1 s2.0 S019689042100861X Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Energy Conversion and Management 246 (2021) 114685

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Energy-exergy analysis of an integrated small-scale LT-PEMFC based on


steam methane reforming process
Zaixing Wang a, Junkui Mao a, b, *, Zhenzong He a, b, Fengli Liang a, *
a
College of Energy and Power Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
b
Key Laboratory of Thermal Management and Energy Utilization of Aircraft, Ministry of Industry and Information Technology, Nanjing 210016, China

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

Keyword: This study reports a novel integrated system for power generation with on-site produced hydrogen and heat
LT-PEMFC recovery. The main equipment of the integrated system includes a steam methane reforming (SMR) reactor for
Steam methane reforming hydrogen production, a palladium membrane (Pd-M) unit for hydrogen purification, a waste heat collector (HC)
On-site hydrogen production
unit for heat recovery, and a low-temperature proton exchange membrane fuel cell (LT-PEMFC) for power
Thermodynamic performance analysis
generation. ASPEN Plus and MATLAB/Simulink jointed simulation model is used to establish an accurate model
for the whole integrated system. Moreover, for the SMR reactor, it is found that the RPlug reactor simulation
module is more appropriate compared with the RGibb module. Then a detailed parametric analysis is carried out
to estimate the effects of hydrogen utilization, the SMR reactor operating temperature, and the current density of
the PEMFC stack on the overall energy and exergy efficiency of the integrated system. Finally, thermodynamic
analysis is performed on the integrated system with S/C of 3.0, SMR reaction temperature of 700 ◦ C, PEMFC
hydrogen utilization rate of 0.8, and power generation of 520 W, delivering energy and exergy efficiencies as
high as 47.4% and 45.7% under these operating parameters. The thermodynamic analysis shows that the PEMFC
stack is the primary source of the system exergy destruction (326.27 W) with 34.33% of the relative exergy
destruction ratio.

reaction. Minutillo et al. [15] put forward two biogas-based poly-gen­


1. Introduction eration systems to produce high-pressure hydrogen, and the off-gas from
the combustor supplied the required heat for the two systems. Janneil
Low-temperature proton exchange membrane fuel cell (LT-PEMFC) et al. [9] investigated the performance of a microgeneration system with
technology has been spotlighted as a more effective electrochemical an SMR unit and a combustor providing the heat required for SMR
converter for power generation than other traditional power generation reactions.
technologies for its advantages: low noise, quick start-up, and high- How to separate H2 from the syngas of an SMR reactor to get high
power density [1–5]. LT-PEMFC is a riper and more commercialized pure H2 is another problem. The LT-PEMFC demands high-pure H2
fuel cell than other fuel cells [6–9] since it works at a lower temperature (99.999%) to operate because of the poor tolerance towards impurities
range from 60 to 80 ◦ C. in the fuel. However, the syngas produced from the SMR reaction not
However, LT-PEMFC has not been widely commercialized because it only have H2 but also contain a large number of by-products like carbon
relies on pure hydrogen as fuel. Hydrogen is traditionally stored in high monoxide (CO). What is worse, PEMFC has a low tolerance to CO, and a
pressure or cryogenic liquid due to its physical properties, resulting in a low CO (100 ppm) concentration can poison the anode catalyst of
high cost and limited storage density [10,11]. Using the steam methane PEMFC [8]. In the literature [12,16,17], palladium-based membrane
reforming (SMR) method to convert fossil fuel into hydrogen effectively (Pd-M) has high permeability and infinite selectivity for H2. As the
solves the hydrogen storage issue [12–14]. SMR process is a robust pressure differential of the Pd-M membrane increases, H2 permeability is
endothermic reaction requiring a continuous hot source to supply heat improved. Hence, the syngas should operate at high pressure to provide
to an SMR reactor. Therefore, effectively and continuously providing sufficient pressure differential for H2 to permeate through. This excep­
thermal for the reaction is one of the most significant problems. tional property of Pd-M makes the whole system operate under high
The external combustor can supply the heat required for an SMR pressure

* Corresponding authors at: College of Energy and Power Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China (J. Mao).
E-mail address: mjkpe@nuaa.edu.cn (J. Mao).

https://doi.org/10.1016/j.enconman.2021.114685
Received 24 May 2021; Accepted 22 August 2021
Available online 3 September 2021
0196-8904/© 2021 Elsevier Ltd. All rights reserved.
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

Nomenclature Ncell Cell numbers, 6


ηel /ψ el Gross electrical energy/exergy efficiency
Abbreviation ηCHH /ψ CHH Overall energy/exergy efficiency
HC Heat collector unit P Pressure (bar)
LT-PEMFC Low-temperature proton exchange membrane fuel cell Δs Molar entropy change
Pd-M Palladium-based membrane Tst The PEMFC stack operating temperature (℃), 65
S/C Steam-to-carbon, 3 μH2 H2 utilization ratio, 0.8
SMR Steam methane reforming μO 2 O2 utilization ratio, 0.5
WGS Water-gas shift XCH4 CH4 conversion rate
YH2 H2 yield
Symbols
εex,i Exergy efficiency
Acell Active area of the cell (cm2), 280
θIr,i Exergy destruction ratio
Ex Exergy (W)
θ∗Ir,i Relative exergy destruction ratio
Δgf0 Standard molar Gibbs free energy change

