )C5 + C3 BoC4 Ffl (6)
another heat-integrated multi-tubular reactor. A conceptual diagram of hl
these three heat-integrated multi-tubular reactors is shown in Fig. 2.
These reactors consist of a bundle of tubes placed in a shell. In reactors I hl = 0.023Re0.8 0.4
l Prl (Kl / dt ) (7)
and II, ammonia is synthesized in the tubes (also known as exothermic- ⎧
side), and glycerol is vaporized in the shell (also known as evaporation- ⎪


< 0.65 convective boiling region,
side). The heat required for the evaporation-side is continuously sup­ Co
C1 = 1.136, C2 = − 0.9, C3 = 667.2, C4 = 0.7, C5 = 0.3
(8)
⎪ > 0.65 nucleate boiling region
plied from the exothermic-side. In reactor III, glycerol steam reforming ⎪

C1 = 0.6683, C2 = − 0.2, C3 = 1, 058, C4 = 0.7, C5 = 0.3
takes place in the tubes (also known as endothermic-side), and methane
tri-reforming occurs in the shell (exothermic-side). The exothermic-side ( )0.8 ( )0.5
provides the heat required for the endothermic-side. As shown in Fig. 1 where Co = 1− x
x
ρg
ρl
q
is convection number, Bo = Gλ is boiling
(b), the outlet stream of the endothermic-side runs to the PSA. In order 2
to produce ammonia, pure hydrogen exited from the top of the PSA unit number, Frl = ρ2Ggdt is Froude number with all flow as liquid, and Ffl is a
l
is mixed with nitrogen, and recycled in parallel to the reactors I and II.



⎪ ρ υi ri in endothermic − side of reactor III
( ) ⎨ b
kc,i,k ak Ck ysi,k − ygi,k = ∑N
for solid − phase (9)

⎩ ρb
⎪ ηj υi,j rsj in exothermic − side of reactor I, II, III
j=1

Also, the waste gases (including: CH4, CO, CO2, and H2O) exited from fluid-dependent parameter which is calculated using the scheme sug­
the bottom of the PSA, as well as methane makeup gas and oxygen, are gested by Kandlikar (1990).
pressurized by a compressor, and fed to the entrance of the exothermic-
side of the reactor III to produce syngas. 3.1.3. Partial mass balances for component i

4
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

( )
d us,k Ck ygi,k ( ) ( )
d Ck us,k Cp,k Tgk ( ) πdt
+ kc,i,k ak Ck ygi,k − ysi,k = 0 for ​ fluid − phase (10) + hk ak Tgk − Tsk + βγ U1− 2 (Tg2 − Tg1 ) = 0 for ​ fluid
dz dz Ac,k
− phase (15)
at z = 0 ygi,k = yg,0
i,k (11)

at z = 0 Tgk = Tg,0 (16)


where υi takes values of +1 and − 1, respectively, depending on whether k

component i is produced or consumed; and υi,j stands for stoichiometric In the above equations, the subscript k represents the side number of
coefficient of component i in reaction j. each reactor, so that k = 1 and 2 stands for tube- and shell-side,
respectively.
3.1.4. Total mass balance {
1 for ​ shell − side ​ of ​ reactor ​ III
( ) ( ) β= (17)
d us,k Ck ∑nk − 1 for ​ tube − side ​ of ​ reactor ​ I, II, III
+ a k Ck kc,i,k ygi,k − ysi,k = 0 (12)
dz i=1 {
1 for ​ tube − side
γ= (18)
at z = 0 us,k = u0s,k (13) Ntube for ​ shell − side

3.1.6. Momentum balance


3.1.5. Energy balances
{ }
dP (1 − εk )2 μk us,k (1 − εk )u2s,k ρk ρ us,k dus,k
εk k + 150 + 1.75 + k =0 (19)
dz ε2k d2p,k ε2k dp,k εk dz

⎧ ∑ n
⎪ ( )

⎪ ρb υi − ΔHf,i ri in ​ endothermic − side ​ of ​ reactor ​ III ​

⎨ i=1
( )
hk ak Tsk − Tgk = for ​ solid − phase (14)

⎪ ∑
N
⎪ ηj ( − ΔHr )j rsj
⎩ ρb
⎪ in ​ exothermic − side ​ of ​ reactor ​ I, II, III
j=1

Table 1
Auxiliary correlations.
Parameter Equation Ref.

Density of fluid mixture P ∑ n


g
ρ = y Mi
RT i=1 i
g
Viscosity of fluid mixture ∑n y μ Wilke (1950)
μ = ∑n i ig
i=1 j=1 yj φij [ ( )0.5 ( )−
[ ( )] 0.25 ]2
Mi − 0.5 μi Mi
φij = 8 1 + 1+
Mj μj μj
Gas viscosity μi = Ai + Bi T + Ci T2 Yaws (1999)
Molar heat capacity of fluid mixture n

Cp = ygi Cp,i
i=1
Gas molar heat capacity Cp,i = Ai + Bi T + Ci T2 + Di T3 + Ei T4 Coker (2007)
Overall two-phase heat transfer coefficient for reactor I & II 1 1 1
= + TP
UTP hi h
Overall heat transfer coefficient for reactor III 1 1 1
= +
U hi ho
Heat transfer coefficient between the gas phase and tube wall Kg Cussler and Cussler (2009)
h = 0.023Re0.8 Pr0.3
( dt ) g
Heat transfer coefficient between the gas phase and packed-tube wall dp K Leva et al. (1948)
h = 3.5Re0.7 exp − 4.6 for ​ cooling ​ the ​ gas ​ phase
( dt ) dt
dp Kg Leva (1950)
h = 0.813Re0.9 exp − 6 for ​ heating ​ the ​ gas ​ phase
dt dt
g
Heat transfer coefficient between the gas phase and catalyst K McCabe et al. (1993)
h = 1.17Re0.585 Pr0.33
dp
1, 000Cp μ
Pr = g ∑n g
K i=1 yi Mi
Mass transfer coefficient between the catalyst and fluid phases − 2/3 g
kci = 1.17Re− 0.42 Sci u Cussler and Cussler (2009)
dp ug ρ
Re =
μ
μ
Sci =
ρDim
Molecular diffusion coefficient in a multi-component gas mixture 1− y Fairbanks and Wilke (1950)
Dim = ∑ yi
j
j∕
=i
Dij √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
Binary diffusion coefficient 0.00143T1.75 1/Mi + 1/Mj Poling et al. (2001)
Dij = √̅̅̅ 1/3 1/3 2
2P[νci + νcj ]

5
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

Table 2 Table 3
Kinetic parameters of the glycerol steam reforming reaction (Cheng et al., 2011). Base scenario for operating conditions of glycerol steam reforming,
Species A (mmol m− 2 − 1
s kPa− (a+b)
) E (kJ mol− 1) a b
methane tri-reforming, and ammonia synthesis reaction as well as the
geometry of reactors I, II, and III.
C3H8O3 3.81 59.8 0.48 0.34
H2 37.2 60.2 0.34 0.27 Parameter Value
CO2 11.6 61.8 0.39 0.41 Glycerol steam reforming reaction
CO 308.6 73.9 0.74 − 0.39
CH4 2,834.2 118.9 0.88 0.34 WHSV (h− 1) 20
H2O/C3H8O3 molar ratio 5
Inlet gas temperature (K) 740
Inlet gas pressure (atm) 5
at z = 0 Pk = P0k (20) Catalyst surface (m2/g)a 4.3
Catalyst diameter (m)a 1.15 × 10− 4

