You are on page 1of 13

Journal of Environmental Chemical Engineering 7 (2019) 103018

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Utilization of biogas from different substrates for SOFC feed via steam T
reforming: Thermodynamic and exergy analyses

Kantilal Chouhan, Shishir Sinha, Shashi Kumar, Surendra Kumar
Department of Chemical Engineering, Indian Institute of Technology Roorkee, Roorkee, 247667, Uttarakhand, India

A R T I C LE I N FO A B S T R A C T

Keywords: Solid oxide fuel cell utilizes CO and H2 as fuel for electrical power generation. These are also tolerant to various
Biogas contaminants in fuel. With this view, detailed thermodynamic equilibrium analysis of steam reforming of biogas
Steam reforming of varying compositions (CH4 = 45–80%) have been carried out at various temperature (573–1273 K) and steam
Syngas production to CH4 molar ratio (0–2) at 1 atm pressure. CH4 conversion of ≥ 99% and complete elimination of carbon
Solid oxide fuel cells
formation have been obtained at 1 atm, 1073 K and S/C = 1.6 for all biogas compositions studied. Fortunately,
Energetic analysis
Exergetic analysis
under these conditions, combined moles of H2 and CO per mole of CH4 remain constant for all biogas compo-
sitions. Besides, thermal and exergy efficiencies of reforming process have also been computed and these are
(92.39–91.46%) and (75.57-69.16%) respectively for CH4/CO2 from 45/55 to 80/20. Energy requirement per
mole of biogas (108.99–179.25 kJ) has been obtained. It is our view that the product gas of steam reforming of
biogas produced from various feedstocks, can be utilized as a fuel for Solid Oxide Fuel Cells and can produce
electrical power between 157.5 and 280 kW for methane percent between 45 and 80 respectively.

1. Introduction Biogas reforming is an attractive and promising method for hy-


drogen and syngas production which can be used as fuel in solid oxide
Energy security is a major challenge of sustainable development fuel cells (SOFCs) as well as for Fischer Tropsch synthesis [1–3,8]. The
particularly in view of continuously increasing global population and technologies which can be used for biogas reforming are steam re-
energy demands. Presently, our energy systems are based on the fossil forming, partial oxidation, dry and autothermal reforming [8,14]. Main
fuel reserves which are depleting and unsustainable, but are responsible product of reforming process is the mixture of CO and H2 i.e. syngas.
for the emissions of greenhouse gases and pollutants leading to global Reforming process with or without separation unit can be coupled with
warming, climate change and environment pollution [1–3]. It is ulti- fuel cell for electrical power generation at small to medium scales,
mate way to exploit and utilize renewable energy sources for the sus- depending on the type of fuel cell used. In this way, it could address
tainable development. The guaranteed supply of renewable fuel at problem of electrical power unavailability/shortage in remote locations
steady rate cannot be ensured. Biomass feedstock for biogas digester where continuous feedstock for biogas digester is easily available.
can be guaranteed to some extent because of continuous availability of Most of research have been carried out on dry reforming of biogas
its feedstocks. Landfills are also the source of biogas. Biogas is a pro- containing CH4/CO2 = 50/50 and 60/40 [14–27]. Dry biogas re-
mising renewable energy source that is typically obtained by anaerobic forming (DBR) has a severe disadvantage of quick deactivation of cat-
fermentation of a wide range of biodegradable organic fractions of alyst due to coke/carbon formation which can be inhibited by adding
agro-industrial waste, municipal solid waste, animal dung, sewage small amount of steam.
sludge etc. The major components of biogas are methane and carbon Steam biogas reforming (SBR) is the combination of dry reforming
dioxide; both are principal greenhouse gases (GHGs) which leads to (DR) and steam reforming (SR) of methane (CH4). It seems to be a
global climate change. It also contains trace amounts of H2S, H2O, N2, promising option for the production of syngas for fuel cell applications.
O2 and Siloxanes [4–6]. Biogas composition depends on its feedstocks. It requires an efficient catalyst which favors both DR and SR reactions.
Its compositions for various feedstocks are shown in Table 1a. It con- Both reactions are highly endothermic in nature though DR being more
tains 45–80% CH4 and 55–20% CO2 depending on its source/feedstocks endothermic than SR. Many research works on SBR can be found in
for biogas digester [7–13]. literature. But these are based on model biogas having CH4/CO2 = 50/


Corresponding author.
E-mail address: skumar.iitroorkee@gmail.com (S. Kumar).

https://doi.org/10.1016/j.jece.2019.103018
Received 28 November 2018; Received in revised form 25 February 2019; Accepted 9 March 2019
Available online 11 March 2019
2213-3437/ © 2019 Elsevier Ltd. All rights reserved.
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

Nomenclature [–]
N2 Number of chemical components present in outlet stream
aij Number of atoms of the jth element present in each che- [–]
mical component i [–] NS Number of chemical reaction components present in the
Aj Total number of atoms of kth element present in the feed stream
in
[–] nCH 4
Moles of methane in inlet stream [mol]
out
cP, i Heat capacity of ith component [kJ/mol.K] nCO Moles of carbon monoxide in outlet stream [mol]
out
Ex chem Total Chemical exergy [kJ/mol] n H2 Moles of hydrogen in outlet stream [mol]
Ex chem, i Standard molar chemical exergy of ith component in gas ni Number of moles of component i [–]
phase [kJ/mol] niin Moles of component i in inlet stream [mol]
Ex dest Exergy destruction [kJ/mol] niout Moles of component i in outlet stream [mol]
in
Ex mix Exergy of mixing [kJ/mol] nsteam Moles of steam in inlet stream [mol]
Ex phy Physical exergy [kJ/mol] P Pressure [atm]
ExT Total exergy [kJ/mol] P0 Standard pressure [atm]
ExT , in Total exergy of inlet stream [kJ/mol] QR Net change in total enthalpies of inlet and outlet streams
ExT , out Total exergy of outlet stream [kJ/mol] [kJ/mol]
Gfi0 Standard Gibbs free energy of formation of ith component R Universal gas constant [=8.314 kJ/kmol.K]
[kJ/mol] Sgen Entropy generated [kJ/mol.K]
GT Total Gibbs free energy of the system [kJ] ΔS Change in entropy [kJ/mol.K]
HTin, i Enthalpy of ith component in the inlet stream [kJ/mol] T Temperature [K]
HTout
,i Enthalpy of ith component in the outlet stream [kJ/mol] yi Mole fraction of component i in the gaseous stream [–]
HTin Total enthalpy of inlet stream at temperature T [kJ/mol]
HTout Total enthalpy of outlet stream at temperature T [kJ/mol] Abbreviation
H298, i Molar enthalpy of ith component at a temperature of 298 K
[kJ/mol] BG Biogas
ΔHL Latent heat of steam in the feed [ = 40.69 kJ/mol] DR Dry reforming
ΔH Change in enthalpy [kJ/mol] GFE Gibbs free energy
j Atomic element [–] SBR Steam reforming of biogas
LHVCH4 Lower heating value of methane [=802.26 kJ/mol] SR Steam reforming
LHVCO Lower heating value of carbon monoxide [ = 282.97 kJ/
mol] Greek Symbols
LHVH2 Lower heating value of H2 [ = 241.80 kJ/mol]
k Total number of atomic elements comprising the reacting ϕi Fugacity coefficient of component i [–]
system [–] μi Chemical potential of component i [kJ/mol]
N Total number of chemical components in the reacting ηEx Exergy efficiency [–]
system [–] ηTh Thermal efficiency [–]
N1 Number of chemical components present in inlet stream

50 and 60/40 compositions only. Nickel is widely used catalyst for susceptible to deactivation due to carbon formation and sintering at
methane steam reforming on industrial scale. But in dry reforming, high temperature. A list of catalysts used for steam reforming of biogas
carbon deposition occurs which leads to its deactivation. Carbon/coke (SBR) is presented in Table 1b.
formation can be inhibited by using different catalysts, preparation Galvagno et al. [39] carried out thermodynamic equilibrium ana-
techniques, supports and promoters as reported in the literature lysis of SBR having CH4/CO2 equal to 60/40. Syngas composition and
[14,20–23]. thermal efficiency of reforming process were reported. Roy et al. [37]
Experimental studies have been carried out intensively for use of Ni studied SBR process and reported the syngas production index in the
based catalysts with different supports namely, ɣ-Al2O3, MgO, Al2O3,
La2O3-CeO2-ZrO2 [7,28–31]. Bimetallic Ni catalysts i.e. Ni-Pt, V-Ni, Ni-
Cr with different supports, have also been investigated [32–34]. Ex- Table 1b
pensive noble catalysts i.e. Rh, Ru, and Pd-Rh [31,35–38] are found to List of catalysts used in steam reforming of biogas.
be active and stable while Ni based catalysts are inexpensive and S.No. Catalysts Authors Reference

1. 2 wt% Ru/Al2O3 Sato et al. (2010) [44]