Nevertheless, CH4 conversion and H2 yield are influenced intensively [23] studied the effect of the catalyst loading on SMR process, and the
by the operating pressure of an SMR reactor. Shagdar et al. [18] inves­ results demonstrated that a catalyst loading of 4 mg/cm2 at 400 ◦ C
tigated the impact of operating pressure on the SMR reaction under achieved 97% CH4 conversion. However, when the loading decreased to
1.02–30 bar, and the results showed that CH4 conversion decreased from 2 mg/cm2, the operating temperature needed to be raised to 1000 ◦ C to
99.67% to 61.22%. Fausto et al. [19] researched the influence of attain the same conversion degree. Hence, CH4 conversion could be
working pressure on CH4 conversion in a membrane reactor employing improved by increasing the catalyst loading at a lower temperature.
SMR technology, and the simulation results showed that an increase in In fact, for the study on the influencing factors of CH4 conversion,
the operating pressure corresponded to a decrease in CH4 conversion. many researchers ignored the effect of catalyst loading and focused on
The above analysis indicates an inconsistency between Pd-M and the influence of operating reaction temperature and pressure. For
SMR reactor: the former requires high pressure to keep high perme­ example, Ye et al. [25] presented a simulation model of hydrogen pro­
ability for H2, whereas the latter requires low pressure to get high CH4 duction via RGibbs submodule in Aspen Plus. But compared with the
conversion. Therefore, this study focuses on coordinating the relation­ RGibbs reactor, the RPlug reactor module in Aspen Plus software can
ship between CH4 conversion in the SMR reactor and H2 permeability in better meet the requirement for establishing the catalyst loading model.
Pd-M under high operating pressure from a modeling point of view. Moreover, in the RPlug reactor module, a mathematical model of the
Actually, CH4 conversion can be improved by increasing the oper­ reforming process is established based on the Langmuir-Hinshelwood-
ating temperature and the catalyst loading in the SMR reactor [20–23]. Hougen-Watson (LHHW) methodology [20] that reveals the reaction
To increase the operating temperature, Falco et al. [24] analyzed the kinetics of the SMR process in detail [21]. For this reason, the RPlug
effect of operating temperature on the thermal power required for the reactor is relatively more precise and closer to the experimental results.
SMR process, and simulation results showed that the total heat power The above-mentioned is aimed at the core components only, but a
increased from 770.7 kW to 1033.4 kW in a reaction temperature range reasonable system is needed to connect each component to realize the
of 400–600 ◦ C. Thus, the increasing operating temperature can raise CH4 high efficiency of the system. Jannelli et al. [9] constructed a micro-
conversion and total heating power but poses a challenge to the heat cogeneration system that consisted of a natural gas steam reforming
supply for the whole system. On the contrary, increasing the catalyst unit, a catalytic burner, and a power unit, but the off-gas was exhausted
loading is a relatively more straightforward method. Arzamendi et al. directly into atmosphere. Actually, the heat from the off-gas could be

Fig. 1. The micro-PEMFC based integrated steam methane reforming system. (SMR: Steam methane reformer; H/L-WGS: High/Low temperature-water gas shift
reactor; HX: Heat exchanger; CL: Cooler; Cp: Compressor; P: Pump).

2
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

Table 1 2. System description and modeling


Operating and design parameters used in the integrated system.
Parameters Units Value 2.1. System configuration
The inlet fuel temperatureTfuel ℃ 25
The methane reformate gas-fueled PEMFC system consists of the
The inlet air temperatureTair ℃ 25
SMR subsystem and the power generation subsystem. Fig. 1 shows the
The inlet water temperatureTH2 O ℃ 25
configuration of the integrated system considered in this research. In
Operating temperature of the SMR ℃ 700
order to explain the operation process of the integrated system more
Operating temperature of the H-WGS ℃ 450
directly, several units are omitted in this configuration, such as the
Operating temperature of the L-WGS ℃ 300
values and the mass controllers. As depicted in Fig. 1, thermal integra­
Operating temperature of Pd membraneTPd ℃ 300
tion is provided by mainly adopting the combustor off-gas to preheat all
Steam-to-carbon ratioS/C – 3
inlet streams and partially using a Pd-M membrane outlet to preheat the
Operating pressure of Pd membranePPd bar 10
water inlet.
Catalyst type of steam reforming – Ni/Al2O3
In the SMR subsystem, the preheated methane and the steam are sent
to the SMR reactor, where an endothermic reaction occurs and produces
recovered and used to preheat the input fuel. Sayar et al. [26] studied a the syngas, consisting of CH4, H2O, H2, CO, and CO2. And the high-
methane fuel processor system that included a catalytic burner, evapo­ temperature combustion off-gas provides the required reaction heat.
rator, and auto thermal reformer (ATR). The external burner supplied The working temperature is maintained at 700 ◦ C, and the S/C rate is
the heat required by the reforming system and the off-gas preheated all assumed to be 3. To further improve the H2 yields and decrease the CO
inlet feedstock. Through the externally heat-integrated technology, the amount, the syngas is sent to a high temperature water gas shift (H-
system efficiency could be up to 86.3%. WGS) reactor and a low temperature water gas shift (L-WGS) reactor.
Through the above analysis, increasing the catalyst loading and Then, in the Pd-M unit, the hydrogen-rich gas is separated into two
recycling the combustor off-gas may be a great way to improve the fluxes: the first (F11) is conveyed to thecombustor as fuel, and the
system efficiency. Thus, this study presents a LT-PEMFC system inte­ second (F10) is 99.999% high purity hydrogen sent to the anode of
grated with an SMR reactor, adopting the combustor off-gas to preheat PEMFC. In this research, the SMR reactor operates at 10 bar to provide
feedstock and evaluating the thermodynamic performance of the inte­ sufficient differential pressure for the hydrogen to permeate through the
grated system through an exergy analysis tool. First, the integrated Pd-M membrane.
system is measured using thermodynamic calculation (including energy Water (W1) is firstly pumped to the No.3 cooler (CL-3) to exchange
and exergy analysis) through ASPEN Plus and MATLAB/Simulink plat­ heat with the high-temperature hydrogen and then with the reformed
form jointed simulation model. Then, a more appropriate model for this gas at the outlet of WGS. The preheated water changes into steam after
study is selected by comparing the performance of the RGibbs and RPlug heat exchange with the combustor off-gas (C1) through HX-2. This
reactor. Next, based on the RPlug reactor module, the effect of the combustor is fed with Pd-M membrane exhaust gases, dry air, the
catalyst loading on CH4 conversion is investigated carefully. Finally, exhaust gases from the PEMFC stack, and a small amount of fuel. Finally,
thermodynamic analysis is performed on the integrated system with S/C the exhausts (C6) are sent to the heat collector (HC) to produce helpful
of 3.0, SMR reaction temperature of 700 ◦ C, PEMFC hydrogen utilization heat. Table 1 provides the parameters of the operation and configuration
rate of 0.8, and power generation of 520 W. of the integrated system.
Fig. 2 shows the simulation flow diagram for the cogeneration sys­
tem. The material flows adopt the positive direction analysis method to
combine the two kinds of flows, whereas the energy flows adopt the

Fig. 2. Simulation flow diagram for the integrated system.