Catalyst density (kg m− 3)a 1,870


3.1.7. Auxiliary correlations
Methane tri-reforming reaction
A set of empirical correlations for estimating physical properties, CH4 molar flow rate (mol s− 1) 515
heat and mass transfer coefficients, and diffusion coefficient are listed in O2/CH4 molar ratio 0.6
Table 1. Inlet gas temperature (K) 1,000
Inlet gas pressure (atm) 20
Catalyst shape b 10-HOLE rings
Catalyst size (m)b 19 × 16
3.2. Kinetics model
Catalyst density (kg m− 3)b 1,870
Ammonia synthesis reaction
3.2.1. Ammonia synthesis
N2 molar flow rate (mol s− 1) 41.1
Nitrogen is converted to ammonia at high temperatures and pres­ N2/H2 molar ratio 1/3
sures through the magnetic iron oxide catalyst. Inlet gas temperature (K) 665
Inlet gas pressure (atm) 300
N2 + 3H2 ↔ 2NH3 ΔH298
r
​K
= − 92.4 ​ kJ ​ mol− 1
(21) Catalyst diameter (m)c 2.85 × 10− 3

Catalyst density (kg m− 3)c 1,870


The rate of this reaction was reported by Dyson and Simon (1968) as Number of tubes 250
follows: Tube diameter (m) 0.0254
Shell diameter (m) 0.5
( )[ ( )0.5 ( )0.5 ]
− E a3H a2NH3 Reactor length (m) 6
rNH3 = 2k0 exp K2a aN2 2 2 − (22) Bed porosity 0.4
RT aNH3 a3H2
Number of tubes 500
Tube diameter (m) 0.0254
log 10 Ka = − 2.691122log 10 T − 5.519265 × 10− 5 T + 1.848863 Shell diameter (m) 0.7
(23) Reactor length (m) 7
2, 001.6
× 10− 7 T2 + + 26, 899 Bed porosity 0.4
T
a
Cheng et al. (2011).
where k0 = 8.849 × 1014 , E = 170, 560 kJ mol− 1 , and Ka is equilibrium b
Rahimpour et al. (2012).
c
constant presented by Gillespie and Beattie (1930). Dyson and Simon (1968).

3.2.2. Glycerol steam reforming considered in this study, the carbon formation reactions including (30)–
The overall reaction (1) operates over the Ni/Al2O3 catalyst at at­ (32) have been neglected.
mospheric pressure and temperature of 723–823 K, which includes The rate of glycerol steam reforming was presented by Cheng et al.
glycerol decomposition and water-gas shift reactions as shown in Eqs. (2011) based on the power-law equation as follows:
(24) and (25) (Cheng et al., 2011): ( )
− Ei ai
ri = Ai exp P Pbi Acs (33)
C3 H8 O3 → 3CO + 4H2 ΔH298
r
​K
= 251 ​ kJ ​ mol− 1
(24) RT C3 H8 O3 H2 O

CO + H2 O ↔ CO2 + H2 ΔH298
r
​K
= − 41.1 ​ kJ ​ mol− 1
(25) where A and E denote the pre-exponential factor and activation energy,
respectively, and the reaction is also ath order concerning C3H8O3 and
However, reactions (24) and (25) may also be accompanied by re­ bth order concerning H2O. The kinetic parameters of glycerol steam
actions (26)–(32) as the side reactions (Cheng et al., 2011): reforming are tabulated in Table 2.
C3 H8 O3 + 5H2 ↔ 3CH4 + 3H2 O (26)
3.2.3. Tri-reforming of methane
CO + 3H2 ↔ CH4 + H2 O (27) The methane tri-reforming reactions, including methane steam
reforming, reverse CO2 methanation, water-gas shift, and complete
CO2 + 4H2 ↔ CH4 + 2H2 O (28) methane oxidation occur according to Eq. (34)‒(37) at 800–1,700 K and
10–30 bar over the Ni/MgO/CeZrO2/Al2O3 catalyst (Alipour-Dehkordi
CO2 + CH4 ↔ 2CO + 2H2 (29) and Khademi, 2019). It should be noted that since methane dry
reforming can be derived from reaction (34) minus reaction (36), so this
CH4 ↔ 2H2 + C (30) is not an independent reaction (Khademi et al., 2021). The accuracy of
these reactions was validated by Arab Aboosadi et al. (2011).
C + H2 O ↔ CO + H2 (31)
/
CH4 + H2 O ↔ CO + 3H2 ΔH298 r
​K
= 206 ​ kJ mol (34)
2CO ↔ C + CO2 (32)
/
The formation of CH4, as a by-product, must be taken into account CH4 + 2H2 O ↔ CO2 + 4H2 ΔH298
r
​K
= 164.9 ​ kJ mol (35)
according to reactions (26)–(29). On the other hand, from the findings of /
Wang et al. (2010), it can be prevented the carbon deposition at high CO + H2 O ↔ CO2 + H2 ΔH298
r
​K
= − 41.1 ​ kJ mol (36)
steam to glycerol ratios. Since the high steam to glycerol ratio is

6
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

/ /
CH4 + 2O2 ↔ CO2 + 2H2 O ΔH298 ​K
= − 802.7 ​ kJ mol (37) Fexo
(43)
r H2
H2 CO ​ ratio =
Xu and Froment (1989) reported a kinetic rate formula (Eq. 38–40) Fexo
CO

for methane steam reforming reactions (34)–(36), and De Smet et al.


F0N2 − Fexo
(2001) modified the kinetic rate proposed by Trimm and Lam (1980) for Nitrogen ​ conversion =
N2
(44)
methane complete oxidation (Eq. (41)) over the Ni-based catalyst. F0N2
( )
k1 P3H2 PCO 1 F0C3 H8 O3 − Fendo
r1 = 2.5 PCH4 PH2 O − × 2 (38) Glycerol ​ conversion =
C 3 H 8 O3
(45)
PH2 KI ∅ F0C3 H8 O3
( )
k2 P4H2 PCO2 1 F0CH4 − Fexo
r2 = 3.5 PCH4 P2H2 O − × (39) Methane ​ conversion =
CH4
(46)
PH2 KІІ ∅2 F0CH4
( )
k3 PH2 PCO2 1
r3 = PCO PH2 O − × (40)
PH2 KІІІ ∅2

k4a PCH4 PO2 k4b PCH4 PO2


r4 = ( )2 + (41)
1 + KCHC4 PCH4 + KOC2 PO2
1 + KCHC4 PCH4 + KOC2 PO2
3.4. Numerical solution

PH2 O
∅ = 1 + KCO PCO + KH2 PH2 + KCH4 PCH4 + KH2 O (42) The governing equations consist of algebraic and ordinary differen­
PH2 tial equations, in which the ordinary differential equations are dis­
The parameters of kinetic constants according to the Arrhenius law cretized into non-linear algebraic equations using the backward finite
difference method. In this regard, the length of the reactor is divided into


⎪ ṁNH3 LHVNH3 + ṁsyngas LHVsyngas

⎪ ∑3 i = steam, N2 , O2



⎪ ṁglycerol LHVglycerol + ṁCH4 LHVCH4 + ṁi hi

⎪ i=1




⎨ for heat − integrated intensified process
energy efficiency = (47)

⎪ ṁNH3 LHVNH3

⎪ i = steam, N2

⎪ ∑2

⎪ ṁ LHV CH4 + ṁi hi

⎪ CH 4 i=1




for methane steam reforming − based ammonia synthesis process

Table 4
Validation of model prediction with experimental data for glycerol steam reformer.
Molar ratio Temperature of feed stream