Table 1a 2. 18 wt% Ni/Al2O3 Izquierdo et al. [45]
(2012)
Composition of biogas obtained from different feedstocks.
3. 15 wt% Ni/γ-Al2O3 Appari et al. (2014) [28]
Source/ Biogas composition (% by volume) References 4. 4 wt % Rh/La-Al2O3 Ahmed et al. (2015) [35]
Substrate 5. Porous Ni-Cr plate (90 wt% Ni and Bui et al. (2015) [34]
CH4 CO2 N2 H2S O2 NH3 10 wt% Cr)
(ppm) 6. 1.31 wt% Pd-Rh/CeZrO2-Al2O3 Roy et al. (2015) [37]
7. 3 wt% V-Ni/Al2O3 Kowalik et al. [33]
Sewage sludge 55-65 35-45 <1 0-62.9 < 1 < 1 [4,6,7] (2015)
Agricultural 45-75 25-55 0-5 10-180 0.01-2 < 1 [4,9,10] 8. Ni/BaTi0.8Sn0.2O3 Miyake et al. (2015) [30]
and organic 9. 7.8 wt% Ni/CeO2 Italiano et al. [46]
waste (2015)
Landfill gas 45-62 24-40 1-17 15-427 1-2.6 – [2,4,11,15] 10. 10 wt% Ni/La2O3–CeO2-ZrO2 Angeli et al. (2016) [31]
Slaughter house 78.5 18-20 1-5 – – – [12] 11. Ni-Pt/CeO2 Turchetti et al. [32]
waste (3 wt% Pt and 10 wt % Ni) (2016)

2
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

N
range of 0.78–1.25 for Pd-Rh catalyst supported on CeZrO2-Al2O3. Peng
et al. [40] reported thermal efficiency values 90.7–93.8% for steam ∑ ni aij = Aj ; j = 1,2, 3……k
i=1 (2)
reforming of methane. Cruz et al. [41] studied exergy analysis of biogas
dry reforming and found overall exergy efficiency of 55%. Roy et al. Solid carbon formation over the catalyst surface has been observed
[38] reported overall thermal efficiency of 42% for a SOFC integrated in many experimental studies of steam reforming of biogas and me-
steam biogas reforming process. Kumar et al. [42] carried out ther- thane [22,28,30,33,45,48]. Since carbon has negligible vapor pressure
modynamic analysis of steam reforming of butanol-ethanol mixture. at operating conditions (T and P), only solid carbon is taken into con-
Thermal and exergy efficiencies of reforming process were reported. sideration. The carbon is determined from element balance equations.
Ashrafi et al. [43] calculated the equilibrium concentrations/com- There are many software packages (Chemcad, HSC Chemistry,
positions for steam reforming of model biogas having 60% CH4 and DETCHEM, Aspen Plus etc.) having various reactor modules based on
40% CO2 using Gibbs free energy (GFE) minimization method namely, GFE minimization, for the calculation of equilibrium compositions of
IPSEpro steady state program with S/C ratio varying from 2 to 5. Roy components. RGibbs reactor module of Aspen plus v8.4 has been used
et al. [38], and Galvagno et al. [39] carried out thermodynamic equi- to obtain equilibrium compositions (CH4, CO2, H2O, CO, H2 and C) in
librium analysis for syngas production via steam reforming based on this work. Unlike, ideal gases, fugacity coefficient (ϕi) of real gases is
Gibbs free energy minimization method using Aspen Plus simulation. not equal to unity and can be determined by using appropriate equation
Peng-Robinson equation of state was employed to calculate thermo- of state like Peng-Robinson (PR), Redlich-Kwong (RK), Soave mod-
dynamic and physical properties of reaction components. Galvagno ification of Redlich-Kwong (SRK). Peng–Robinson (PR) is generally
et al. [39] determined syngas compositions via steam reforming (S/ found suitable for estimating thermodynamic properties of hydro-
C = 1.2 and 2.5) and partial oxidation (O2/C = 0.12 and 0.25) using a carbons and light gases [39,42,49,50]. Therefore, Peng-Robinson (PR)
fixed biogas composition (60% CH4 and 40% CO2). Roy et al. [38] in- equation of state has been selected as property method in present study.
vestigated steam reforming of biogas (CH4/CO2 = 40/60, 50/50 and For simulation by RGibbs reactor module of Aspen Plus, one has to
60/40) and S/C ratio (1 to 2 with 0.25 interval). However, the steam specify (i) moles of CH4, CO2 and H2O (ii) operating temperature and
reforming of biogas with wide range of compositions, normally ob- pressure, and (iii) Thermodynamic property method. After simulation
tained by anaerobic fermentation of different substrates, have not yet by RGibbs reactor, moles of CH4, CO2, H2O, CO, H2 and C are obtained.
been done and so it is not available in the literature. It is to be mentioned here that the product of reforming are to be
In the present work the thermodynamic equilibrium analysis of specified by the user prior to simulation based on the experimental
steam reforming of biogas for a wide range of compositions (CH4/ work and trial simulations [42].
CO2 = 45/55 to 80/20) have been carried out. Main objective was to
determine the single end use of the product gas (H2 and CO) as a fuel for 2.2. Thermal efficiency and energy analysis
SOFC. For this purpose steam to methane ratio (S/C) has been varied
from 0 to 2 for various temperatures between 573 and 1273 K at at- Thermal efficiency is a measure of the performance of energy uti-
mospheric pressure. Besides, energy requirement, thermal and exergy lization by the system. It is based on the first law of thermodynamics
efficiencies of the process and consequently power generated by SOFC and indicates how efficiently the reformer converts the energy supplied
have also been calculated. (input) into useful work. The thermal efficiency is defined as the ratio of
the net work output to the total energy input [51]. In the case of the
biogas steam reforming, thermal efficiency can be written as
2. Methodology
n Hout
2
out
LHVH2 + nCO LHVCO
ηThermal (%) = in in
× 100
2.1. Thermodynamic equilibrium analysis nCH 4
LHVCH4 + QR + nsteam ΔHL (3)

QR = HTout − HTin (4)


Thermodynamic equilibrium analysis of any chemical reaction
system can be carried out using two ways, one being stoichiometric N1 N1 T
⎡ ⎤
method and other is non stoichiometric method. Stoichiometric method HTin = ∑ niin HTin,i = ∑ niin ⎢H298,i + ∫ cP,i dT⎥
requires knowledge of all independent chemical reactions and their i=1 i=1 ⎣ 298 ⎦ (5)
equilibrium constant. For each independent reaction, one equation re-
N2 N2 T
lating mole fraction of components and equilibrium constant can be ⎡ ⎤
developed. A set of equations corresponding to independent reactions
HTout = ∑ niout HTout,i = ∑ niout ⎢H298,i + ∫ cP,i dT⎥
i=1 i=1 ⎣ 298 ⎦ (6)
can be solved simultaneously to determine equilibrium mole fractions
of components [47]. However, these equations are highly nonlinear. Where LHV is lower heating value of fuel, which is the amount of heat
Therefore, in case of large number of independent equations, it be- released when a unit amount of fuel at 298 K temperature is subjected
comes very difficult to solve them simultaneously using numerical to complete combustion and products are returned to same tempera-
techniques. Besides, non-stoichiometric method is based on the concept ture, provided that the H2O remains in vapor phase. The LHV values of
of Gibbs free energy minimization. According to this concept, total GFE methane (LHVCH4 ), hydrogen (LHVH2 ) and carbon monoxide (LHVCO )
of an isobaric and isothermal reaction system, reaches to its minimum are 802.26, 241.80 and 282.97 kJ/mol respectively and latent heat of
value at equilibrium. This criterion becomes the basis to obtain the steam ( ΔHL ) is 40.69 kJ/mol [50]. QR represents net change in total
composition of reaction components at equilibrium. For the sake of enthalpies of inlet and outlet streams of reformer at reaction tempera-
in
brevity, this method is briefly described below. ture and pressure. HTout
, i and HT , i are the enthalpy of component i in inlet
The total GFE of a reaction system can be expressed by Eq. (1). and outlet streams at temperature T respectively. cP, i represents the heat
capacity of component i expressed as a function of temperature in the
N N
form of a polynomial [52,53]. H298, i is molar enthalpy of component i at
GT = ∑ ni μi =∑ ni (Gfi0 + RTln (yi ϕi P /P0)) 298 K [49,50].
i i (1)

Minimization of the objective function represented by Eq. (1) can be 2.3. Exergy efficiency
carried out with constraints based on the element balance as given by
Eq. (2). Exergy is based on the second law of thermodynamics that deals
with the quality of energy and is concerned about the usefulness and

3
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

the work potential of the energy contained in fuel processing system. [56]. It converts chemical energy of a gaseous fuel directly into elec-
When a system (at T, P) is brought to thermodynamic equilibrium with trical energy and heat by combining fuel with an oxidant (O2/air)
its surrounding/environment (at T0, P0), the maximum useful work electrochemically. It operates at high temperature between
obtained is known as exergy of the system. Exergy gets wasted or de- 973–1273 K. SOFC has a great advantage of using both hydrogen and
stroyed because of occurrence of irreversibility associated with the carbon monoxide as the fuel. Moreover, it provides high fuel flexibility
system. Once exergy is wasted, it can never be recovered, this exergy and tolerance to contaminations [8,57,58]. It consists of a solid oxide
destroyed is viewed as the lost work potential of the system during the electrolyte (ceramic such as Yttria stabilized zirconia), which is sand-
process. wiched between two electrodes. Anode and cathode are generally made
For a stream flowing through the reformer, the total exergy can be of Ni-YSZ (mixture of Ni and Yttria stabilized zirconia), and LSM
expressed as the sum of its all three constituents namely physical ex- (Strontium doped lanthanum manganite) respectively. Oxygen is sup-
ergy, chemical exergy and exergy of mixing [53]. plied to the cathode where it combines with incoming electrons from an
ExT = Ex phy + Ex chem + Ex mix external circuit and form oxide ions which diffuse to the anode through
(7)
electrolyte. At anode, oxide ions react with H2 and CO to produce H2O
The physical exergy is defined as the total amount of work that can and CO2, releasing electrons [58,59].
be obtained when the stream at operating state (T, P) is brought to the Reactions at anode:
environment state (T0, P0) by physical process and there exists thermal H2 + O2− → 2H2O + 2 e−
and mechanical equilibrium [53–55]. The physical exergy of the stream CO + O2− → 2CO2 + 2 e−
is calculated as follows: Reaction at cathode:
Ex phy = ΔH − T0 ΔS O2 + 4 e− → 2 O2−
(8)
The theoretical potential of a fuel cell can be calculated by Nernst
Where, equation.
NS T 0.5
ΔH = ∑ yi ∫ cP,i dT ΔGrxn R TFC ⎛ p H2 pO2 ⎞
ENernst = − + ln ⎜ ⎟
i=1 T0 (9) 2F 2F p (16)
⎝ H2 O ⎠
NS