3
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

reverse analysis method. For example, the target electrical power is first
XNin2 − XNout2
used as the input parameter for the material flows analysis. Then the XNchannel = (10)
power generation subsystem calculates the required hydrogen flow
2
ln(XNin2 /XNout2 )
based on the input. Finally, the hydrogen production subsystem calcu­ Here, parameters XHsat2 O and XN2 denote the mole fractions of H2 O and
lates the required fuel flows. Electrical power is considered as the output
N2 , λair stands for the air stoichiometric ratio.
parameter for the energy flows analysis. The integrated system simula­
Vact , the activation loss voltage occurs at both PEMFC electrodes.
tion using Aspen Plus and MATLAB/Simulink co-simulation approach and
Based on the investigation of references, the authors adopt the equation
the key parameters are presented in the following figure.
published by Jay et al. [29]:
Vact = V0 + Va (1 − exp(− c1 i)) (11)
2.2. Fuel cell subsystem
where V0 is the voltage drop at zero current density.
2.2.1. Electrochemical model of the PEMFC stack
Here, on the basis of previous work [1], combined with relevant V0 = 0.279 − 8.5 × 10− 4 (Tst − 298.15) + 1.4454 × 10− 7 Tst
(12)
references, a relatively accurate mathematical model is established, × [ln(Pca − Psat sat
H2 O ) + 0.5ln(0.1173(Pca − PH2 O ))]
which is closer to the experimental results.
So as to build up the mathematical model, the following general
2
Va = (− 1.618 × 10− 5 Tst + 1.618 × 10− 2 )(PO2 /0.1773 + Psat
H2 O )
assumptions are put forward: + (1.8 × 10 Tst − − 4
0.166)(PO2 /0.1773 + Psat (13)
H2 O )

+ (− 5.8 × 10− 4 Tst + 0.5736)


• All gases behave ideally.
• The PEMFC stack is in steady-state operation condition. i is the operating current density,
• The water vapor is assumed to saturation.
• The temperature is distributed equally inside the stack, which is i=
Ist
(14)
equal to the coolant outlet temperature. Acell
• Pressure drop inside the stack is ignored. The ohmic overpotential Vohm can be expressed as:

The single cell output voltage Vcell can be determined as below: Vohm = iRint (15)

Vcell = ENerst − Vloss (1) The function relationship between internal electrical resistance Rint ,
membrane thickness tm , and membrane conductivity σ m is as follows:
Vloss = Vact + Vohm + Vcon (2) tm tm
Rint = =
σm (0.005139λm − 0.00326) × exp(350 × (1/303 − 1/Tst ))
in which ENerst is the Nernst voltage, also called reversible voltage; the
(16)
voltage loss Vloss is made up of three parts: the activation overpotential
Vact , the ohmic polarization Vohm , and the concentration polarization At last, the concentration overpotential Vcon is shown as below:
Vcon . ( )c3
i
The Nernst potentialENerst is defined as below [27]: Vconc = i c2 (17)
imax
( √̅̅̅̅̅̅̅̅ )
Δg0f Δs Rg Tst P H 2 P O2
ENerst = − − (Tst − 298.15) + ln (3) where
2F 2F 2F P H2 O



⎪ (7.16 × 10− 4 Tst − 0.166)(PO2 /0.1773 + Psat
H2 O )
where PH2 , PO2 and PH2 O are the partial pressure of H2 , O2 and H2 O in unit ⎪



of bar, which are as follows [28]: ⎪
⎨ + (− 1.45 × 10− 3 Tst + 1.68 ) if PO2 /0.1773 + Psat
H2 O < 2 atm
⎡ ⎤ c2 =

⎪ (8.66 × 10 Tst − 0.068)(PO2 /0.1773 + Psat
− 5

⎪ H2 O )
⎢ ⎥ ⎪

⎢ ( 1) ⎥ ⎪
(4) ⎪ − 4
PH2 = 0.5Psat
H2 O ⎢ − 1⎥ ⎩ + (− 1.6 × 10 Tst + 0.54 ) else
⎣ ⎦
exp 1.653i
T 1.334
X sat
H2 O
st c3 = 2, imax = 2.2
(18)
( ( ))
0.291i
PO2 = Pop 1 − XHsat2 O − XNchannel exp 10.832 (5) In which, the relation between stack output electricity Pst , the cur­
2
Tst
rent density i, the single cell voltage Vcell , and the number of cells Ncell can
be expressed as:
log(Psat
H2 O ) = − 2.1749+0.0295(Tst − 273.15)− 9.1837 × 10
− 5
×(Tst − 273.15)2
Pst = Ncell iAcell Vcell (19)
+1.4454 × 10− 7 ×(Tst − 273.15)3
(6) Thermodynamic model of the PEMFC stack
We have detailed the thermal model by the thermodynamic principle
where in our previous research work [1]. Therefore, we will only give the vital
Psat thermodynamic equations as follows.
(7) By thermodynamic energy balance, the PEMFC stack thermal model
H2 O
XHsat2 O =
Pop
can be calculated as:
XNin2 = 0.79(1 − XHsat2 O ) (8) dTst H˙in − Ḣ out − Q˙cl − Pst
= (20)
dt mst Cp,st
1 − XHsat2 O
XNout2 = ( )( ) (9)
1+ 0.21 λair − 1 Ḣ in = Ḣ in,an + Ḣ in,ca (21)
0.79 λair

where Ḣin,an , Ḣin,ca are the input enthalpy of the anode and cathode,

4
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

respectively.
where Gt is the total Gibbs free energy, G0i is the standard Gibbs free
Ḣ out = Ḣ out,an + Ḣ out,ca (22) energy of species i, ni is the mole of species i, fi0 is the standard-state
fugacity, fi is the fugacity at operating conditions.
where Ḣout,an , Ḣout,ca are the anode and cathode output enthalpy,
respectively. 2.3.2. RPlug reactor module
The cooling fan removes the generated heat from the stack to In this work, the reaction kinetics are shaped by Xu et al. [22], and
maintain the stack operating at a proper temperature. Q˙cl , the heat flow the corresponding commercial catalyst is Ni/Al2O3 [21]. In this paper,
dissipated by the cooling fan can be expressed as: the RPlug reactor module is developed based on the following
assumptions:
Q˙cl = nair cp,air (Tst − T1 ) (23)
• The porosity of the catalyst bed is constant, and the particle size of
where nair is the air coolant mole flow rate, and T1 is the air temperature. the catalyst is uniform.
dTst /dt is the temperature change with respect to time. When the • The reactor operates at a constant temperature, and the internal
temperature is no longer changing with time, the temperature value Tst temperature is evenly distributed.
is the operating temperature of the PEMFC stack. The average value of • Do not consider the impact of carbon deposition.
mst Cp,st is about 35 kJ/K. • Ignore the reactor pressure drop.
The reacted air and hydrogen are defined as:
nIst The kinetic rate equations and kinetic parameters utilized in the
nreacted
H2 = (24) RPlug reactor module are shown below, respectively,
2F
( )
nIst k1 PCH4 PH2 O - PCO P3H2 /Ke1
nreacted = (25) R1 = (32)
O2
4F P2.5
H2 DEN 2
The hydrogen utilization ratio and oxygen utilization ratio are also ( )
estimated from Eqs. (26) and (27), respectively [30]: R2 =
k2 PCO PH2 O - PCO2 PH2 /Ke2
(33)
P H2 DEN 2