723 K 823 K

Experimental dataa Model prediction AE Experimental dataa Model prediction AE


H2/C3H8O3 1.0 0.8 0.2 3.3 2.8 0.5
CO2/C3H8O3 0.47 0.45 0.02 1.6 1.3 0.3
CO/C3H8O3 0.069 0.075 0.006 0.28 0.39 0.11
CH4/C3H8O3 0.017 0.009 0.008 0.23 0.17 0.06
a
Cheng et al. (2011).

and Van’t Hoff equation were reported by Alipour-Dehkordi and Kha­


demi (2020).
Table 5
3.3. Base-scenario definition Verification between model prediction and industrial data presented by Abashar
(2000) for ammonia synthesis reactor.
To characterize and simulate the reactor’s behaviour, the geometry Parameter Industrial data Model prediction RE
and operating conditions for evaporation-, exothermic-, and
Outlet composition (mol%)
endothermic-side are assigned as a base scenario in Table 3. Also, to N2 17.80 17.72 0.45
investigate the role of parameters affecting the performance of the re­ H2 53.90 53.69 0.39
actors, H2/CO ratio, nitrogen, glycerol, and methane conversion as well NH3 19.10 19.34 1.25
as energy efficiency of the process are defined as assessment criteria by CH4 6.30 6.31 0.16
Ar 2.90 2.94 1.38
equations 43–47. Outlet temperature (K) 728 737.52 1.30
N2 conversion 0.3091 0.3138 1.52

7
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

Table 6 2,100 unequally-spaced points, and a numerical solution of all non-


Validation of model prediction with the experimental data for methane tri- linear and linear algebraic equations is found by the “fsolve” function
reformer. in MATLAB software (R2020a) based on the trust-region-dogleg method.
Feed temperature CO2 molar CH4
(K) fraction conversion 4. Results and discussion
1,023 Model prediction 13.2 98.7
Experimental 12.94 97.9 4.1. Model reliability analysis
dataa
RE 2.00% 0.81%
To validate the mathematical and kinetic models used for glycerol
1,123 Model prediction 10.9 99.9
Experimental 10.71 99.99 steam reforming, the molar ratio of products to reactant (glycerol) ob­
dataa tained by simulation at two feed temperatures of 723 and 823 K are
RE 1.77% 0.09% compared with the experimental data provided by Cheng et al. (2011)
a
Kang et al. (2007). under the geometry and operating conditions reported by them. Table 4
shows that the developed model can provide a convincing approxima­
tion of the experimental data with maximum absolute error (AE) of 0.5
for hydrogen to glycerol molar ratio at 823 K. Since the

Fig. 3. Variation of (a) glycerol and methane conversion, and (b) gas temperature in endothermic- and exothermic-side of reactor III under adiabatic and heat-
integrated conditions.

Fig. 4. Variation of dryness fraction, nitrogen conversion, and gas temperature in exothermic-side of reactor I and II.

8
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

Fig. 5. Effect of WHSV in endothermic-side of reactor III on (a) glycerol and methane conversion, ammonia production, H2/CO ratio, and (b) gas temperature in
exothermic-side of reactor II and III.

consumption/generation rate of steam was not presented in the kinetics good concordance between the experimental data and numerical
model and its concentration is predicted through the atomic balance, so modeling results under the operating and geometry conditions
some notable error is observed in some cases. mentioned by Kang et al. (2007), due to the maximum relative error
As indicated in Table 5, a comparison between the model prediction (RE) of 2% for the molar fraction of carbon dioxide at 1,023 K.
and industrial data presented by Abashar (2000), is done to confirm the Therefore, the given model can also be applied to simulate this heat-
accuracy of the kinetic and mathematical models for ammonia synthesis integrated intensified process. However, experimental proof-of-concept
reaction. A good agreement between these results is recognized with is required to validate the simulation results, get more certainty in the
relative error (RE) for outlet composition not exceeding 1.5%. model’s precision, and ensure the safe operability of this process.
Besides, the experimental data reported by Kang et al. (2007) at
equilibrium state is applied to validate the mathematical model and the
kinetic rate of methane tri-reforming reaction (see Table 6). There is a

Fig. 6. Effect of (a) O2/CH4, (b) O2/CH4, (c) H2O/glycerol, and (d) N2/H2 ratio on the performance of heat-integrated intensified process.

9
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

4.2. Simulation results reactors I and II. In the dimensionless length range of 0–0.1, the tem­
perature profile in both reactors rises to 695 K, which can be attributed
A comparison between heat-integrated and adiabatic conditions for to rapid ammonia synthesis reaction rate as a result of high reactants
reactor III, as shown in Fig. 3, is carried out in terms of glycerol and inlet concentration. After that, the molar fraction of nitrogen continu­
methane conversion as well as gas temperature in exothermic- and ously reduces, lowering the rate of nitrogen conversion, and conse­
endothermic-side. In the adiabatic reactor (AR), no heat is transferred quently the low heat generated through ammonia synthesis leads to a
from the solid wall of the exothermic-side into the endothermic-side. So, decrease in temperature. At the entrance of reactor I, a breakingpoint in
the heat required for glycerol steam reforming is taken from feed gas the temperature profile is observed due to the shifting from convective
sensible heat, and temperature prfile reduces quickly to reach 565 K (see to nucleate boiling region in the evaporation-side. Nitrogen conversion
Fig. 3(b)). This leads to a decrease in the reaction rate, and resulting in a at the end of reactors I and II reaches 27% and 26%, respectively. The
lower degree of glycerol conversion, i.e., 28% (see Fig. 3(a)). Instead, dryness fraction in reactor I increases up to 0.5, two-phase flow gradu­
the heat required for glycerol steam reforming in the heat-integrated ally changes to vapor-phase flow at the end of reactor II, where the
reactor is supplied by the other side; the gas temperature remains at a dryness fraction will be equal to 1.0.
higher level compared to the AR (see Fig. 3(b)), and consequently, the
glycerol conversion reaches 100% (see Fig. 3(a)). 4.3. Effect of WHSV
At the first half of the heat-integrated reactor and AR, the methane
conversion increases rapidly (see Fig. 3(a)), due to temperature rise, and The effect of weight hourly space velocity (WHSV) in glycerol steam
after that approaches equilibrium conversion. Although the heat- reforming reaction on the methane and glycerol conversion as well as
integrated reactor and AR almost have the same behaviour in terms of ammonia production and H2/CO ratio is indicated in Fig. 5(a). On the
methane conversion, the exothermic-side temperature in the heat- endothermic-side, the gas residence time decreases due to an increase in
integrated is much lower than that in the AR (see Fig. 3(b)), leading the WHSV, and this in turn, reduces the glycerol conversion. Simulta­
to a delay in Ni/MgO/CeZrO2/Al2O3 catalyst deactivation. At the neously, the endothermic-side heat transfer coefficient is raised with
entrance of the heat-integrated reactor, the temperature profile in the WHSV, leading to more heat transfer rate from the solid wall and
exothermic-side increases to 1,440 K and then decreases rapidly to consequently decreasing the temperature in the exothermic-side (see
1,140 K as a result of heat transferred from the solid wall; but in the AR, Fig. 5(b)). This causes a reduction in methane conversion. More
which acts as a traditional autothermal tri-reformer, a hot spot tem­ ammonia production and a higher H2/CO ratio are observed at higher
perature (i.e., 1,540 K) is first observed at the dimensionless length of WHSV due to producing more hydrogen.
0.2 and then gradually approaches 1,400 K due to fuel depletion.
Therefore, this heat-integrated process can also improve the conven­ 4.4. Effect of feed components molar ratio
tional autothermal tri-reformer in terms of catalyst lifetime by removing
heat from the solid wall. The effect of the O2/CH4 ratio on the performance of reactor III is
The comparison between the heat-integrated reactor and AR in­ illustrated in Fig. 6(a). Increasing the O2/CH4 ratio leads to more
dicates that the methane tri-reforming reaction can provide the heat available oxygen for the methane oxidation reaction, eventually
required to drive the glycerol steam reforming, and coupling of these increasing the gas temperature and methane conversion. This causes
two reactions can be feasible and beneficial for the following reasons: (i) more heat transfer to the endothermic-side and increases the gas tem­
saving much energy by transferring 69.23 kW heat from the exothermic- perature on this side, thereby increasing the glycerol conversion. It
to endothermic-side in order to drive the glycerol steam reforming, (ii) should be noted that O2/CH4 ratio above 0.64 has no considerable effect
achieving the highest conversion of methane and glycerol, and (iii) on both conversions. On the other hand, increasing the O2/CH4 ratio
keeping the temperature of methane tri-reforming reaction lower than decreases the H2/CO ratio due to more CO2 production, driving the
normal to delay in catalyst deactivation. reverse water-gas shift reaction and producing more CO as a result of the
Fig. 4 indicates the variations of the nitrogen conversion, dryness Le Chatelier’s principle.
fraction and gas temperature in the exothermic-side along the length of The maximum gas temperature on the endothermic-side increases
from 750 to 1,050 K as a result of increasing the O2/CH4 ratio from 0.4 to
0.8. Although the higher temperature in endothermic-side benefits the
glycerol conversion, coke deposition is more likely to occur at higher
temperatures during reaction time, thereby leading to catalyst deacti­
vation. In this regard, Cheng et al. (2011) reported that coke was
deposited on Ni/Al2O3 catalyst at 823 K after 4 h, even under excess
steam/glycerol ratios. High temperatures also increase the cost of
reactor construction, energy consumption, and operational costs.
Therefore, in this study, 823 K is selected as the maximum allowable
temperature for glycerol steam reforming reaction to prevent coke
deposition. At this temperature, the O2/CH4 ratio equal to 0.59 is found,
and a synthesis gas with H2/CO ratio of 2.6 is produced. The results
show that the O2/CH4 = 0.59 is the best ratio to reach methane and
glycerol conversion above 90%.
The O2/CH4 ratio, as indicated in Fig. 6(b), affects the performance
of reactors I and II. More ammonia is produced with O2/CH4 ratio due to
increasing the glycerol conversion (see Fig. 6(a)) and more hydrogen. In
addition, the ammonia production rate remains constant for O2/CH4 >
0.64 according to reaching the highest degree of glycerol conversion, i.
e., 100%. The outlet dryness fraction of reactor II raises with O2/CH4
ratio, and eventually approaches 1.0 during the transition from two-
phase flow to pure vapor as a result of a higher heat transfer coeffi­
cient inside the tubes. Also, no change in the outlet temperature of the
Fig. 7. Effect of splitting factor on the performance of the reactor I and II. evaporation-side of reactor II is observed and it remains at the boiling