T
P⎞ Where, ΔGrxn is molar Gibbs free energy of reaction at standard pres-
ΔS = ∑ yi ⎜∫ cP,i dT − R ln P ⎟ sure, F is Faraday constant, R is gas constant, TFC is fuel cell temperature
0
i=1 ⎝ T0 ⎠ (10) and pi is partial pressure of component i in the fuel cell.
The chemical exergy is the work obtainable from the system when The actual cell voltage is always lower than theoretical voltage
the substance under consideration is brought to chemical equilibrium because of three overpotential losses, which occur due to irreversibility
with the environment at reference state at constant T and P reversibly during fuel cell operations. These three losses are activation over-
[42,54]. The chemical exergy of the multicomponent stream can be potential (Vact ) , concentration overpotential (Vconc ) and ohmic over-
expressed by Eq. (11). potential losses (Vohmic ). The actual cell voltage can be determined by
the following expression.
NS
Ex chem = ∑ yi Exchem,i Vactual = ENernst − Vact − Vconc − Vohmic (17)
i=1 (11)
Activation losses are related to the energy barrier that has to be
The exergy of mixing is represented by Eq. (12). It always has a overcome by species for electrochemical reaction to occur.
negative value. Concentration losses are associated to mass transfer resistance of re-
NS actant species. The electrochemical reactions take place at the active
Ex mix = RT0 ∑ yi lnyi sites in the pores of electrodes. The reacting species should diffuse
i=1 (12) through pores to reach these sites. This offers mass transfer resistance
The overall exergy balance of the system under steady state process, which leads to voltage loss. Ohmic overpotential losses occur due to the
can be expressed as electrical resistance offered to the ionic flow through electrolyte and
resistance offered to the electron flow through the interconnections and
ExT , in = ExT , out + Exdest (13) electrodes.
Where, Ex dest represents the exergy destruction that is viewed as the The required molar flow rate of fuel can be determined by following
wasted or lost work potential due to irreversibility of the process that expression.
occurs due to the chemical reactions and large temperature difference I = 2 * F * nfuel * Uf (18)
between the reactants and the products of the reformer [53]. If irre-
versibility is more in a process, there will be larger entropy generation The power output of fuel cell (PFC ) is the product of actual cell
and exergy destruction. voltage (Vact ) and current (I).
Ex dest = T0 Sgen (14) PFC = Vactual × I (19)

Exergy efficiency represents realistic performance of the system. It is Where, I is the current, F is Faraday constant, nfuel is input molar flow
defined as the ratio of useful exergy output to the exergy applied (input) rate of H2 and CO, and Uf is fuel utilization factor. SOFC power gen-
to the process. Therefore, the overall exergy efficiency of a process is eration has been computed by taking total voltage loss of 35 percent of
given by Eq. (15). the theoretical voltage (ENernst ) and fuel utilization factor (Uf ) equals to
70 percent.
ExT , out
ηEx (%) = × 100
ExT , in (15)
3. Results and discussions

2.4. Solid oxide fuel cell and use of product gas as feed In the present study, the thermodynamic simulations have been
performed for biogas steam reforming for a wide range of biogas
Solid oxide fuel cell offers a clean and environment friendly tech- compositions depending on various feedstocks having CH4 from 45% to
nology for electrical power generation with high efficiency of 55–65% 80% and CO2 from 55% to 20%. The reactant stream contains CH4, CO2,

4
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

and H2O, and the product stream contains CH4, CO2, H2O, CO, H2 and moles of CH4, in − moles of CH4, out
Conversion of CH4 = × 100%
solid carbon. It should be noted that a careful selection of product moles of CH4, in (20)
components is very important; otherwise it could lead to misleading
simulation results. Our selection of components of product stream of moles of H2, out
Yield of H2 =
reformer is based on the experimental studies available in the literature moles of CH4, in (21)
[26,39,43,60]. The thermodynamic equilibrium compositions of com-
moles of COout
ponents of product stream are determined by employing R-Gibbs re- Yield of CO =
actor module. This reactor is based on the total GFE minimization, is moles of CH4, in (22)
used for multiple reactions and is operated at constant T and P condi- It is to be noted that yield of CO is generally defined on the basis of
tions. Peng-Robinson equation of state is used for computation of the total moles of CH4 and CO2 in inlet stream, but we have taken basis as
thermodynamic properties of components in the process. Some as- moles of CH4 in inlet stream to show the effect of amount/percentage of
sumptions have been made to carry out simulation work, are given as CH4 in biogas composition. Thermal efficiency and exergy efficiency
have been computed for all operating conditions to find out most en-
(i)Reforming process is considered as a steady state process. ergy efficient process operating conditions for each biogas composition.
(ii)Reactor operated non-adiabatically. According to different thermodynamic studies reported in the lit-
(iii)Pressure drop across reactor is assumed equal to zero. erature, authors have used various software tools to program their
(iv) Reforming is maintained at constant temperature by utilizing an model as discussed in section 2.1. Therefore, prior to proceed further,
external heat supply. the simulations done by R-Gibbs reactor module have been validated
(v) Biogas contains methane and carbon dioxide only. Rest of the against thermodynamic as well as experimental studies available in the
constituents of biogas have been removed prior to steam reforming literature on steam reforming of biogas (SBR). Sufficient efforts are
of biogas. made to validate our methodology with available thermodynamic stu-
dies as shown in Table 2. Our simulation results are found to be in good
Simulations have been carried out to find out the effect of tem- agreement with those reported in the literature. It is worth mentioning
perature, steam to CH4 ratio (S/C) and biogas composition on steam that Roy et al. [38] have not considered the carbon formation for
reforming of biogas (SBR). The operating parameters selected for si- thermodynamic study, but in fact it occurs at 573-923 K depending
mulations are temperature (573–1273 K), S/C ratio (0–2) and CH4 upon S/C ratio (1 and 1.5) and biogas composition (CH4/CO2 = 50/50
(45–80%) in biogas which consists of CH4 and CO2 only. and 60/40). Because of that, our simulation results do not match at low
The performance of SBR process has been analyzed in terms of temperature (823 and 873 K); but at higher temperature (> 973 K), our
conversion of CH4, yield of H2 and CO, as defined below- results are in good agreement. We have also compared our simulation
results with reported experimental studies as presented in Table 3.
Results match closely with Avraam et al. [60], Appari et al. [28], Cipiti

Table 2
Validation of simulation results with computational and thermodynamic studies available in literature.
S.No. Authors Operating Parameters Remarks
conditions
Product Composition (%) XCH4 XCO2 H2/CO H2 yield Carbon
(%) (%) ratio or H2 formation
CH4 CO2 H2O H2 CO fraction* (%)

1. Ashrafi CH4/CO2 = 1.5 0.35 18.36 dry 65.55 15.74 98.30 −33.23 4.16 3.17 0.00 (0.00) Simulations results are in
et al. [43] S/C = 3.5 (0.35) (18.42) basis (65.58) (15.65) (98.32) (-33.83) (4.19) (3.18) good agreement with
P = 1 atm authors
T = 973 K
2. Galvagno CH4/CO2 = 1.5 – 6.55 11.57 54.15 27.73 ⁓99 – 1.95 – 0.00 (0.00) Simulations results are in
et al. [39] S/C = 1.2 (6.53) (11.60) (54.14) (27.73) (98.93) (1.95) good agreement with
P = 1 atm authors
T = 1073K
3. Chiodo CH4/CO2 = 1.5 7.45 16.69 22.37 41.85 11.64 64.05 – – – 0.00 (0.00) Difference may be due to
et al. [62] S/C = 1.8 (7.52) (16.13) (22.27) (42.66) (11.43) (64.25) round off error taken by
P = 1 atm authors.
T = 873 K
4.(a) Roy et al. CH4/CO2 = 1.5 – – – – – 60.10 −5.97 3.29 0.53* not Authors have not
[38] S/C = 1.5 (65.26) (-3.79) (3.71) (0.55*) considered considered carbon
P = 1 atm (7.57) formation for
T = 873 K thermodynamic study, but
(b) CH4/CO2 = 1.0 – – – – – 62.74 5.77 – – not in fact it occurs at
S/C = 1.5 (68.77) (7.27) considered T = 873 K.
P = 1 atm (7.79)
T = 873 K
5. (a) Roy et al. CH4/CO2 = 1.5 – – – – – 91.97 24.31 2.40 0.61* not Though the authors have
[38] S/C = 1.5 (91.80) (24.19) (2.40) (0.61*) considered not considered carbon
P = 1 atm (0.00) formation. Therefore, it is
T = 973 K nil. Fortunately, at these
(b) CH4/CO2 = 1.0 – – – – – 98.29 37.57 – not operating conditions,
S/C = 1.5 (98.20) (37.55) considered carbon formation does not
P = 1 atm (0.00) occur when it is considered
T = 1023 K in thermodynamic
simulation.

Values in parentheses are equilibrium values obtained in this study.