nreacted
(26)
H2
μ H2 = ( )
nsupplied k3 PCH4 P2H2 O − PCO2 P4H2 /Ke3
(34)
H2
R3 = 3.5
P H2 DEN 2
nreacted
(27)
O2
μ O2 =
nsupplied KH2 O PH2 O
O2
DEN = 1 + KCO PCO + KH2 PH2 + KCH4 PCH4 + (35)
P H2
According to the voltage and thermal equations mentioned above,
zero-dimensional modeling for the PEMFC subsystem is established by here, the partial pressure (Pj ) and the adsorption coefficient (Kj ) of jth
the MATLAB/Simulink platform. component (j = CH4, H2O, CO2, CO, H2); the reaction rate coefficients
(ki ) and the equilibrium constant (Kei ) of ith reaction (i = 1, 2, 3). The
2.3. Steam methane reforming subsystem

SMR process can be summarized as follows [24]: Table 2


Values of parameters appearing in the kinetic reactions [21].
SMR: CH4 + H2 O ↔ CO + H2 ΔH298 = 206 kJ/mol (28)
Parameters Units Value

WGS : CO + H2 O ↔ CO2 + H2 ΔH298 = - 41 kJ/mol (29) Reaction Parameters


k0i k01 mol Pa0.5 / (kg s) 5.79 × 1016

global SMR : CH4 + 2H2 O ↔ CO2 + 4H2 ΔH298 = 165 kJ/mol k02 9.33 × 107

(30) k03 1.29 × 1017


Ei E1 kJ / mol 217.01
Eq. (28) represents SMR reaction, which is a highly endothermic E2 68.2
reaction; SMR reaction is accompanied by a side reaction, water gas shift E3 215.84
(WGS) reaction (Eq. (29)), that occurs at a lower temperature. Eq. (30)
means the global SMR process is a robust endothermic procedure, Adsorption Parameters
K0j K0CH4 Pa− 1 66.6
requiring a continuous hot source to supply heat to the SMR reactor.
K0H2 O 1.77 × 1010
This chemical reaction process takes place in the SMR reactor, and
Aspen Plus software is used in this research to model this process. The K0H2 6.15 × 10− 4

modeling equations for the other significant components used in Aspen K0CO 8.25
Plus are given in Appendix. K0CO2 –
ΔHj ΔHCH4 kJ / mol − 38.28
2.3.1. RGibbs reactor module ΔHH2 O 88.68
The RGibbs reactor conducts the Gibbs free energy minimization ΔHH2 − 82.90
method for determining the concentration of the syngas. The lower of ΔHCO − 70.65
the total Gibbs free energy of the system indicates that the mixture of ΔHCO2 –
chemical substances is close to thermodynamic equilibrium. For the Equilibrium constants
entire chemical reaction process, the total Gibbs energy is multiplied by Kei Ke1 Pa2
(
26830
)
1.198 × 1023 exp −
the sum of i-th components [31,32]: (
T
)
Ke2 4400
1.767 × 10− 2 exp

N ∑
N
fi T
t
G = ni G0i + R(T) ni ln 0 + ns Gs (31) Ke3
(
− 22430
)
fi 2.117 × 1021 exp
i=1 i=1
T

5
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

temperature dependence of ki and Kj is described by an Arrhenius type Table 3


function, and Table 2 sums up the parameters that appear in the kinetic Standard molar chemical exergy of selected substances [35,37,38].
reactions. Substance State Exergy (kJ/mol)
( )
Ei CH4 g 831.65
ki = k0i × exp − (36)
RT H2 O l 0.921
g 9.501
( )
ΔHj CO g 275.35
Kj = K0j × exp − (37)
RT CO2 g 20.162
H2 g 238.088
The SMR performances are evaluated based on CH4 conversion rate
N2 g 0.639
XCH4 and H2 yield YH2 , which are shown below:
O2 g 3.921
nCH4 ,in − nCH4 ,out
XCH4 (%) = (38)
nCH4 ,in
where h and s represent the specific enthalpy (kJ/mol) and entropy (kJ/
nH2 ,out
YH2 (%) = (39) mol K), and the subscript zero means the standard environment state.
nCH4 ,in
The specific chemical exergy exch of a substance is determined by:
∑ ∑
where nCH4 ,in , nCH4 ,out and nH2 ,out correspond to the inlet CH4 flow rate, exch = n
χ n ⋅ exch + χ n ⋅ lnxn (46)
outlet CH4 and H2 flow rate in mol/min, respectively.
where χ n is the mole fraction of a component, exnch is the standard
3. Thermodynamic evaluation of the integrated system chemical exergies. The standard chemical exergies of the species n are
listed in Table 3.
In this part, the fundamental definition of exergy and the related
conservation equations are first introduced. Then, based on this theo­
3.2. Performance evaluation parameters of the system
retical analysis, the thermodynamic mathematical model of exergy
analysis is established. Finally, according to the thermodynamic math­
The integral system is evaluated according to different efficiency
ematical model, some integrated system performance evaluation pa­
parameters (i.e., electrical energy and exergy efficiency). The gross
rameters can be defined (i.e., electrical and exergy efficiencies).
electrical energy efficiency ηel and the overall energy efficiency ηCHH
(CHH, combined electrical and heat energy) are expressed as
3.1. 3.1 Exergy analysis of the system [34,39,40]:
Wnet
When performing exergy analysis, the first step is to define the ηel = (47)
(nCH4 ,Ref + nCH4 ,CB ) ⋅ LHVCH4
standard environmental state. In this research, the standard environ­
ment state is fixed at 25 ◦ C and 1 bar.
Wnet + QHC
At present, for a control volume, the exergy balance equation is ηCHH = (48)
(nCH4 ,Ref + nCH4 ,CB ) ⋅ LHVCH4
commonly expressed as [3]:
Exin = Exout + Exheat + Woutput + Ir (40) Wnet = Pst − PBOP (49)