10
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

Fig. 8. Effect of inlet gas temperature in (a) exothermic-side of reactor III, (b) exothermic-side of reactor III, (c) endothermic-side of reactor III, and (d) exothermic-
side of reactor I & II on the performance of heat-integrated intensified process.

Fig. 9. Effect of inlet gas pressure in (a) endothermic-side of reactor III, and (b) exothermic-side of reactor I and II on the performance of heat-integrated inten­
sified process.

point of glycerol. methane conversion. The ammonia production is diminished with


Fig. 6(c) shows the effect of steam/glycerol ratio, at a given WHSV (i. steam/glycerol ratio due to less available hydrogen. Besides, an increase
e., 20 1/h), on the performance of the heat-integrated intensified pro­ in the outlet dryness fraction of reactor II is found by decreasing the
cess. The glycerol conversion increases with steam/glycerol ratio up to H2O/glycerol ratio from 12 to 5 due to the higher heat transfer coeffi­
6.5 as a result of an increase in the reaction rate, and after that reaches cient in the exothermic-side of reactors I and II; and further reducing of
100% as a result of fuel depletion. The methane conversion also in­ H2O/glycerol ratio causes complete evaporation of glycerol liquid phase
creases because of lower heat transfer from the solid wall and higher (dryness fraction = 1.0). On the other hand, it has been reported in the
temperature levels on the exothermic-side. Although increasing the literature (Silva et al., 2015) that an abundant amount of steam has a
steam/glycerol ratio from 3 to 12 leads to an increase in the glycerol positive effect on the performance of the catalyst in glycerol steam
conversion up to 16%, it has a slight effect, lower than 4%, on the reforming due to the lack of carbon deposition. Nevertheless, it seems

11
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

Fig. 10. Effect of number of tubes in reactor I and II on the (a) performance of heat-integrated intensified process and (b) temperature profile in exothermic-side of
reactor II.

4.6. Effect of gas temperature


Table 7
Comparison between conventional and heat-integrated intensified process.
The changes in the performance of the heat-integrated intensified
Parameter Value process with inlet gas temperature in the exothermic-side of reactor III
Proposed Conventional are presented in Fig. 8(a) and (b). Rising the temperature in the
process processa exothermic-side from 800 to 1,200 K makes the methane tri-reforming
The main source of fuel (kg /tNH3 ) 1,872 kg 756.29 kg Natural reaction at a faster rate, and increases the methane conversion by
Glycerol gas
Steam (kg /tNH3 ) 7,316.5 2,499.35
32%. In addition, glycerol conversion also increases due to more heat
Nitrogen (kg /tNH3 ) 824 914.67
transfer from the solid wall. Results indicate that maximum glycerol
conversion, i.e., 100%, is obtained at 1,000 K, where the H2/CO ratio is
Ammonia production (kg) 1,000 1,000
equal to 2.52. The H2/CO ratio reduces with temperature due to
Water for cooling of ammonia reactor – 400.71
(kg /tNH3 ) increasing the rate of reverse water-gas shift reaction, thereby resulting
Emissions to air (kg /tNH3 ) 0.0 5,500.36 in hydrogen consumption and CO production. Although higher con­
Energy efficiency 31% 40% versions are achieved at higher inlet temperatures, the gas temperature
a
in the endothermic-side of reactor III exceeds the maximum allowable
Chisalita et al. (2020).
temperature (i.e., 823 K) at inlet gas temperatures above 1,000 K (see
Fig. 8(a)). Therefore, due to the achievement of methane and glycerol
that H2O/glycerol ≅ 5 is the best ratio under the base scenario, but using conversion above 90% and exhibiting a longer Ni/Al2O3 catalyst life­
a stochastic optimization algorithm can help to find the correct ratio. As time, 950–1,000 K is the best inlet gas temperature in the exothermic-
indicated in Fig. 6(c), the H2/CO ratio in the synthesis gas increases with side of reactor III.
the steam to glycerol ratio due to the higher conversion of methane. As shown in Fig. 8(b), more ammonia is produced with inlet gas
Fig. 6(d) presents the variations of the performance of the reactors I temperature, due to higher glycerol conversion (see Fig. 8(a)) and
and II versus N2/H2 ratio. More ammonia is produced with an increase in consequently more hydrogen production; as such, dryness fraction in­
the N2/H2 ratio up to 0.8 due to increasing the feed molar flow rate, and creases from 0.7 to 1 as a result of greater heat transfer rate from the
then decreases due to reducing hydrogen conversion. Adding more heat solid wall. At inlet gas temperature above 1,000 K, no change in
to the evaporation-side can either change the glycerol phase from liquid ammonia production and dryness fraction is observed. However, the
to vapor or raise the temperature of the pure vapor. At the end of the dryness fraction more than 98% is also taken at 950–1,000 K.
evaporation-side of reactor II, the dryness fraction increases as the N2/ Although increasing inlet gas temperature in the endothermic-side of
H2 ratio increases from 0.25 to 0.37 at the glycerol boiling point tem­ reactor III from 723 to 823 K has no considerable effect on glycerol
perature. As such, at N2/H2 ratios above 0.37, sensible heat leads to an conversion, the maximum temperature in this side increases from 815 to
increase in the vapor phase temperature from 625 to 648 K. Therefore, 868 K (see Fig. 8(c)). As shown in this figure, to prevent the Ni/Al2O3
N2/H2 ratio ≅ 0.8 is recommended for this process in terms of more catalyst deactivation, 723–745 K is the appropriate inlet gas tempera­
ammonia production and higher glycerol temperature in the evapora­ ture for this side.
tion-side. The variations of the performance of the reactors I and II versus inlet
gas temperature in the exothermic-side of these reactors are indicated in
4.5. Effect of splitting factor Fig. 8(d). Higher inlet gas temperature is favored for the production of
ammonia and glycerol in the vapor-phase. The glycerol temperature
Although the change in the splitting factor (fraction of total feed, exiting from the evaporation-side rises to 730 K by increasing the inlet
containing H2 and N2 entering reactor I) has no significant effect on the temperature of exothermic-side, due to further heat transfer from the
dryness fraction, it has a noticeable influence on the nitrogen conversion solid wall. As a result, it is evident that increasing the temperature of
and ammonia production (see Fig. 7). Increasing splitting factor leads to glycerol vapor can be remarkably effective in reducing the size of the
lower residence time in reactor I, thereby decreasing nitrogen conver­ heat-exchanger, which is located between reactors II and III.
sion in this reactor. On the other hand, this term increases nitrogen
conversion in reactor II due to higher residence time. Therefore, the
highest ammonia production is obtained at the intersection of the ni­ 4.7. Effect of operating pressure
trogen conversion profiles, in which the splitting factor is equal to 0.52.
In other words, equally dividing the feed stream into two reactors in The reactor wall can be constructed of high-pressure resistant ma­
parallel leads to maximum ammonia production. terials to allow the reactor to operate at different pressures on both sides.