* H2 fraction =H2, out / (CH4+CO2+CO+H2)out.

5
K. Chouhan, et al.

Table 3
Validation of simulation results with experimental studies available in literature.
S. No. Authors Catalyst Operating Parameters Remarks
Conditions
Product Composition XCH4 XCO2 H2/CO H2 Yield COout/ Carbon
(%) (%) (%) ratio or (CH4+CO2)in formation (%)
H2 fraction* or
dry basis CO Yield

1. Ashrafi Ni based CH4/CO2 = 1.5 Dry basis 82.87 −23.34 4.04 2.63 (3.18) ** - (0.00) Lower conversion results in decreased H2 yield.
et al. [55] S/C = 3.5 CH4 = 3.88 (0.35) (98.32) (-33.83) (4.19)
P = 1 atm CO2 = 18.69 (18.43)
T = 973 K H2 = 62.05 (65.58)
CO = 15.38 (15.65)
2. Avraam 5 wt% Ru/Al2O3 CH4/CO2 = 1 Dry basis - (30.51) 2.10 - (2.69) - (0.65) - (0.00) Experimental results are in good agreement with
et al. [56] S/C = 2 CH4 = 0.00 (0.07) ∼100 (2.06) simulation results.
P = 1 atm CO2 = 15.63(14.83) (99.68)
T = 1073 K H2 = 57.14 (57.31)
CO = 27.23 (27.74)
3. Appari 15 wt% Ni/ γ- CH4/CO2 = 1.487 H2 = 31.0(32.62) 97.75 5.6 (6.88) 2.74 - (2.90) - (0.62) - (0.00) Efficient catalyst provides results close to
et al. [28] Al2O3 S/C = 2.02 CO = 11.3 (11.57) (98.12) (2.82) equilibrium values.
P = 1 atm
T = 973 K

6
4. Cipiti 7.5 wt% Ni/CeO2 CH4/CO2 = 1.5 CH4 = 0.27 (0.00) 99.70 6.0 (10.20) 2.80 - (2.93) - (0.64) - (0.00) Simulation results are in good agreement with
et al. [57] S/C = 3.0 CO2 = 9.6 (8.98) (98.93) (2.75) experimental study.
P = 1 atm H2O = 29.15 (31.03)
T = 1173 K H2 = 44.95 (43.97)
CO = 16.03 (16.02)
5. Roy 0.09 wt% [Pd(7)- CH4/CO2 = 1.5 T = 1023 K – 93.9 4.9 (33.79) 3.26 2.72 (2.70) 0.50 (0.72) 4.29 (0.00) Low conversion of CO2 gives lesser CO yield and
et al. [37] Rh(1)] /CeZrO2- S/C = 1.5 (97.50) (2.25) higher H2/CO ratio. Carbon formation was reported
Al2O3 P = 1 atm T = 1073 K – 96.80 10.6 2.91 2.90 (2.71) 0.61 (0.76) 3.20 (0.00) on catalyst surface.
(99.26) (40.45) (2.15)
T = 1123 K – 99.06 17.6 2.74 3.05 (2.69) 0.6 (0.78) 2.94 (0.00)
(99.77) (45.70) (2.06)
6. Roy et al. 1.31 wt% [Pd(7)- CH4/CO2 = 1.5 T = 973 K – 89.42 −22.78 3.94 0.59* (0.61*) - (0.65) 1.63 (0.00) Catalyst does not favor dry reforming reaction
[38] Rh(1)] /CeZrO2- S/C = 1.5 (91.80) (24.19) (2.40) which provides lower conversion of CO2 and higher
Al2O3 P = 1 atm T = 1023 K – 95.47 −7.88 3.36 0.60* (0.62*) - (0.72) 2.16 (0.00) H2/CO ratio. Carbon formation occurred on catalyst
(97.50) (33.79) (2.25) surface.
T = 1073 K – 97.19 0.69 3.05 0.60* (0.62*) - (0.76) 2.88 (0.00)
(99.26) (40.45) (2.15)
T = 1123 K – 98.15 11.15 2.76 0.60* (0.62*) - (0.78) 2.20 (0.00)
(99.77) (45.70) (2.06)

Values in parentheses are equilibrium values obtained in this study.


*H2 fraction = H2, out / (CH4+CO2+CO+H2)out.
**It may be noted that Ashrafi et al. [43] have calculated CO selectivity instead of CO yield. Formula used by them is CO selectivity = COout / (CO+CO2)out. Their computed CO selectivity was 0.45 (46), which is close to
our result.
Journal of Environmental Chemical Engineering 7 (2019) 103018
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

et al. [61] and Roy et al. [38] except that the carbon formation has been presence of CO2 in BG results in higher conversion of CH4
observed by Roy et al. [37,38]. (573–1023 K). This difference in conversion of CH4 for different BG
After validation of simulation results, thermodynamic equilibrium compositions disappears at 1073 K and above. The conversion of CH4
compositions have been computed for different parameters as men- reaches to ≥99% for all BG compositions at 1073 K.
tioned previously in this section. All the results are analyzed to find out Fig. 3(b) depicts the variational trends of conversion of CH4 as a
operating conditions to get ≥ 99% conversion of CH4 with reasonably function of BG composition and steam to CH4 ratio at constant re-
high H2 yield and that too at the temperature where the problems of forming temperature of 1073 K for SBR process. For biogas having 45%
sintering of catalyst as well as carbon formation do not occur. Fig. 1(a) CH4, conversion of CH4 (XCH4) increases very slowly with increasing S/
and (b) are shown to illustrate this criterion of selecting the operating C ratio from 96.64% to 99.73% for S/C ratio of 0 and 2 respectively,
conditions. Fig. 1(a) shows the variational trends between conversion because of higher amount of CO2 (55%) in BG. High presence of CO2
of CH4 and temperature over S/C ratio (0–2) for biogas composition of results in the occurrence of RWGS reaction (R4) along with R1-R3. It
CH4/CO2 = 55/45. Conversion of CH4 increases with temperature for should be noted that for BG compositions (50–80% CH4), XCH4 de-
all S/C ratios because of endothermic nature of reactions R1-R3 as creases marginally with the increase in S/C ratio; reaches its minimum
listed in Table 4. In the region of 573–873 K, at constant temperature, it value at S/C = 0.2, 95.89% (for 50% CH4), 95.52% (for 55% CH4); at
decreases with increasing S/C ratio due of lesser occurrence of reaction S/C = 0.4, 94.80% (60% CH4); at S/C = 0.6, 94.19% (65% CH4),
R3 which is verified by reduced carbon formation observed in SBR 93.78% (70% CH4) and 93.47% (75% CH4); at S/C = 0.8, 92.85% (80%
process. CH4). This reduction of XCH4 with increase in S/C ratio occurs because
In contrast, at constant temperature in the range of 873–1073 K, adding steam results into lesser occurrence of the methane decom-
CH4 conversion starts increasing as S/C ratio is raised due to en- position reaction R3 (CH4 → C + 2 H2) which can be verified by higher
dothermic reactions R1 and R2; which can be confirmed by increased carbon formation at lower S/C ratio. After reaching minimum value,
production of H2 as shown in Fig. 1(b). It should be noted that above XCH4 goes up smoothly to attain its peak value which varies between
1073 K, CH4 conversion rises marginally with S/C ratio. H2 yield also 99.73 to 99.39% for BG compositions of 45% CH4 and 80% CH4 re-
has its maximum value at 1073 K which is true for S/C = 0.8–2.0. spectively, at S/C = 2. It is worth noting that further increasing S/C
Moreover, the increment in H2 yield is also marginal above S/C = 1.6 ratio beyond 1.6, yields into negligible increment in XCH4 value. Also, it
and there is no carbon formation at this temperature. Therefore, S/ attains its value ≥99% for all BG compositions at S/C = 1.6.
C = 1.6 and T = 1073 K are chosen to show our simulation results in
the subsequent sections.
3.3. Yield of hydrogen

3.1. Carbon formation


Fig. 4(a) shows the effect of reforming temperature and BG com-
position on the yield of hydrogen which is defined as moles of hydrogen
Carbon formation over catalyst surface has detrimental effect on the
catalyst performance which leads to low activity, deactivation, sin-
tering and reduced lifetime of catalyst. It also causes the increased
pressure drop across reactor. Therefore, it is essential to determine the
optimum operating conditions thermodynamically which prevent the
carbon formation in the SBR process. The possible reactions responsible
for carbon formation are decomposition of methane (R3), Boudouard
reaction (R5), carbon monoxide and carbon dioxide reduction reactions
(reverse R6 and R7).
Fig. 2 illustrate the carbon formation and carbon free zones at all
temperature and S/C ratios for all biogas compositions (CH4/
CO2 = 45/55 to 80/20). At S/C = 0-0.6, carbon formation occurs at all
temperature (573–1473 K), for S/C = 0.8 and above 1108 K, it starts
becoming extinct from the reaction network. This substantial change
lies in the range of 523–1108 K over S/C between 0.8 and 3.8. The last
carbon formation is observed at 523 K and S/C = 3.8 and above these
conditions It is absolutely eliminated from the SBR process thermo-
dynamically.