where Exin and Exout are the total of the mole flow exergies of all input where nCH4 ,Ref and nCH4 ,CB are the mole flow rate of methane to the
and output fluids, Exheat and Woutput are the exergy transfer by heat and reformer and the combustor, LHVCH4 is the low heating value of CH4
work, Ir displays the total exergy destructions or the irreversibilities. (802.3 kJ/mol), QHC is the heat recovered by the heat collector, and PBOP
For the exergy balance, the following equation is also used in some is the total power consumed by the balance of the plant components (i.
literature [33,34]: e., compressor, pump).
The gross electrical exergy efficiency ψ el and the overall exergy ef­
Exin - Ex,out + ExQ − ExW − I = 0 (41) ficiency ψ CHH are given as [37]:

where, ψ el =
Wnet
(50)
(nCH4 ,Ref + nCH4 ,CB ) ⋅ exCH4
ExQ is deduced from the Carnot cycle. When Carnot engine works
between high temperature T and low temperature T0 , the heat source Wnet + QHC
ψ CHH = (51)
absorbed by Carnot engine is Q: (nCH4 ,Ref + nCH4 ,CB ) ⋅ exCH4
( )
T0 where exCH4 is the standard chemical exergy of CH4 (831.65 kJ/mol).
ExQ = 1 − Q (42)
T The exergy efficiency εex,i , exergy destruction ratio θIr,i , and relative
exergy destruction ratio θ∗Ir,i for an ith component are introduced to show
the exergy destruction of each component more clearly. [15,41]:
ExW represents the exergy transfer through work:
Q
Exiout + Exin W
i + Exi + Exi
ExW = Wnet (43) εex,i = (52)
in
Exi
The total exergy transfer Ex is calculated using the following form
[15,35,36]: Ir,i
θIr,i = (53)
Exin
Ex = n⋅ex = n⋅(exph + exch ) (44) Fuel

The specific physical exergy exph of a substance is shown as: θ∗Ir,i =


I
∑r,i (54)
Ir,i
exph = (h − h0 ) − T0 (s − s0 ) (45)

6
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

Fig. 3. The research procedure for the integrated system in this work.

Fig. 4. Comparison polarization curves between simulated results and experimental value at (a) varying pressure [29] and (b) varying temperature [42].

where Ir,i is the destruction of a component, Exin


Fuel is the total exergy of
4.1.1. Validation of PEMFC model

fuel flowing into the integrated system, and Ir,i is the total exergy As aforesaid, it is necessary to verify the established PEMFC math­
destruction. ematical model because PEMFC is a complicated system. Under different
To better display the structure of this paper, the research procedure operating conditions, such as different pressures and temperatures,
is shown in Fig. 3: whether the simulated polarization curve is consistent with the experi­
mental results can verify the PEMFC model.
1) Firstly, the hydrogen flow rate is determined according to the target As shown in Fig. 4(a), the polarization curves of PEMFC simulation
electrical power of the PEMFC stack, which is also the target results and experimental value published by Jay et al. [29] with various
hydrogen production rate of the SMR reformer; operating pressures (1–5 bar) are compared. The figure shows that the
2) Secondly, when the integrated system adopts the Pd-M unit, the single cell voltage Vcell with current density i displays similar trends
working pressure of SMR is high; between the present and experimental results published by the
3) Then, compared with the RGibbs reactor, the RPlug reactor module researcher. For example, the cell voltage boosts continuously as the
in Aspen Plus software can better meet the requirement for estab­ operating pressure rises when the current density remains constant. By
lishing the catalyst loading model. Thus increasing the catalyst contrast, the polarization curve constantly decreases as the current
loading on the RPlug module is a simple way to get high CH4 con­ density increases when the operating pressure remains. Besides the
version, small CH4 inlet flow rate, and low heat loss of the system; similar trends, the figure also indicates the simulation and experimental
4) Finally, the efficiency of the integrated system under these operating value tally very well in error permissible.
parameters is studied in detail. Furthermore, the same trend is also obtained in Fig. 4(b). To further
verify the accuracy of the PEMFC model, a comparative analysis of the
4. Results and discussion simulation and experiment results from F.L et al. [42] are performed.
The polarization curves of PEMFC results under three different tem­
4.1. Model validation peratures (39 ◦ C, 56 ◦ C, and 72 ◦ C) are investigated. According to the
comparison results, the changing trend of the voltage polarization
In this part, the experimental results published by other researchers curves at different temperatures is the same as that at various pressures.
are cited to verify the simulation reliability of the critical elements, such However, the OCV (Open Circuit Voltage) from the experimental is
as the PEMFC stack and the SMR reactor, to lay a reliable base for the higher than the simulation results. At 72 ◦ C, for instance, the value of
results and discussion. OCV under experimental conditions is about 1.05 V, while the value of
simulation is less than 1.0 V. The reason could be that the voltage losses

7
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

Fig. 5. Comparison of modeling results with experimental value [21]: (a) reactor outlet flowrate of methane; (b) methane conversion; (c) reactor outlet flowrate of
hydrogen; (d) hydrogen yield.