12
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

Since the operating pressure of ammonia synthesis and glycerol steam with the lower heating value of 4,000 kJ/kg.
reforming reactions has a direct effect on ammonia production, the in­
fluence of this parameter on the performance of reactors is investigated 5. Conclusion
in Fig. 9. Increasing the gas pressure in the endothermic-side of reactor
III up to 5 atm increases the glycerol conversion, but a considerable This study was focused on a heat-integrated intensified process to
improvement in glycerol conversion is not found at higher pressures produce ammonia and syngas as valuable products. In this process,
more than 5 atm (see Fig. 9(a)). This trend can be attributed to this fact glycerol steam reforming continuously provided the hydrogen required
that further enhancement in the pressure, from a certain point, does not for ammonia synthesis, and the liquid glycerol was simultaneously
induce a dominant positive impact on the glycerol conversion with vaporized by heat generated from ammonia synthesis. Methane tri-
respected to the thermodynamic limitation. On the other hand, the reforming was also used as a heat source to drive the glycerol reform­
ammonia production is not significantly affected by changes in pressure. ing, and at the same time, the effluent gas exiting from the glycerol
The influence of inlet gas pressure in the exothermic-side of reactor I reforming side was consumed in the tri-reforming reaction. The
and II on ammonia production and dryness fraction (see Fig. 9(b)) shows behaviour of the reactors was revealed through a 1D heterogeneous plug
that the operating pressure is the most effective parameter during flow model. The backward finite difference approach was applied to
ammonia synthesis, so that rising the pressure from 150 to 300 atm discretize and solve the governing equations through the trust-region-
increases the ammonia production more than two folds, due to the Le dogleg approach.
Chatelier’s principle. In addition, the adequate dryness fraction (e.i., Based on detailed modeling and studying the impact of various pa­
1.0) can be achieved at pressures above 300 atm. rameters on the performance of this process, the acceptable range of
operating conditions for this system was found as follows: steam/glyc­
4.8. Effect of tubes number erol ratio ≅ 5.0, O2/CH4 ratio ≅ 0.6, N2/H2 ratio ≅ 0.8, splitting factor ≅
0.5, operating pressure of glycerol steam reforming ≅ 500 kPa, oper­
Fig. 10 illustrates the alterations of outlet temperature in ating pressure of ammonia synthesis ≥ 300 atm, 950 < inlet gas tem­
evaporation-side, dryness fraction, ammonia production, and tempera­ perature of methane tri-reforming < 1,000 K, 723 < inlet gas
ture profile of ammonia synthesis in response to the number of tubes in temperature of glycerol steam reforming <745 K, 670 < inlet gas tem­
reactor I and II. At constant feed molar flow rate, increasing the number perature of ammonia synthesis <750 K. Numerical results showed that
of tubes reduces the molar flow rate of reactants in each tube, leading to more than 90% of glycerol and methane, and more than 25% of nitrogen
a severe reduction in temperature profile (see Fig. 10(b)) and a decrease can be converted to syngas and ammonia. In addition, glycerol dryness
in ammonia production (see Fig. 10(a)). This can be attributed to slow fraction equal to 1.0, and a synthesis gas with H2/CO ratio above 2.0 was
rate of glycerol reforming reaction as a result of low glycerol inlet obtained, which is compatible with the Fischer-Tropsch and methanol
concentration. Although decreasing the number of tubes to less than 240 synthesis processes.
reduces the heat transfer surface area, more heat is transferred to the In general, the advantages of this heat-integrated intensified process
evaporation-side as a result of higher operating temperature on the compared to the conventional ammonia synthesis process are as follows:
exothermic-side. This concept increases the outlet temperature in the (1) reducing the process capital cost due to the removal of the desul­
evaporation-side of reactor II from 625 to 674 K and brings the dryness phuriser, steam reformer (furnace), water-gas shift converter, and
fraction to 1.0 (see Fig. 10(a)). condenser; and elimination of the inter-stage cooler for ammonia syn­
thesis reactor, (2) simultaneous production of ammonia and synthesis
4.9. Comparison between conventional and proposed process gas through glycerol valorization as a credible alternative to reforming
of natural gas, (3) providing the auto-thermality within the reactors by
Table 7 shows a comparison between ammonia production routes coupling glycerol steam reforming and methane tri-reforming, as well as
through the heat-integrated intensified process and conventional ammonia synthesis reaction and evaporation of liquid glycerol, (4)
methane steam reforming data reported by Chisalita et al. (2020) in mitigation of the environmental pollution due to no flue gas emissions
terms of greenhouse gas emissions, energy efficiency, the amount of fuel into the atmosphere compared to the conventional process (i.e.,
consumption, and cooling system for ammonia synthesis reactor. 5,500.36 kg/tNH3 ), and (5) use of greenhouse gases (CO2, CH4, and H2O)
Although the fuel consumption in the proposed process is higher than produced via glycerol steam reforming as the feedstock of methane tri-
the conventional, the main source of fuel for the proposed ammonia reforming.
production process is supplied from glycerol as a renewable carbon
resource which can reduce the price of biodiesel and make its production CRediT authorship contribution statement
process more economical. Correspondingly, more steam is also
consumed by the proposed process. In the conventional process, 400.71 Mohammad Hasan Khademi: Conceptualization, Project adminis­
kg of water per each tonne of ammonia is used to cool the ammonia tration, Supervision, Writing – original draft, Writing – review & editing.
synthesis reactor, while in the proposed system, no external cooling Mohammad Lotfi-Varnoosfaderani: Methodology, Software, Valida­
medium is required and the reactor is cooled by glycerol vaporization. tion, Investigation, Writing – original draft.
During the conventional process, large amounts of greenhouse gases
are released to the atmosphere through the furnace stack, but in the
proposed system, the greenhouse gases produced by glycerol steam Declaration of competing interest
reforming are reused in the methane tri-reforming reaction to supply the
heat of reaction. Thus, this leads to a reduction in gas emissions to air The authors declare that they have no known competing financial
from 5,500.36 kg per tonne of NH3 to zero. Although ammonia pro­ interests or personal relationships that could have appeared to influence
duction via the conventional method yields a higher energy efficiency (i. the work reported in this paper.
e., 40%) compared to the proposed method (i.e., 31%), it should be
noted that the energy of air and fuel for heating the furnace and elec­ Acknowledgement
tricity for gas compression in calculating energy efficiency of the con­
ventional process are not considered due to lack of information on this We thank Alireza Palizvan (Master’s student in chemical engineering
issue. In addition, the proposed process produces synthesis gas, as a at University of Isfahan) for assistance with methodology, and for
valuable product for methanol and Fischer-Tropsch synthesis process, comments that greatly improved the manuscript.