3.2. Conversion of CH4

Several simultaneous and consecutive reactions occur in Steam re-


forming of biogas (SBR) as listed in Table 4. Fig. 3(a) illustrates the
effect of reaction temperature and biogas composition on conversion
efficiency of CH4. For all biogas compositions, it increases sharply when
temperature is raised from 573 K to 1023 K because of endothermic
nature of reactions (R1-R3) listed in Table 4. Further increase in re-
forming temperature beyond 1023 K, conversion of CH4 goes up
smoothly to attain maximum value (99.99%) for all BG compositions at
1273 K. For the reforming temperature range of 573–973 K, it decreases
as amount of CH4 in BG increases at a constant temperature, which is
because of the fact that amount of CO2 in BG decreases with the in-
crease in amount of CH4 in BG. In lower temperature range
(573–773 K), methane decomposition reaction (R3) is more pro-
nounced. This can be verified by carbon/coke formation and small Fig. 1. Effect of temperature and S/C ratio on (a) conversion of CH4 and (b) H2
amount of CO produced as compared to H2 produced. Higher the Yield for CH4/CO2 = 55/45.

7
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

Table 4
Chemical reactions involved in steam reforming of biogas.
Reaction No. Reactions Reaction types ΔHR (kJ/mol) at 298 K, 1 atm

R1 CH4 + CO2 ↔ 2CO + 2H2 Dry reforming of methane 247.28


R2 CH4 + H2O ↔ CO + 3H2 Steam reforming of methane 206.11
R3 CH4 ↔ C + 2H2 Decomposition of methane 74.85
R4 CO2 + H2 ↔ CO + H2O Reverse water gas shift reaction 41.17
R5 2CO ↔ C + CO2 Boudouard reaction −172.43
R6 C + H2O ↔ CO + H2 Carbon gasification 131.26
R7 CO2 + 2H2 ↔ C + 2H2O Carbon dioxide reduction reaction −90.09
R8 CO2 + 4H2 ↔ CH4 + 2H2O Methanation reaction −164.94

Fig. 2. Carbon formation and carbon free zones at various temperatures and S/
C ratios for BG compositions 45/55 to 80/20.

produced per mol of CH4 fed to the reformer at the pressure of 1 atm.
The H2 yield increases with the rise in reforming temperature from
573 K, reaches its peak value at temperature of 1023 K which is 2.52
(45% CH4), 2.60 (50% CH4), 2.67% (55%), and at 1073 K peak H2
yields are 2.74 (60% CH4), 2.80 (65% CH4), 2.86 (70% CH4), 2.91 (75%
CH4) and 2.95 (80% CH4) and then its value reduces slightly with the
increase in temperature up to 1273 K. Increasing H2 yield with in-
creasing temperature is because of the fact that the reactions re-
sponsible for hydrogen production (R1-R3, R6), are endothermic in
nature and their dominance increases with temperature. After reaching
peak value, H2 yield decreases with increasing temperature beyond
1023 K because of the RWGS reaction (R4) which consumes H2 formed
from reactions (R1-R3, R6) along with CO2 to produce CO and H2O
which is verified by increased amount of CO and H2O produced in this
temperature range. It should also be noted that H2 yield remains also
same for all BG compositions in lower temperature range 573–873 K. At Fig. 3. Conversion of CH4 for different BG compositions (a) effect of tempera-
923 K and above, the difference between H2 yields of different BG ture at constant S/C = 1.6 and (b) effect of S/C ratio at constant temperature of
compositions get widen. This phenomenon can be explained by the fact 1073 K.
that for the BG with higher content of CH4, shows higher occurrence of
reaction R2 than that of R1 because of lower CO2 content in BG which is mixture is fed to SOFC for electrical power generation. Fig. 5(a) re-
rich in CH4. Hence, It may be stated that the richer the BG in CH4, presents the effect of change in reforming temperature and BG com-
higher the H2 yield at 1023 K and above at S/C = 1.6. positions on the yield of CO while SBR process is operated at 1 atm
Fig. 4(b) depicts the variational trends of the H2 yield as a function pressure and S/C = 1.6. In lower temperature region of 573–723 K, CO
of S/C ratio and BG composition. It can be noticed that for all BG yield which represents the moles of CO produced per mol of CH4 in BG
compositions, the H2 yield increases with the increase in S/C ratio fed to the reformer, remains constant near to zero value for all BG
because of dominance of reaction R2 at 1073 K. Higher the amount of compositions. It occurs due to CH4 decomposition reaction R3 becomes
CH4 in BG results in higher H2 yield with increase in S/C ratio at very prominent in this region which leads to formation of carbon and
1073 K. This can be explained by the occurrence of reaction R4 in re- H2. Increasing reforming temperature beyond 723 K, CO yield increases
verse direction in which excess steam reacts with CO formed/produced sharply as reforming temperature is raised from 723 K to 1273 K; which
by reactions (R1-R3) to produce CO2 and H2, ultimately resulting into may be due to endothermic nature of reactions namely; dry reforming
higher yield of H2 and CO. reaction R1, steam reforming reaction R2 and carbon gasification re-
action R6.
3.4. Yield of CO Fig. 5(b) shows the variational trends of the CO yield as a function
of S/C ratio and BG compositions. For 45% CH4, CO yield decreases
Yield of CO becomes very important parameter when product gas with increasing S/C ratio which may be due to the dominance of WGS

8
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

Fig. 4. H2 yield for different BG compositions (a) effect of temperature at Fig. 5. CO yield for different BG compositions (a) effect of temperature at
constant S/C = 1.6 and (b) effect of S/C ratio at constant temperature of constant S/C = 1.6 and (b) effect of S/C ratio at constant temperature of
1073 K. 1073 K.

reaction (reverse R4). The CO yield for all other BG compositions (50- (SBR) process is a key parameter. The energy cost reflects in the product
80% CH4) first increases and reaches its peak value and then decreases cost. Higher the amount of energy requirement, greater the cost of
with the increase in S/C ratio at 1073 K and 1 atm. It can be interpreted production of any desired chemical. The energy requirement for SBR
that at constant S/C ratio, higher the amount of CO2 (lower the amount process for different BG compositions at various operating conditions
of CH4) in BG results in higher yield of CO because of higher amount of i.e. reforming temperature and S/C ratio; is computed by using Eq. (4).
CO2 reacts which CH4 to produce two times moles of CO in reaction R1 Fig. 7(a) shows the energy requirement trends for BG having
as compared to that produced by reaction R2. 45–80% CH4 at S/C = 1.6 for change in temperature. According to
Fig. 7(a), the energy requirement increases sharply with the increasing
3.5. Yield of (H2+CO) temperature from 573 to 1023 K which may be explained by the pro-
gress of endothermic reactions R1-R3, R4 and R6. Thereafter, it goes up
Fig. 6(a) illustrates the effect of temperature and BG compositions smoothly, but rise in energy requirement beyond 1073 K is very small
on the yield of (H2+CO) at constant S/C = 1.6. For all BG composi- because of nearly complete conversion of CH4 at this temperature. It is
tions, the yield of (H2+CO) increases with temperature but remains worth noting that as the BG gets rich in CH4, energy requirement in-
same for all BG compositions at the same temperature in the tem- creases because of the higher amount of CH4, which gets consumed in
perature range from 573 to 823 K. In the region of 873–1023 K, dif- reactions R1-R4, which leads to rise in overall endothermicity of re-
ference in (H2+CO) yield for different BG composition is well pro- forming process. The energy requirement for SBR process operating at
nounced. It remains same for all BG compositions at 1073 K and above. 1073 K, varies between 108.99 and 179.25 kJ.
Operating conditions of 1073 K and S/C = 1.6, provides highest yield of Fig. 7(b) depicts the effect of S/C ratio on the energy requirement
(H2+CO) which ranges between 3.96 and 3.98 mol of (H2+CO) per (QR) for steam reforming of biogas with different compositions at
mol of CH4. constant temperature of 1073 K. At S/C = 0, the amount of energy re-
Fig. 6(b) depicts the effect of S/C ratio and BG compositions on the quirement for reforming process is highest for BG having 45% CH4; as
yield of (H2+CO) at constant temperature of 1073 K. For all BG com- BG gets rich in CH4, energy requirement decreases because of lower
positions, the yield of (H2+CO) increases with increasing S/C ratio amount of CO2 available for dry reforming of methane reaction R1
from 0 to 1.6, and remains same above S/C = 1.6 for all BG composi- which is the most endothermic reaction in the reaction network as is
tions. obvious from Table 4. For all other BG compositions (50–80% CH4),
energy requirement first increases and reaches its peak value of
3.6. Energy requirement 123.90 kJ (S/C = 0.4), 133.06 kJ (S/C = 0.6), 142.30 kJ (S/C = 0.8),
151.59 kJ (S/C = 1.0), 160.87 kJ (S/C = 1.0), 170.35 kJ (S/C = 1.2),
Energy requirement associated with steam reforming of biogas and 179.71 kJ (S/C = 1.4) for BG having 50% to 80% CH4 (step size 5)