of the PEMFC mathematical model in the present work are established condition of a high CH4 inlet flow rate and low reforming temperature,
through the theoretical and empirical formulas, and these formulas are the average error is less than 3% within the acceptable range (less than
idealized and simplified. For example, the type and thickness of the 10%), so the reactor model in the present paper is suitable and suffi­
proton exchange membrane used in the experiment are not given in the ciently well validated.
literature, so the simulation values of these two parameters can only
take empirical values. 4.2. Determine the major parameters
In the end, with the above analysis, the two figures indicate that the
simulation results agree with the experimental value well, and the 4.2.1. Determination of the hydrogen production rate
feasibility of the proposed PEMFC model is further verified. The methane reformate gas-fueled PEMFC system consists of two
parts: the SMR and PEMFC electricity generation subsystems. The power
4.1.2. Validation of RPlug reactor module generation subsystem determines the hydrogen flow rate of PEMFC and
The SMR reactor is numerically solved in Aspen Plus using the RPlug ultimately determines the reforming gas flow rate required by the re­
reactor module by the LHHW methodology. The LHHW kinetic expres­ form subsystem.
sion is based on the reference [22], and the rate expressions for the SMR The expected net electric power of the integrated system is 500 W;
reactions are given in Eqs. (32)–(35). Table 2 sums up the parameters for however, considering the power-consuming components (i.e., pumps),
the kinetic reactions over Ni/Al2O3 catalyst [21], and the reformer the generating capacity of the PEMFC stack is set as 520 W in this study.
operates at varying inlet flow rates and temperatures. Based on the mathematical model in Section System configuration, the
The simulation results and experimental value in the reference are operating parameters are listed in Table 4; in the range of 0.0–1.6 A/
presented in Fig. 5. The dashed lines correspond to the simulation re­ cm2, the cell voltage Vcell and output power Pst curves versus the current
sults, while the points represent the reference paper. From Fig. 5, the density i are shown in Fig. 6. For example, at 1.15 A/cm2, the PEMFC
errors between the experiment and simulation results are minimal,
especially under the condition of a high CH4 inlet flow rate and low Table 4
reforming temperature. Notably, at 616.85 ◦ C, the error decreases Operating parameters used in the PEMFC stack modeling.
gradually as the inlet flow rate is increased. The reason for the slightly
Parameters Units Value
larger error at the low flow rate is that, in the experimental environ­
ment, the lower the inlet flow rate is, the longer the contact time be­ Active area of cellAcell cm2 280
tween the reaction gas and the catalyst is, and the more sufficient the Cell numbersNcell – 6
reforming reaction is, so both CH4 conversion XCH4 and H2 yield YH2 are Hydrogen utilization ratioμH2 – 0.8
high. Unfortunately, the simulation model cannot describe the contact Oxygen utilization ratioμO2 – 0.5
time between the catalyst and the reaction steam. Nonetheless, the Operating pressurePop bar 1.1
changing trend of simulation and experiment is consistent. Under the Operating temperatureTst ℃ 65

8
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

Fig. 6. Curves of the cell voltage Vcell and output power Pst versus the current
Fig. 8. Effective of reaction pressure on CH4 conversion and H2 yield for LHHW
density i.
model and Gibbs model.

stack reaches the peak output power of 924.14 W, and the corresponding example, for the Gibbs model, with the reaction pressure changes from 1
single cell voltage is 0.478 V. With a current density of approximately to 10 bar, the inlet CH4 flow rate changes from 0.075 to 0.113 mol/min
0.477 A/cm2, a 522.2 W of the electrical power can be further obtained, and XCH4 decreases from 97.2% to 61.6%, respectively. Similar obser­
which is appropriate for the target power of 520 W. Correspondingly, vations have been reported previously [18,19], and the reason why the
the reacted hydrogen flow rate is 0.2491 mol/min, which is calculated SMR reaction is sensitive to the reaction pressure was reported by Xu
by Eq. (24). et al. [22]. That is because the high reaction pressure has an adverse
impact on the forward reaction rates. The pressure rise reduces the
4.2.2. Determination of the reactor module amount of reaction products, which makes XCH4 decreases. Subsequently,
As mentioned above, there are two different reactor modules in XCH4 and YH2 rapidly decrease as the reaction pressure is augmented.
Aspen Plus software, the RGibbs, and the RPlug module. The former Moreover, the two models have similar performance trends with the
adopts Gibbs free energy minimization methodology, while the latter pressure change in Figs. 7 and 8; besides, CH4 conversion of the LHHW
uses LHHW methodology (hereafter called the Gibbs model and the model is always less than that of the Gibbs model at any pressure point.
LHHW model). By comparing the performance of the two reactors under This difference is because the Gibbs model is an ideal model which
different pressures and temperatures, the appropriate model for this adopts the Gibbs free energy minimization theory. Contrarily, the LHHW
study is determined. model is established on the kinetic rate equations. Different kinds of
To directly produce high purity hydrogen for the PEMFC stack, the catalysts show various activities, resulting in diverse kinetic rate equa­
integrated system employs the Pd-M reactor, which operates at high tions and different reforming performances. The kinetic rate equations
pressure. For that reason, Figs. 7 and 8 display the performance of two adopted in this study are reported by Oliveira et al. [21], and the
type reactors at different reaction pressures, varying from 1 to 10 bar. mathematical model presents a lower XCH4 by comparing with the report
Both models operate at a S/C feed molar ratio of 3:1, the target hydrogen of Xu et al. [22]. However, the performance could be improved by
production flow rate of 0.2491 mol/min, and the reaction temperature increasing the catalyst loading along the gas flowing direction.
of 700 ◦ C. Under the basic parameters set in this study, to achieve the The Pd-M reactor operates at high pressure, and the pressure is set as
target hydrogen production, the inlet CH4 flow rate increases along with 10 bar in this study. As seen in Fig. 9, the effect of the catalyst loading on
the pressure increases, leading to the decreases of XCH4 and YH2 . For

Fig. 7. Effective of reaction pressure on methane inlet flow rate for LHHW Fig. 9. Effective of the catalyst loading on methane conversion and methane
model and Gibbs model. inlet flow rate for the LHHW model.