13
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

Nomenclature

a Catalyst specific surface area, m2 m− 3


Ac Cross-section area, m2
Acs Catalyst surface, m2 kgcat
− 1

Bo Boiling number
C Total concentration, mol m− 3
Co Convection number
Cp Molar heat capacity of the fluid, J mol− 1 K− 1
Dij Binary diffusion coefficient, m2 s− 1
Dim Diffusivity in a multicomponent gas mixture, m2 s− 1
dt Diameter of tube, m
dp Diameter of catalyst pelet, m
E Activation energy, kJ mol− 1
F Molar flow rate, mol s− 1
Frl Froude number with all flow in liquid
Ffl Fluid-dependent parameter
g gravitational acceleration, m s− 2
G Mass flux, kg m− 2 s− 1
h Heat transfer coefficient, W m− 2 K− 1
K Thermal conductivity, W m− 1 K− 1
Ka Equilibrium constant for ammonia synthesis reaction
1
Ki Adsorption equilibrium constant for methane tri-reforming (i = CH4, H2, CO, H2O, OC2 , CHC4 ), bar−
Kn Reaction equilibrium constant for methane tri-reforming (n = I, II, III)
kc Catalyst-fluid mass transfer coefficient, m s− 1
− 1 − 1
kj Rate constant for methane tri-reforming reaction (j = 1, 2, 3, 4a, 4b), mol kgcat s
M Molecular weight
N Total number of reactions
Ntube Number of tubes
n Total number of chemical components
P Pressure, Pa
Pr Prandtl number
Pi Partial pressure of species i, kPa
q Heat flux, W m− 2
R Universal gas constant, J mol− 1 K− 1
Re Reynolds number
− 1 − 1
r Rate of reaction, mol kgcat s
Sc Schmidt number
T Temperature, K
U Overall heat transfer coefficient, W m− 2 K− 1
us Superficial velocity, m s− 1
x Dryness fraction
y Molar fraction
z Axial coordinate, m

Greek letters
ΔHf Heat of formation, J mol− 1
ΔHr Heat of reaction, J mol− 1
μ Fluid viscosity, kg m− 1 s− 1
ε Bed porosity
η Effectiveness factor
λ Latent heat of vaporization, J mol− 1
ρ Fluid density, kg m− 3
ρb Catalytic bed density, kg m− 3
υ Stoichiometric coefficient
υc Molecular diffusion volume, cm3 mol− 1

Superscripts
0 Inlet condition
b Boiling point
endo Stands for the endothermic-side
exo Stands for the exothermic-side
g Gas-phase

14
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

s Catalyst-phase
TP Two-phase

Subscripts
i Chemical component
j Reaction No
k Side No. (1 = tube-side, 2 = shell-side)
l Liquid

References as nitrogen carriers. Nat. Energy 312 (3), 1067–1075. https://doi.org/10.1038/