9
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

Fig. 6. [H2+CO] yield for different BG compositions (a) effect of temperature


at constant S/C = 1.6 and (b) effect of S/C ratio at constant temperature of
1073 K. Fig. 7. Energy requirement for SBR process for different BG compositions (a)
effect of temperature at constant S/C = 1.6 and (b) effect of S/C ratio at con-
stant temperature of 1073 K.
respectively. After reaching peak value, QR decreases very marginally
with the increase in S/C ratio. This can be explained by prominent
energy input to the SBR process is recovered in the form of useful
occurrence of steam reforming reaction R2 at higher S/C ratio; which
products H2 and CO and remaining 8% is exhausted with other gaseous
needs less energy than dry reforming reaction R1.
products for all BG compositions. Therefore, it can be concluded that
SBR process is throughout endothermic at all operating conditions,
SBR process with BG compositions (45–80% CH4) can be used to pro-
as shown in Fig. 7. Energy requirement corresponding to reforming
duce almost constant combined moles of H2 and CO which can be di-
temperature of 1073 K and steam to CH4 ratio of 1.6, for SBR process is
rectly used as a fuel in SOFC for electrical power generation.
listed in Table 5.
At constant reforming temperature of 1073 K, thermal efficiency of
SBR process with different BG compositions (45–80% CH4), has been
3.7. Thermal efficiency computed to determine the effect of change in S/C ratio which is shown
in Fig. 8(b). For all BG compositions, thermal efficiency of SBR process
Thermal efficiency is computed by using Eq. (3) for all BG compo- first increases with increasing S/C ratio because of increased production
sitions at (i) different operating temperatures of reformer while keeping of H2 and CO. It is worth noting that steam reforming of BG having
S/C = 1.6, and (ii) different S/C ratios at 1073 K; these are represented lower content of CH4 yields in higher thermal efficiency as compared to
by Fig. 8(a) and (b), respectively. It is clear from Eq. (3) that thermal BG with higher CH4 content, which is because of the fact that high
efficiency of SBR process is directly proportional to the production of amount of CO2 in BG results in higher conversion of CH4 into H2 via
moles of H2 and CO, and is inversely proportional to the sum of LHV of DRM reaction R1 and WGS reaction (reverse R4). As S/C ratio in-
fuel, overall heat of reaction and amount of steam in the feed. As shown creases, the difference between thermal efficiency values of different
in Fig. 8(a), thermal efficiency of SBR process increases with the in- BG compositions narrows down. After reaching peak value, it goes
creasing temperature from 573 to 973 K because of the endothermicity down smoothly with further increase in S/C ratio. This reducing trend
of process which favors higher production of H2 and CO and thereafter, of thermal efficiency is due to the fact that the amount of heat required
it goes up smoothly on further increasing the temperature from 973 to for vaporization of H2O, increases with the increasing S/C ratio, but
1073 K, attains almost constant value between 92.39% and 91.46% for increase in production of H2 and CO is very marginal. It results in
BG containing 45% and 80% CH4 respectively. The rise in thermal ef- higher amount of heat required for vaporization of water as compared
ficiency of SBR is very marginal (insignificant) from 1073 to 1273 K. to the energy available from H2 and CO production with increasing S/C
The maximum thermal efficiencies for SBR process, achieved lie be- ratio.
tween 92.83% and 92.25% for BG containing 45% CH4 and 80% CH4 Thermal efficiency for SBR process at reforming temperature of
respectively. 1073 K and S/C ratio of 1.6, have been reported in Table 5, and it varies
Thermal efficiency is a measure of performance of system in terms from 92.39% to 91.46% for BG having CH4 content of 45 and 80%
of utilization of energy supplied. These results reveal that 92% of the respectively.

10
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

Table 5
Results of the steam reforming of biogas (SBR) process at S/C = 1.6, T = 1073 K, and P = 1 atm.
Parameters CH4 in Biogas

45% 50% 55% 60% 65% 70% 75% 80%

XCH4 (%) 99.63 99.54 99.45 99.35 99.25 99.15 99.04 98.93
XCO2 (%) 40.75 40.27 38.98 96.69 33.12 27.71 19.48 6.43
H2 yield (mol/mol) 2.49 2.58 2.66 2.74 2.80 2.86 2.91 2.95
CO yield (mol/mol) 1.49 1.40 1.31 1.24 1.17 1.11 1.06 1.01
(H2+CO) yield (mol/mol) 3.98 3.98 3.97 3.98 3.97 3.97 3.97 3.96
H2/CO ratio 1.67 1.85 2.03 2.21 2.39 2.57 2.75 2.94
C (mol/mol) Nil Nil Nil Nil Nil Nil Nil Nil
Energy requirement (kJ/mol) 108.99 119.36 129.60 139.72 149.73 159.66 169.49 179.25
Thermal efficiency (%) 92.39 92.24 92.10 91.97 91.84 91.71 91.59 91.46
Exergy efficiency (%) 75.57 74.45 73.40 72.42 71.52 70.67 69.89 69.16
SOFC Power generation (kW) (based on 1 mol/s of biogas) 157.46 175.12 192.74 210.30 227.82 245.28 262.70 280.06

SBR of different BG compositions (45–80% CH4) results in nearly same


value of exergy efficiency at constant temperature. At 873 K and above,
SBR of BG containing 45% CH4 shows the highest exergy efficiency.
Increase in exergy efficiency is very sharp up to 973 K, thereafter it goes
up smoothly with further rise in reforming temperature. The exergy
efficiency gets reduced as BG becomes rich in CH4, and BG with 80%
CH4 has lowest exergy efficiency among all BG compositions as shown
in Fig. 9(a). It may be due to the fact that total moles of H2 and CO
produced for all BG compositions are same but input exergy is higher
for BG having higher CH4 content, which ultimately results in lower
exergy efficiency of steam reforming with BG rich in CH4 content. There
is very marginal increase in exergy efficiency for different BG

Fig. 8. Thermal efficiency of SBR process for different BG compositions (a)


effect of temperature at constant S/C = 1.6 and (b) effect of S/C ratio at con-
stant temperature of 1073 K.

3.8. Exergy efficiency

The exergy efficiency is based on the second law of thermodynamics


and is computed by using Eq. (15) for steam reforming of biogas at
different S/C ratio, reforming temperature and biogas composition.
Fig. 9(a) represents the variational trends of the exergy efficiency of
SBR process at S/C = 1.6 with change in process variable such as
temperature and molar feed ratio of S/C. Exergy efficiency calculation
is carried out by considering both H2 and CO as products, because this
mixture can be used as a fuel to power SOFC.
Fig. 9(a) shows that the exergy efficiency increases with the in-
creasing reforming temperature because of the progress of reactions R1-
Fig. 9. Exergy efficiency of SBR process for different BG compositions (a) effect
R3 in forward direction which leads to the increased production of of temperature at constant S/C = 1.6 and (b) effect of S/C ratio at constant
moles of H2 and CO. In the lower temperature region from 573 to 823 K, temperature of 1073 K.