9
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

the SRM reaction is examined adopting the LHHW model at a reaction appropriate design conditions to discuss further the integrated system
pressure of 10 bar and a reaction temperature of 700 ◦ C. The catalyst operation; parameters such as the hydrogen utilization factor of the
loading increases from M to 10 M, XCH4 increases from 60.8% to 96.9%, PEMFC stack and the reactor operating temperature are analyzed. In the
and nCH4 ,in decreases from 0.128 to 0.082 mol/min, respectively. How­ following parts, the implications of these principal parameters of the
ever, at sufficiently high SRM catalyst loading, the increase in XCH4 and system are investigated.
the decrease in nCH4 ,in are not apparent due to the limitation of chemical
equilibrium to endothermic reactions. A similar trend is also obtained by 4.3.1. Effect of hydrogen utilization
Arzamendi et al. [23]. Increasing the catalyst loading from 6 M to 10 M, The H2 consumption rate is fixed at 0.2491 mol/min as the electric
XCH4 increases by about 5% and nCH4 ,in decreases by around 4%, power output is constant at 520 W. Thus, the change of H2 utilization of
respectively. These results indicate that the performance of the LHHW the PEMFC stack could affect the parameters of CH4 feed rate and un­
model is not changed remarkably in this catalyst loading range. converted H2 for the combustor. As shown in Fig. 11, the improvement
Consequently, the catalyst loading for the LHHW model is fixed at 6 M in of H2 utilization μH2 reduces the CH4 feed to be supplied for the reformer
the following study. subsystem. From Eq. (26), if the H2 required by the PEMFC stack re­
Fig. 10 illustrates the impact of reaction temperature on XCH4 of three mains constant, the increase of μH2 will lead to the decrease of the H2
models: the LHHW model with a catalyst loading of M (the LHHW supply rate. Similar trends are consistent with those reported by Pur­
model-M), the LHHW model with a catalyst loading of 6 M (the LHHW nima et al. [43]. One can further obtain that the lower μH2 in the PEMFC
model-6 M), and the Gibbs model. Considering the reaction pressure of stack, the more unused H2 is supplied to the combustor. Constantly, in
10 bar, three models display a similar movement: XCH4 rapidly increases the case of high μH2 , additional CH4 is required to be supplied to the
as the reaction temperature is augmented first, then tends to be stable, combustor to provide extra heat needed by the integrated system. And
and approaches 100% at the end. The reason is that endothermic re­ an additional CH4 supply that starts at μH2 of 0.5 is also shown in Fig. 11.
actions are limited by chemical equilibrium, as reported in the literature As a result, the overall efficiency (ηCHH /ψ CHH ) and gross electrical effi­
[24]. When the reaction temperature is fixed, Figs. 7–9 show that the ciency (ηel /ψ el ) are compared and shown in Fig. 12. With the improve­
higher XCH4 becomes, the smaller nCH4 ,in is. To reduce nCH4 ,in , XCH4 of 90% ment of μH2 , both kinds of electrical efficiency are increased, tardily from
is chosen as the judgment criterion (the blue dotted line in Fig. 10). 25.8% to 32.7% for energy efficiency and from 24.8% to 31.5% for
Based on this criterion, the lowest temperatures in the three models are exergy efficiency.
about 700 ◦ C, 750 ◦ C, and 820 ◦ C, respectively (the special blue symbol). Similarly, with the improvement of μH2 , the overall energy efficiency
From the simulation results, XCH4 reaches 92% for the LHHW model-6 M
at 700 ◦ C; for the LHHW model-M and the Gibbs model, the operating
temperatures have to be raised to about 750 ◦ C and 820 ◦ C to attain this
conversion rate. Because the SMR reaction is a strongly endothermic
reaction process, so the reactor needs an external heat source. If the
reactor operates at a higher reaction temperature, more heat needs to be
provided. Compared with the other two models, the LHHW model-6 M
requires a lower temperature.
Hence, it has been determined that the LHHW model is the best as the
reactor module through the above analysis. Simultaneously, the optimal
operating parameters for the model are obtained.

4.3. Evaluation of operating parameters effects

There are mutual coupling influences and constraints between the


design conditions of the SMR subsystem and the design parameters of
the power generation system. Therefore, it is necessary to choose
Fig. 11. The effect of hydrogen utilization of PEMFC on the fuel flow rate of the
integrated system.

Fig. 10. Effective of reaction temperature on methane conversion for the


LHHW model and the Gibbs model. Fig. 12. The system efficiency under various hydrogen utilization.

10
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

increases from 42.0% to 49.2%, and the overall exergy efficiency boosts But there is an interesting phenomenon: with the improvement of the
from 40.4% to 47.5%. Due to the energy recovery equipment, the waste operating temperature, both overall energy and exergy efficiency first
heat carried by the off-gascan be recovered, making the system effi­ decrease and then increase. In the temperature range from 640 ◦ C to
ciency boost about 17%. The overall efficiency shows the same trend 670 ◦ C, the overall energy efficiency drops from 48.1% to 44.5%, and
mainly due to the decrease in total CH4 supply and increases in μH2 . In exergy efficiency falls from 46.3% to 42.9%; on the contrary, the energy
this study, with the fixed total output power of PEMFC, the H2 pro­ efficiency improves from 44.5% to 48.5% and exergy efficiency rises
duction by the SMR process is reduced with the increase of μH2 . from 42.9% to 46.8% in the temperature range from 670 to 750 ◦ C,
Consequently, the CH4 feed for the SMR reformer system is also reduced. respectively. The reasons for this phenomenon are as follows: when the
operating temperature is below 670 ◦ C, the catalyst activity is low at this
4.3.2. Effect of the SMR reactor operating temperature temperature, resulting in low CH4 conversion. As Fig. 5 shows, the
Different reactor operating temperatures affect the catalyst activity maximum CH4 conversion is about 40% at the low-temperature range,
and then affect CH4 conversion. For example, from Fig. 13, the leading to a large amount of unconverted fuel in the syngas of the
improvement of reaction temperature reduces the CH4 supply for the reforming subsystem.
reformer system. This phenomenon results from a high SMR operating Moreover, when the temperature increases, CH4 conversion is
temperature, which means a higher fuel conversion, and less CH4 rapidly improved at the low-temperature range. Subsequently, the un­
needed to be feed to meet the H2 requirement. And the similar phe­ converted fuel is fed into the combustor and discharges a large com­
nomenon is reported by the reference [44]. At a temperature above bustion waste heat. Then, through the heat collector device, the waste
720 ◦ C, the flow rate of the CH4 supply becomes stable because the heat is recovered. In the low temperature range, the recovered heat
maximum conversion rate is achieved (see Fig. 10). Correspondingly, for power decreases rapidly, resulting in decreased overall efficiency as the
the gross electrical efficiency, both electrical energy and exergy effi­ operating temperature increases. On the contrary, the higher the tem­
ciency rapidly increase at first, tend to be stable, and finally reach 31.7% perature range is, the higher the CH4 conversion is. Afterward, the
and 30.7%, respectively, with the reaction temperature increasing as higher conversion makes a few unconverted fuels fed into the combustor
shown in Fig. 14. The reason for that is as follows: the H2 supply is and generates a relatively small combustion exhaust flow. Thus, the
constant because the total electric power of PEMFC is fixed at 520 W. As recovered thermal power changes slowly, but the gross electrical effi­
the operating temperature increases, CH4 conversion increases first and ciency gradually rises, improving total efficiency.
then stabilizes, resulting in the supply of CH4 decreasing first and then
remaining unchanged. 4.3.3. Effect of the PEMFC stack current density
The effect of changing the current density of the PEMFC stack on the
fuel supply flow rate, energy efficiency, and exergy efficiency is pre­
sented in Figs. 15 and 16. It can be seen that both the CH4 supply flow for

Fig. 13. The effect of reaction temperature of SMR reactor on the fuel flow rate.

Fig. 15. The effect of current density on the fuel flow rate.