s41560-018-0268-z.
Ghani, R., Iranshahi, D., 2019. Conceptual comparison of four configurations in the
Abashar, M.E.E., 2000. Application of heat interchange systems to enhance the
thermal coupling of ammonia synthesis and 2-butanol dehydrogenation. Appl.
performance of ammonia reactors. Chem. Eng. J. 78, 69–79. https://doi.org/
Therm. Eng. 154, 238–250. https://doi.org/10.1016/j.applthermaleng.2019.03.057.
10.1016/S1385-8947(99)00175-8.
Ghani, R., Boostani, F., Iranshahi, D., 2019. Analysis of the combined ammonia
Alipour-Dehkordi, A., Khademi, M.H., 2019. Use of a micro-porous membrane multi-
production and cyclohexane dehydrogenation by a novel bifunctional reactor.
tubular fixed-bed reactor for tri-reforming of methane to syngas: CO2, H2O or O2
Energy Fuels 33, 6717–6726. https://doi.org/10.1021/acs.energyfuels.9b01088.
side-feeding. Int. J. Hydrogen Energy 44, 32066–32079. https://doi.org/10.1016/j.
Gilbert, P., Alexander, S., Thornley, P., Brammer, J., 2014. Assessing economically viable
ijhydene.2019.10.097.
carbon reductions for the production of ammonia from biomass gasification.
Alipour-Dehkordi, A., Khademi, M.H., 2020. O2, H2O or CO2 side-feeding policy in
J. Clean. Prod. 64, 581–589. https://doi.org/10.1016/J.JCLEPRO.2013.09.011.
methane tri-reforming reactor: the role of influencing parameters. Int. J. Hydrogen
Gillespie, L.J., Beattie, J.A., 1930. The thermodynamic treatment of chemical equilibria
Energy 45, 15239–15253. https://doi.org/10.1016/j.ijhydene.2020.03.239.
in systems composed of real gases. I. An approximate equation for the mass action
Anderson, K., Bows, A., Mander, S., 2008. From long-term targets to cumulative emission
function applied to the existing data on the haber equilibrium. Phys. Rev. 36,
pathways: reframing UK climate policy. Energy Pol. 36, 3714–3722.
743–753. https://doi.org/10.1103/PhysRev.36.743.
Arab Aboosadi, Z., Jahanmiri, A.H., Rahimpour, M.R., 2011. Optimization of tri-reformer
Global Status Report - Global CCS Institute [WWW Document], n.d. URL. https://www.gl
reactor to produce synthesis gas for methanol production using differential evolution
obalccsinstitute.com/resources/global-status-report/, 9.7.21.
(DE) method. Appl. Energy 88, 2691–2701. https://doi.org/10.1016/j.
Iliuta, I., Iliuta, M.C., Fongarland, P., Larachi, F., 2012. Integrated aqueous-phase
apenergy.2011.02.017.
glycerol reforming to dimethyl ether synthesis-A novel allothermal dual bed
Arora, P., Hoadley, A.F.A., Mahajani, S.M., Ganesh, A., 2016. Small-scale Ammonia
membrane reactor concept. Chem. Eng. J. 187, 311–327. https://doi.org/10.1016/j.
production from biomass: a techno-enviro-economic perspective. Ind. Eng. Chem.
cej.2012.02.001.
Res. 55, 6422–6434. https://doi.org/10.1021/ACS.IECR.5B04937.
Kandlikar, S.G., 1990. A general correlation for saturated two-phase flow boiling heat
Azarhoosh, M.J., Farivar, F., Ale Ebrahim, H., 2014. Simulation and optimization of a
transfer inside horizontal and vertical tubes. J. Heat Tran. 112, 219–228. https://
horizontal ammonia synthesis reactor using genetic algorithm. RSC Adv. 4,
doi.org/10.1115/1.2910348.
13419–13429. https://doi.org/10.1039/c3ra45410j.
Kang, J.S., Kim, D.H., Lee, S.D., Hong, S.I., Moon, D.J., 2007. Nickel-based tri-reforming
Bicer, Y., Dincer, I., 2018. Life cycle assessment of ammonia utilization in city
catalyst for the production of synthesis gas. Appl. Catal. Gen. 332, 153–158. https://
transportation and power generation. J. Clean. Prod. 170, 1594–1601. https://doi.
doi.org/10.1016/j.apcata.2007.08.017.
org/10.1016/J.JCLEPRO.2017.09.243.
Khademi, M., Alipour-Dehkordi, A., 2021. Modeling and optimization of methane tri-
Bicer, Y., Dincer, I., Zamfirescu, C., Vezina, G., Raso, F., 2016. Comparative life cycle
reforming reactor under different side-feeding strategies with the aim of maximizing
assessment of various ammonia production methods. J. Clean. Prod. 135,
hydrogen yield. Nashrieh Shimi va Mohandesi Shimi Iran.
1379–1395. https://doi.org/10.1016/J.JCLEPRO.2016.07.023.
Khademi, M.H., Sabbaghi, R.S., 2017. Comparison between three types of ammonia
Carvalho, E.P., Borges, C., Andrade, D., Yuan, J.Y., Ravagnani, M.A.S.S., 2014. Modeling
synthesis reactor configurations in terms of cooling methods. Chem. Eng. Res. Des.
and optimization of an ammonia reactor using a penalty-like method. Appl. Math.
128, 306–317. https://doi.org/10.1016/j.cherd.2017.10.021.
Comput. https://doi.org/10.1016/j.amc.2014.03.099.
Khademi, M.H., Alipour-Dehkordi, A., Tabesh, M., 2021. Optimal design of methane tri-
Chein, R.Y., Wang, C.Y., Yu, C.T., 2017. Parametric study on catalytic tri-reforming of
reforming reactor to produce proper syngas for Fischer-Tropsch and methanol
methane for syngas production. Energy 118, 1–17. https://doi.org/10.1016/j.
synthesis processes: a comparative analysis between different side-feeding strategies.
energy.2016.11.147.
Int. J. Hydrogen Energy 46, 14441–14454. https://doi.org/10.1016/j.
Cheng, C.K., Foo, S.Y., Adesina, A.A., 2011. Steam reforming of glycerol over Ni/Al2O3
ijhydene.2021.01.215.
catalyst. Catal. Today 178, 25–33. https://doi.org/10.1016/j.cattod.2011.07.011.
Khazayialiabad, A., Iranshahi, D., Ebrahimian, S., 2021. A conceptual evaluation of a
Chisalita, D.A., Petrescu, L., Cormos, C.C., 2020. Environmental evaluation of european
new multifunctional reactor containing glycerol steam reforming and nitrobenzene
ammonia production considering various hydrogen supply chains. Renew. Sustain.
hydrogenation. Chem. Eng. Process. - Process Intensif. 164, 108405. https://doi.org/
Energy Rev. 130, 109964. https://doi.org/10.1016/J.RSER.2020.109964.
10.1016/j.cep.2021.108405.
Coker, A.K., 2007. Ludwig’s Applied Process Design for Chemical and Petrochemical
Khodabandehloo, M., Larimi, A., Khorasheh, F., 2020. Comparative process modeling
Plants, Ludwig’s Applied Process Design for Chemical and Petrochemical Plants.
and techno-economic evaluation of renewable hydrogen production by glycerol
Elsevier Inc. https://doi.org/10.1016/B978-0-7506-7766-0.X5000-3.
reforming in aqueous and gaseous phases. Energy Convers. Manag. 225, 113483.
Cussler, E.L., Cussler, E.L., 2009. Diffusion: Mass Transfer in Fluid Systems. Cambridge
https://doi.org/10.1016/j.enconman.2020.113483.
university press.
Leva, M., 1950. Packed-tube heat transfer. Ind. Eng. Chem. 42, 2498–2501. https://doi.
Dashti, A., Khorsand, K., Marvast, M.A., Kakavand, M., 2006. Modeling and simulation of
org/10.1021/ie50492a031.
ammonia synthesis reactor. Pet. Coal 48, 15–23.
Leva, M., Weintraub, M., Grummer, M., Clark, E.L., 1948. Cooling of gases through
De Groote, A.M., Froment, G.F., 1996. Simulation of the catalytic partial oxidation of
packed tubes. Ind. Eng. Chem. 40, 747–752. https://doi.org/10.1021/ie50460a042.
methane to synthesis gas. Appl. Catal. Gen. 138, 245–264. https://doi.org/10.1016/
Ma, F., Hanna, M.A., 1999. Biodiesel production: a review. Bioresour. Technol. 70, 1–15.
0926-860X(95)00299-5.
https://doi.org/10.1016/S0960-8524(99)00025-5.
De Smet, C.R.H., De Croon, M.H.J.M., Berger, R.J., Marin, G.B., Schouten, J.C., 2001.
Macedo, M.S., Soria, M.A., Madeira, L.M., 2019. Glycerol steam reforming for hydrogen
Design of adiabatic fixed-bed reactors for the partial oxidation of methane to
production: traditional versus membrane reactor. Int. J. Hydrogen Energy 44,
synthesis gas. Application to production of methanol and hydrogen-for-fuel-cells.
24719–24732. https://doi.org/10.1016/j.ijhydene.2019.07.046.
Chem. Eng. Sci. 56, 4849–4861. https://doi.org/10.1016/S0009-2509(01)00130-0.
McCabe, W.L., Smith, J.C., Harriott, P., 1993. Unit Operations of Chemical Engineering.
Dincer, I., Bicer, Y., 2018. 2.1 ammonia. Compr. Energy Syst. 2– 5, 1–39. https://doi.
McGraw-hill, New York.
org/10.1016/B978-0-12-809597-3.00201-7.
Medford, A.J., Hatzell, M.C., 2017. Photon-driven nitrogen fixation: current progress,
Dyson, D.C., Simon, J.M., 1968. A kinetic expression with diffusion correction for
thermodynamic considerations, and future outlook. ACS Catal. 7, 2624–2643.
ammonia synthesis on industrial catalyst. Ind. Eng. Chem. Fundam. 7, 605–610.
https://doi.org/10.1021/ACSCATAL.7B00439.
https://doi.org/10.1021/i160028a013.
Mikulčić, H., Skov, I.R., Dominković, D.F., Wan Alwi, S.R., Manan, Z.A., Tan, R.,
Fairbanks, D.F., Wilke, C.R., 1950. Diffusion coefficients in multicomponent gas
Duić, N., Hidayah Mohamad, S.N., Wang, X., 2019. Flexible Carbon Capture and
mixtures. Ind. Eng. Chem. 42, 471–475. https://doi.org/10.1021/ie50483a022.
Utilization technologies in future energy systems and the utilization pathways of
Farniaei, M., Abbasi, M., Rahnama, H., Rahimpour, M.R., Shariati, A., 2014. Syngas
captured CO2. Renew. Sustain. Energy Rev. 114 https://doi.org/10.1016/J.
production in a novel methane dry reformer by utilizing of tri-reforming process for
RSER.2019.109338.
energy supplying: modeling and simulation. J. Nat. Gas Sci. Eng. 20, 132–146.
Nikzad, A., Iranshahi, D., 2021. Analysis of integrated system for ammonia synthesis and
https://doi.org/10.1016/j.jngse.2014.06.010.
methyl formate production in the thermally coupled reactor. Chem. Eng. Process. -
Gálvez, M.E., Frei, A., Halmann, M., Steinfeld, A., 2007. Ammonia production via a two-
Process Intensif. 166, 108418. https://doi.org/10.1016/j.cep.2021.108418.
step Al2O3/AlN thermochemical cycle. 2. Kinetic analysis. Ind. Eng. Chem. Res. 46,
Okoye, P.U., Abdullah, A.Z., Hameed, B.H., 2017. A review on recent developments and
2047–2053. https://doi.org/10.1021/IE061551M.
progress in the kinetics and deactivation of catalytic acetylation of glycerol—a
Gao, W., Guo, J., Wang, P., Wang, Q., Chang, F., Pei, Q., Zhang, W., Liu, L., Chen, P.,
byproduct of biodiesel. Renew. Sustain. Energy Rev. https://doi.org/10.1016/j.
2018. Production of ammonia via a chemical looping process based on metal imides
rser.2017.02.017.