11
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

compositions with temperature at S/C = 1.6. Highest value of the ex- this crucial research area of ‘Renewable Energy’ in the department.
ergy efficiency varies from 75.82 to 69.99% for BG having 45% and
80% CH4 respectively, at 1273 K. References
Fig. 9(b) illustrates the effect of S/C ratio and BG compositions on
exergy efficiency of SBR process at reforming temperature of 1073 K. It [1] A. Rafiee, K. Rajab Khalilpour, D. Milani, M. Panahi, Trends in CO2 conversion and
clearly shows that at 1073 K the exergy efficiency for all BG composi- utilization: a review from process systems perspective, J. Environ. Chem. Eng. 6
(2018) 5771–5794, https://doi.org/10.1016/J.JECE.2018.08.065.
tions increases with the rise in S/C ratio from 0 to 2 with step size of [2] G. Nahar, D. Mote, V. Dupont, Hydrogen production from reforming of biogas:
0.2, which leads to higher production of moles of H2. BG with 45% CH4 review of technological advances and an Indian perspective, Renew. Sustain.
content shows the highest exergy efficiency for steam reforming pro- Energy Rev. 76 (2017) 1032–1052, https://doi.org/10.1016/j.rser.2017.02.031.
[3] E. Bocci, A. Di Carlo, S.J. McPhail, K. Gallucci, P.U. Foscolo, M. Moneti, M. Villarini,
cess. As BG gets rich in CH4 content exergy efficiency reduces due to M. Carlini, Biomass to fuel cells state of the art: a review of the most innovative
higher amount of CH4 in feed which leads to higher input exergy of the technology solutions, Int. J. Hydrogen Energy 39 (2014) 21876–21895, https://doi.
reforming process. The maximum exergy efficiency achieved at S/C = 2 org/10.1016/j.ijhydene.2014.09.022.
[4] S. Rasi, A. Veijanen, J. Rintala, Trace compounds of biogas from different biogas
because of higher production of total moles of H2 and CO. production plants, Energy 32 (2007) 1375–1380, https://doi.org/10.1016/j.
energy.2006.10.018.
3.9. SOFC power generation [5] C.A. Schwengber, F.A. Da Silva, R.A. Schaffner, N.R.C. Fernandes-Machado,
R.J. Ferracin, V.R. Bach, H.J. Alves, Methane dry reforming using Ni/Al2O3 cata-
lysts: evaluation of the effects of temperature, space velocity and reaction time, J.
In SBR process, product gas is a mixture of H2, CO, CO2 and H2O; Environ. Chem. Eng. 4 (2016) 3688–3695, https://doi.org/10.1016/j.jece.2016.07.
H2O can be easily condensed and thus removed from the product gas. 001.
Since, CO2 does not have any adverse effect on SOFCs, so H2O free [6] A.F. Lucrédio, J.M. Assaf, E.M. Assaf, Reforming of a model sulfur-free biogas on Ni
catalysts supported on Mg(Al)O derived from hydrotalcite precursors: effect of La
product gas can be directly used as a fuel to SOFCs for electrical power and Rh addition, Biomass Bioenergy 60 (2014) 8–17, https://doi.org/10.1016/j.
generation. SOFCs produce H2O and CO2 from which CO2 can be se- biombioe.2013.11.006.
parated by condensing H2O. The CO2 thus obtained can be further [7] A. Iulianelli, S. Liguori, Y. Huang, A. Basile, Model biogas steam reforming in a thin
Pd-supported membrane reactor to generate clean hydrogen for fuel cells, J. Power
purified to food grade by cryogenic distillation so that it can be used in Sources 273 (2015) 25–32, https://doi.org/10.1016/j.jpowsour.2014.09.058.
soft drinks and beverage industries. [8] H.J. Alves, C. Bley Junior, R.R. Niklevicz, E.P. Frigo, M.S. Frigo, C.H. Coimbra-
The electrical power generation by SOFC operating at 1 atm pres- Araújo, Overview of hydrogen production technologies from biogas and the ap-
plications in fuel cells, Int. J. Hydrogen Energy 38 (2013) 5215–5225, https://doi.
sure and 1073 K temperature with feed from product gas of SBR process org/10.1016/j.ijhydene.2013.02.057.
(P = 1 atm, T = 1073 K and S/C = 1.6), are shown in Table 5. It could [9] Y. Gao, J. Jiang, Y. Meng, F. Yan, A. Aihemaiti, A review of recent developments in
generate power between 157.46 and 280.06 kW for biogas having hydrogen production via biogas dry reforming, Energy Convers. Manage. 171
(2018) 133–155, https://doi.org/10.1016/j.enconman.2018.05.083.
methane between 45% and 80%, respectively; the computations are [10] A.G. Chmielewski, A. Urbaniak, K. Wawryniuk, Membrane enrichment of biogas
based on biogas molar flow rate of one mole per second as feed to the from two-stage pilot plant using agricultural waste as a substrate, Biomass
reformer. Bioenergy 58 (2013) 219–228, https://doi.org/10.1016/j.biombioe.2013.08.010.
[11] I. Ullah Khan, M. Hafiz Dzarfan Othman, H. Hashim, T. Matsuura, A.F. Ismail,
M. Rezaei-DashtArzhandi, I. Wan Azelee, Biogas as a renewable energy fuel – a
4. Conclusions review of biogas upgrading, utilisation and storage, Energy Convers. Manage. 150
(2017) 277–294, https://doi.org/10.1016/j.enconman.2017.08.035.
Detailed thermodynamic evaluation combined with energetic and [12] C.E. Granada, C. Hasan, M. Marder, O. Konrad, L.K. Vargas, L.M.P. Passaglia,
A. Giongo, R.R. de Oliveira, L. de M. Pereira, F. de Jesus Trindade, R.A. Sperotto,
exergetic analyses for syngas production via steam reforming of biogas Biogas from slaughterhouse wastewater anaerobic digestion is driven by the ar-
(SBR) of varying compositions (CH4:CO2 = 45:55 to 80:20) has been chaeal family Methanobacteriaceae and bacterial families Porphyromonadaceae
presented. Steam reforming of biogas produces the syngas which can be and Tissierellaceae, Renew. Energy 118 (2018) 840–846, https://doi.org/10.1016/
j.renene.2017.11.077.
used as a fuel to solid oxide fuel cell. The best operating conditions for [13] R. Kadam, N.L. Panwar, Recent advancement in biogas enrichment and its appli-
the production of combined moles of H2 and CO for all the biogas cations, Renew. Sustain. Energy Rev. 73 (2017) 892–903, https://doi.org/10.1016/
compositions under study, are P = 1 atm, T = 1073 K and Steam to CH4 j.rser.2017.01.167.
[14] S. Damyanova, B. Pawelec, K. Arishtirova, J.L.G. Fierro, Biogas reforming over
molar ratio = 1.6. Under these conditions, CH4 conversion remains bimetallic PdNi catalysts supported on phosphorus-modified alumina, Int. J.
≥99% and carbon formation is completely absent for all biogas com- Hydrogen Energy 36 (2011) 10635–10647, https://doi.org/10.1016/j.ijhydene.
positions. H2 yield (2.49–2.95), CO yield (1.49-1.01) and H2/CO ratio 2011.05.098.
[15] C.S. Lau, A. Tsolakis, M.L. Wyszynski, Biogas upgrade to syn-gas (H2-CO) via dry
(1.67–2.94) have been achieved for CH4:CO2 = 45:55 to 80:20. Energy
and oxidative reforming, Int. J. Hydrogen Energy 36 (2011) 397–404, https://doi.
requirement per mole of biogas varies between 108.99 and 179.25 kJ. org/10.1016/j.ijhydene.2010.09.086.
Furthermore, this SBR process has shown thermal efficiency (92.39- [16] M. Genoveva Zimicz, B.A. Reznik, S.A. Larrondo, Conversion of biogas to synthesis
gas over NiO/CeO2-Sm2O3 catalysts, Fuel. 149 (2015) 95–99, https://doi.org/10.
91.46%) and exergy efficiency (75.57-69.16%).
1016/j.fuel.2014.09.024.
This study proposes that the utilization of biogas obtained from [17] N.D. Charisiou, G. Siakavelas, K.N. Papageridis, A. Baklavaridis, L. Tzounis,
various sources via steam reforming is an efficient process to produce D.G. Avraam, M.A. Goula, Syngas production via the biogas dry reforming reaction
syngas for SOFC applications. Product gas from this reforming process over nickel supported on modified with CeO2 and/or La2O3 alumina catalysts, J.
Nat. Gas Sci. Eng. 31 (2016) 164–183, https://doi.org/10.1016/j.jngse.2016.02.
contains H2, CO, CO2 and H2O. This product gas can directly fed to the 021.
SOFC as fuel and can generate the electrical power between 157.5 (for [18] P. Ugarte, P. Durán, J. Lasobras, J. Soler, M. Menéndez, J. Herguido, Dry reforming
45% CH4) and 280 kW (for 80% CH4). of biogas in fluidized bed: process intensification, Int. J. Hydrogen Energy 42
(2017) 13589–13597, https://doi.org/10.1016/j.ijhydene.2016.12.124.
[19] X. Chen, J. Jiang, K. Li, S. Tian, F. Yan, Energy-efficient biogas reforming process to
Declaration of interest produce syngas: the enhanced methane conversion by O2, Appl. Energy 185 (2017)
687–697, https://doi.org/10.1016/j.apenergy.2016.10.114.
[20] J. Xu, W. Zhou, Z. Li, J. Wang, J. Ma, Biogas reforming for hydrogen production
There is no conflict of interest of any kind with anyone. over a Ni-Co bimetallic catalyst: effect of operating conditions, Int. J. Hydrogen
Energy 35 (2010) 13013–13020, https://doi.org/10.1016/j.ijhydene.2010.04.075.
Acknowledgements [21] K. Tao, Y. Zhang, S. Terao, N. Tsubaki, Development of platinum-based bimodal
pore catalyst for CO2 reforming of CH4, Catal. Today 153 (2010) 150–155, https://
doi.org/10.1016/j.cattod.2010.02.061.
The authors wish to express their gratitude to the Ministry of [22] A. Kambolis, H. Matralis, A. Trovarelli, C. Papadopoulou, Ni/CeO2-ZrO2 catalysts
Human Resource Development, Govt. of India, New Delhi, for providing for the dry reforming of methane, Appl. Catal. A Gen. 377 (2010) 16–26, https://
doi.org/10.1016/j.apcata.2010.01.013.
financial support for this research work. We humbly dedicate this re-
[23] A. Serrano-Lotina, L. Daza, Highly stable and active catalyst for hydrogen produc-
search work to Late Prof. (Mrs.) Shashi who left to her heavenly abode tion from biogas, J. Power Sources 238 (2013) 81–86, https://doi.org/10.1016/j.
in the recent past. She was indeed responsible for initiating research in jpowsour.2013.03.067.

12
K. Chouhan, et al. Journal of Environmental Chemical Engineering 7 (2019) 103018

[24] O.A. Bereketidou, M.A. Goula, Biogas reforming for syngas production over nickel catalytic steam reforming: 1, Thermodynamic optimization, Energy and Fuels. 22
supported on ceria – alumina catalysts, Catal. Today 195 (2012) 93–100, https:// (2008) 4182–4189, https://doi.org/10.1021/ef800081j.
doi.org/10.1016/j.cattod.2012.07.006. [44] T. Sato, T. Suzuki, M. Aketa, Y. Ishiyama, K. Mimura, N. Itoh, Steam reforming of
[25] B. Li, S. Zhang, Methane reforming with CO2 using nickel catalysts supported on biogas mixtures with a palladium membrane reactor system, Chem. Eng. Sci. 65
yttria-doped SBA-15 mesoporous materials via sol-gel process, Int. J. Hydrogen (2010) 451–457, https://doi.org/10.1016/j.ces.2009.04.013.
Energy 38 (2013) 14250–14260, https://doi.org/10.1016/j.ijhydene.2013.08.105. [45] U. Izquierdo, V.L. Barrio, J.F. Cambra, J. Requies, M.B. Güemez, P.L. Arias, G. Kolb,
[26] S.A. Chattanathan, S. Adhikari, M. McVey, O. Fasina, Hydrogen production from R. Zapf, A.M. Gutiérrez, J.R. Arraibi, Hydrogen production from methane and
biogas reforming and the effect of H2S on CH4 conversion, Int. J. Hydrogen Energy natural gas steam reforming in conventional and microreactor reaction systems, Int.
39 (2014) 19905–19911, https://doi.org/10.1016/j.ijhydene.2014.09.162. J. Hydrogen Energy 37 (2012) 7026–7033, https://doi.org/10.1016/j.ijhydene.
[27] P. Djinović, I.G.O. Črnivec, A. Pintar, Biogas to syngas conversion without carbo- 2011.11.048.
naceous deposits via the dry reforming reaction using transition metal catalysts, [46] C. Italiano, A. Vita, C. Fabiano, M. Laganà, L. Pino, Bio-hydrogen production by
Catal. Today 253 (2015) 155–162, https://doi.org/10.1016/j.cattod.2015.01.039. oxidative steam reforming of biogas over nanocrystalline Ni/CeO2 catalysts, Int. J.
[28] S. Appari, V.M. Janardhanan, R. Bauri, S. Jayanti, Deactivation and regeneration of Hydrogen Energy 40 (2015) 11823–11830, https://doi.org/10.1016/j.ijhydene.
Ni catalyst during steam reforming of model biogas: an experimental investigation, 2015.04.146.
Int. J. Hydrogen Energy 39 (2014) 297–304, https://doi.org/10.1016/j.ijhydene. [47] N. Katiyar, S. Kumar, S. Kumar, Thermodynamic analysis for quantifying fuel cell
2013.10.056. grade H2 production by methanol steam reforming, Chem. Eng. Technol. 36 (2013)
[29] M.M. Danilova, Z.A. Fedorova, V.I. Zaikovskii, A.V. Porsin, V.A. Kirillov, 581–590, https://doi.org/10.1002/ceat.201200540.
T.A. Krieger, Applied catalysis B : environmental porous nickel-based catalysts for [48] A. Vita, L. Pino, F. Cipitì, M. Laganà, V. Recupero, Biogas as renewable raw material
combined steam and carbon dioxide reforming of methane, Appl. Catal. B, Environ. for syngas production by tri-reforming process over NiCeO2 catalysts: optimal op-
147 (2014) 858–863, https://doi.org/10.1016/j.apcatb.2013.10.005. erative condition and effect of nickel content, Fuel Process. Technol. 127 (2014)
[30] M. Miyake, N. Tomiyama, K. Iwamoto, K. Nagase, S. Nishimoto, Y. Kameshima, 47–58, https://doi.org/10.1016/j.fuproc.2014.06.014.
Biogas reforming over BaTi1-xSnxO3-supported Ni-based catalysts recovered from [49] R.H. Perry, D.W. Green, J.O. Maloney, Perry’s Chemical Engineers’ Handbook, se-
spent Ni-metal-hydride batteries, Int. J. Hydrogen Energy 40 (2015) 8341–8346, venth ed., McGraw-Hill, New York, 1999.
https://doi.org/10.1016/j.ijhydene.2015.04.106. [50] Y.A. Cengel, M.A. Boles, Thermodynamics: An Engineering Approach, sixth ed.,
[31] S.D. Angeli, L. Turchetti, G. Monteleone, A.A. Lemonidou, Applied catalysis B : McGraw-Hill, New York, 2008.
environmental catalyst development for steam reforming of methane and model [51] B. Kumar, S. Kumar, S. Sinha, S. Kumar, Utilization of acetone-butanol-ethanol-
biogas at low temperature, Appl. Catal. B Environ. 181 (2016) 34–46, https://doi. water mixture obtained from biomass fermentation as renewable feedstock for
org/10.1016/j.apcatb.2015.07.039. hydrogen production via steam reforming: thermodynamic and energy analyses,
[32] L. Turchetti, M.A. Murmura, G. Monteleone, A. Giaconia, A.A. Lemonidou, Bioresour. Technol. 261 (2018) 385–393, https://doi.org/10.1016/j.biortech.2018.
S.D. Angeli, V. Palma, C. Ruocco, M.C. Annesini, Kinetic assessment of Ni-based 04.035.
catalysts in low-temperature methane/biogas steam reforming, Int. J. Hydrogen [52] B. Kumar, S. Kumar, S. Kumar, Thermodynamic analysis of H2production by oxi-
Energy 41 (2016) 16865–16877, https://doi.org/10.1016/j.ijhydene.2016.07.245. dative steam reforming of butanol-ethanol-water mixture recovered from
[33] P. Kowalik, K. Antoniak-jurak, M. Błesznowski, M.C. Herrera, M.A. Larrubia, Biofuel Acetone:Butanol:Ethanol fermentation, Int. J. Hydrogen Energy 43 (2018)
steam reforming catalyst for fuel cell application, Catal. Today 254 (2015) 6491–6503, https://doi.org/10.1016/j.ijhydene.2018.02.058.
129–134, https://doi.org/10.1016/j.cattod.2015.03.002. [53] S. Kumar, N. Katiyar, S. Kumar, S. Yadav, Exergy analysis of oxidative steam re-
[34] Q.T.P. Bui, Y. Kim, S.P. Yoon, J. Han, H.C. Ham, S.W. Nam, C.W. Yoon, Steam forming of methanol for hydrogen producton: modeling study, Int. J. Chem. React.
reforming of simulated biogas over plate Ni-Cr catalysts: influence of pre-oxidation Eng. 11 (2013) 489–500, https://doi.org/10.1515/ijcre-2012-0073.
on catalytic activity, Appl. Catal. B Environ. 166–167 (2015) 335–344, https://doi. [54] N. Hajjaji, M.N. Pons, A. Houas, V. Renaudin, Exergy analysis: an efficient tool for
org/10.1016/j.apcatb.2014.11.045. understanding and improving hydrogen production via the steam methane re-
[35] S. Ahmed, S.H.D. Lee, M.S. Ferrandon, Catalytic steam reforming of biogas - Effects forming process, Energy Policy 42 (2012) 392–399, https://doi.org/10.1016/j.
of feed composition and operating conditions, Int. J. Hydrogen Energy 40 (2015) enpol.2011.12.003.
1005–1015, https://doi.org/10.1016/j.ijhydene.2014.11.009. [55] J. Díez-Ramírez, F. Dorado, A. Martínez-Valiente, J.M. García-Vargas, P. Sánchez,
[36] J.M. Vásquez Castillo, T. Sato, N. Itoh, Effect of temperature and pressure on hy- Kinetic, energetic and exergetic approach to the methane tri-reforming process, Int.
drogen production from steam reforming of biogas with Pd-Ag membrane reactor, J. Hydrogen Energy 41 (2016) 19339–19348, https://doi.org/10.1016/j.ijhydene.
Int. J. Hydrogen Energy 40 (2015) 3582–3591, https://doi.org/10.1016/j.ijhydene. 2016.04.229.
2014.11.053. [56] O.Z. Sharaf, M.F. Orhan, An overview of fuel cell technology: fundamentals and
[37] P.S. Roy, C.S. Park, A.S.K. Raju, K. Kim, Steam-biogas reforming over a metal-foam- applications, Renew. Sustain. Energy Rev. 32 (2014) 810–853, https://doi.org/10.
coated (Pd-Rh)/(CeZrO2-Al2O3) catalyst compared with pellet type alumina-sup- 1016/j.rser.2014.01.012.
ported Ru and Ni catalysts, J. CO2 Util. 12 (2015) 12–20, https://doi.org/10.1016/ [57] U. Lucia, Overview on fuel cells, Renew. Sustain. Energy Rev. 30 (2014) 164–169,
j.jcou.2015.09.003. https://doi.org/10.1016/j.rser.2013.09.025.
[38] P.S. Roy, J. Song, K. Kim, C.S. Park, A.S.K. Raju, CO2conversion to syngas through [58] K. Kendall, Introduction to SOFCs, Elsevier Ltd, 2015, https://doi.org/10.1016/
the steam-biogas reforming process, 2 Util. 25 (2018) 275–282, https://doi.org/10. B978-0-12-410453-2.00001-4.
1016/j.jcou.2018.04.013. [59] A. Choudhury, H. Chandra, A. Arora, Application of solid oxide fuel cell technology
[39] A. Galvagno, V. Chiodo, F. Urbani, F. Freni, Biogas as hydrogen source for fuel cell for power generation - a review, Renew. Sustain. Energy Rev. 20 (2013) 430–442,
applications, Int. J. Hydrogen Energy 38 (2013) 3913–3920, https://doi.org/10. https://doi.org/10.1016/j.rser.2012.11.031.
1016/j.ijhydene.2013.01.083. [60] D.G. Avraam, T.I. Halkides, D.K. Liguras, O.A. Bereketidou, M.A. Goula, An ex-
[40] X.D. Peng, Analysis of the thermal efficiency limit of the steam methane reforming perimental and theoretical approach for the biogas steam reforming reaction, Int. J.
process, Ind. Eng. Chem. Res. 51 (2012) 16385–16392, https://doi.org/10.1021/ Hydrogen Energy 35 (2010) 9818–9827, https://doi.org/10.1016/j.ijhydene.2010.
ie3002843. 05.106.
[41] P.L. Cruz, Z. Navas-Anguita, D. Iribarren, J. Dufour, Exergy analysis of hydrogen [61] F. Cipitì, O. Barbera, N. Briguglio, G. Giacoppo, C. Italiano, A. Vita, Design of a
production via biogas dry reforming, Int. J. Hydrogen Energy 43 (2018) biogas steam reforming reactor: a modelling and experimental approach, Int. J.
11688–11695, https://doi.org/10.1016/j.ijhydene.2018.02.025. Hydrogen Energy 41 (2016) 11577–11583, https://doi.org/10.1016/j.ijhydene.
[42] B. Kumar, S. Kumar, S. Kumar, Thermodynamic and energy analysis of renewable 2015.12.053.
butanol–ethanol fuel reforming for the production of hydrogen, J. Environ. Chem. [62] V. Chiodo, A. Galvagno, A. Lanzini, D. Papurello, F. Urbani, M. Santarelli, S. Freni,
Eng. 5 (2017) 5876–5890, https://doi.org/10.1016/j.jece.2017.10.049. Biogas reforming process investigation for SOFC application, Energy Convers.
[43] M. Ashrafi, C. Pfeifer, T. Pröll, H. Hofbauer, Experimental study of model biogas Manage. 98 (2015) 252–258, https://doi.org/10.1016/j.enconman.2015.03.113.

13

You might also like