15
M.H. Khademi and M. Lotfi-Varnoosfaderani Journal of Cleaner Production 324 (2021) 129241

Patcharavorachot, Y., Chatrattanawet, N., Arpornwichanop, A., Assabumrungrat, S., coupling aqueous phase glycerol reforming and methanol synthesis. Catal. Today.
2019. Optimization of hydrogen production from three reforming approaches of https://doi.org/10.1016/j.cattod.2020.03.043.
glycerol via using supercritical water with in situ CO2 separation. Int. J. Hydrogen Trimm, D.L., Lam, C.W., 1980. The combustion of methane on platinum-alumina fibre
Energy 44, 2128–2140. https://doi.org/10.1016/j.ijhydene.2018.07.053. catalysts-I. Kinetics and mechanism. Chem. Eng. Sci. 35, 1405–1413. https://doi.
Poling, B.E., Prausnitz, J.M., O’connell, J.P., 2001. Properties of Gases and Liquids. org/10.1016/0009-2509(80)85134-7.
McGraw-Hill Education. Unlu, D., Hilmioglu, N.D., 2020. Application of aspen plus to renewable hydrogen
Rahimpour, M.R., Arab Aboosadi, Z., Jahanmiri, A.H., 2012. Synthesis gas production in production from glycerol by steam reforming. Int. J. Hydrogen Energy 45,
a novel hydrogen and oxygen perm-selective membranes tri-reformer for methanol 3509–3515. https://doi.org/10.1016/j.ijhydene.2019.02.106.
production. J. Nat. Gas Sci. Eng. 9, 149–159. https://doi.org/10.1016/j. Wang, X., Wang, N., Li, M., Li, S., Wang, S., Ma, X., 2010. Hydrogen production by
jngse.2012.06.007. glycerol steam reforming with in situ hydrogen separation: a thermodynamic
Rahnama, H., Farniaei, M., Abbasi, M., Rahimpour, M.R., 2014. Modeling of synthesis investigation. Int. J. Hydrogen Energy 35, 10252–10256. https://doi.org/10.1016/j.
gas and hydrogen production in a thermally coupling of steam and tri-reforming of ijhydene.2010.07.140.
methane with membranes. J. Ind. Eng. Chem. 20, 1779–1792. https://doi.org/ Weigand, B., 2004. Analytical Methods for Heat Transfer and Fluid Flow Problems,
10.1016/j.jiec.2013.08.032. Analytical Methods for Heat Transfer and Fluid Flow Problems. Springer Berlin
Sadeghi, M.T., Kavianiboroujeni, A., 2008. The optimization of an ammonia synthesis Heidelberg. https://doi.org/10.1007/978-3-540-68466-4.
reactor using genetic algorithm. Int. J. Chem. React. Eng. 6 https://doi.org/ Weng, Q., Toan, S., Ai, R., Sun, Zhao, Sun, Zhiqiang, 2021. Ammonia production from
10.2202/1542-6580.1789. biomass via a chemical looping–based hybrid system. J. Clean. Prod. 289, 125749.
Saidi, M., Moradi, P., 2020. Conversion of biodiesel synthesis waste to hydrogen in https://doi.org/10.1016/J.JCLEPRO.2020.125749.
membrane reactor: theoretical study of glycerol steam reforming. Int. J. Hydrogen Wilke, C.R., 1950. A viscosity equation for gas mixtures. J. Chem. Phys. 18, 517–519.
Energy 45, 8715–8726. https://doi.org/10.1016/j.ijhydene.2020.01.064. https://doi.org/10.1063/1.1747673.
Sarafraz, M.M., Christo, F.C., 2021. Sustainable three-stage chemical looping ammonia Xu, J., Froment, G.F., 1989. Methane steam reforming, methanation and water-gas shift:
production (3CLAP) process. Energy Convers. Manag. 229, 113735. https://doi.org/ I. Intrinsic kinetics. AIChE J. 35, 88–96. https://doi.org/10.1002/aic.690350109.
10.1016/J.ENCONMAN.2020.113735. Xu, D., Lv, L., Ren, X., Ren, J., Dong, L., 2018. Route selection for low-carbon ammonia
Shakeri, M., Iranshahi, D., Naderifar, A., 2019. A conceptual investigation for the production: a sustainability prioritization framework based-on the combined
simultaneous production of gasoline and ammonia in thermally coupled reactors. weights and projection ranking by similarity to referencing vector method. J. Clean.
Chem. Eng. Process.- Process Intensif. 138, 15–26. https://doi.org/10.1016/j. Prod. 193, 263–276. https://doi.org/10.1016/J.JCLEPRO.2018.05.054.
cep.2019.02.009. Yang, X., Wang, S., Liu, H., Liu, G., He, Y., 2020. Numerical studies of sorption-enhanced
Shan, N., Chikan, V., Pfromm, P., Liu, B., 2018. Fe and Ni dopants facilitating ammonia glycerol steam reforming in a fluidized bed membrane reactor at low temperature.
synthesis on Mn4N and mechanistic insights from first-principles methods. J. Phys. Int. J. Hydrogen Energy 45, 8346–8356. https://doi.org/10.1016/j.
Chem. C 122, 6109–6116. https://doi.org/10.1021/ACS.JPCC.7B12569. ijhydene.2020.01.110.
Silva, J.M., Soria, M.A., Madeira, L.M., 2015. Challenges and strategies for optimization Yaws, C.L., 1999. Chemical Properties Handbook: Physical, Thermodynamic,
of glycerol steam reforming process. Renew. Sustain. Energy Rev. https://doi.org/ Environmental, Transport, Safety, and Health Related Properties for Organic and
10.1016/j.rser.2014.10.084. Inorganic Chemicals/. McGraw-Hill, New York.
Soloveichik, G., 2019. Electrochemical synthesis of ammonia as a potential alternative to Zhang, H., Wang, L., Van Herle, J., Marechal, F., Desideri, U., 2019. Techno-economic
the Haber–Bosch process. Nat. Catal. 25 2, 377–380. https://doi.org/10.1038/ comparison of green ammonia production processes. Appl. Energy, 114135. https://
s41929-019-0280-0. doi.org/10.1016/J.APENERGY.2019.114135.
Thirabunjongcharoen, S., Bumroongsakulsawat, P., Praserthdam, P., Charojrochkul, S., Zhu, X., Imtiaz, Q., Donat, F., Müller, C.R., Li, F., 2020. Chemical looping beyond
Assabumrungrat, S., Kim-Lohsoontorn, P., 2020. Thermally double coupled reactor combustion – a perspective. Energy Environ. Sci. 13, 772–804. https://doi.org/
10.1039/C9EE03793D.

16

You might also like