Fig. 16. The performance of the integrated system under different cur­
Fig. 14. The system efficiency under different reaction temperatures. rent densities.
11
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

Table 5
The exergy performance of the integrated system (Units: W).
Component Exin
i
Exout
i ExQi ExW
i
Ir,i εex,i (%) θIr,i (%) θ∗Ir,i (%)

P1 3.95 4.21 – − 0.97 0.72 85.46 0.04 0.08


CP1 38.72 43.51 – − 8.23 3.44 92.67 0.20 0.36
CP2 0.65 0.73 – − 0.20 0.12 85.97 0.01 0.01
SMR 1833.82 1814.93 – – 18.89 98.97 1.11 1.99
H-WGS 1496.06 1470.01 − 17.64 – 8.41 99.44 0.50 0.89
L-WGS 1447.02 1443.46 − 3.37 – 0.19 99.52 0.01 0.02
Pd-M 1443.46 1416.37 – – 27.08 98.12 1.60 2.85
CB 925.94 614.24 – – 311.70 66.34 18.37 32.79
PEMFC 1278.00 255.92 − 175.81 520.00 326.27 74.47 19.23 34.33
HX-1 1752.75 1745.34 – – 7.41 99.58 0.44 0.78
HX-2 678.00 668.19 – – 9.80 98.55 0.58 1.03
HX-3 1842.34 1833.82 – – 8.52 99.54 0.50 0.90
HX-4 315.95 311.57 – – 4.39 98.61 0.26 0.46
Mixer 1315.58 1299.67 – – 15.92 98.79 0.94 1.67
HC 267.61 134.13 52.17 – 185.65 41.94 10.94 19.53
CL-1 1567.20 1559.82 – – 7.38 99.53 0.43 0.78
CL-2 1479.64 1471.73 – – 7.91 99.47 0.47 0.83
CL-3 1250.28 1243.67 – – 6.61 99.47 0.39 0.70
Sum 950.41 56.03 100.0
System level results 1739.79 134.13 − 144.654 510.6 950.40 ErrorΔIr,i Relative error (%)
0.09 0.01
Wnet + QHC
ψ CHH = = 45.7%
(nCH4 ,Ref + nCH4 ,CB ) ⋅ exCH4
Wnet + QHC
ηCHH = = 47.4%
(nCH4 ,Ref + nCH4 ,CB ) ⋅ LHVCH4

the SMR reformer and combustor increase with the raises of the current system are determined. In this section, the exergy analysis is applied to
density from 0.1 to 1.6 A/cm2 because the H2 demand rises as the cur­ assert the irreversibilities and the losses generated in the integrated
rent density raises from Eq. (24). Therefore, the energy and exergy ef­ system. For the integrated system, the operating temperature of the SMR
ficiency decrease as the current density raises, corresponding to the reactor is set as 700 ◦ C, and the S/C ratio is assumed to be 3. The other
single cell voltages. For the gross electrical efficiency, the energy effi­ operating parameters are shown in Tables 1 and 4.
ciency and exergy efficiency decrease continuously from 34.5% and Table 5 summarizes the exergy values of the system and each
33.3% to 10.0% and 9.5%, respectively. Accordingly, the overall energy component. It is evident from the table that the total exergy destruction
efficiency decreases from 55.6% to 27.1%, and the exergy efficiency is 950.4 W. The relative error between the total exergy destruction at the
decreases from 53.6% to 26.2%. This observation is consistent with system and component levels is only 0.01%, suggesting that the model
other researches that have been reported [3,36,45]. Within a specific system is accurate enough. In line with the general expectations, the
range, the power density of the system improves significantly with the PEMFC stack is the primary source of the system exergy destruction
increase of current density, while the system efficiency is gradually (326.27 W) with 34.33% of the relative exergy destruction ratio and
decreased. 74.47% of the exergy efficiency. This phenomenon is primarily due to
the irreversible heat loss of the stack, that is, the thermal loses from the
4.4. Detail exergy analysis for the intergraded system surface to the surroundings [1].
Further, the second-highest exergy destruction is the combustor
Based on the above research and investigation, the most suitable (311.7 W), accounting for 32.79% of the total exergy destruction. That is
operation parameters (S/C of 3.0, SMR reaction temperature of 700 ◦ C, mainly because the combustor model is under an adiabatic state,
and the PEMFC hydrogen utilization of 0.8) of the whole simulation resulting in higher outlet temperature and more heat loss. And the low

Fig. 17. (a) Exergy destruction and (b) relative exergy destruction ratio of each component.

12
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

Table 6 that each bubble size roughly represents the exergy destruction of each
The thermodynamic properties of the integrated system. component, which is expressed by the scale value on the top axis. And
Node Molar rate (mol/min) Temperature (℃) Pressure (bar) figure (b) displays the proportion of relative exergy destruction ratio of
each piece of equipment in the total.
Fuel
F1 0.086 25.0 10.0 Here, the exergy destruction ratio θIr,i (Eq. (52)) is introduced to
F2 0.086 300.0 10.0 analyze the problem further. As shown in Table 5, the results of the
F3 0.343 300.0 10.0 exergy destruction ratio and the relative exergy destruction are similar.
F4 0.343 500.0 10.0 The PEMFC stack (19.23%), the combustor (18.37%), and the heat
F5 0.501 700.0 10.0
F6 0.501 450.0 10.0
collector (10.94%) are the three highest exergy destruction ratio com­
F7 0.501 450.0 10.0 ponents. Under the operating parameters, the energy and exergy effi­
F8 0.501 300.0 10.0 ciencies are 47.4% and 45.7%, respectively. The energy efficiency is
F9 0.501 300.0 10.0 higher than the exergy efficiency because the LHV of the energy value
F10 0.311 300.0 1.0
(802.3 kJ/mol) is lower than the exergy value (831.6 kJ/mol) of CH4.
F11 0.190 300.0 1.0
F12 0.311 50.0 1.0 Table 6 shows the thermodynamic properties of each node in the
Water system. Furthermore, to demonstrate the exergy flow of the integrated
W1 0.257 25.0 1.0 system more vividly, a Sankey diagram is represented in Fig. 18 below.
W2 0.257 25.0 10.0 As the legend shows, the light blue arrow indicates the exergy flow of the
W3 0.257 135.6 10.0
W4 0.257 180.0 10.0
stream, the red arrow means the exergy destruction of the component,
W5 0.257 180.0 10.0 the pink arrow represents the thermal or the work exergy of the
W6 0.257 300.0 10.0 component, and the black block represents the component. For example,
W7 2.000 25.0 1.0 the PEMFC block: the inlet exergy flow consists of A5 stream (43.96 W)
W8 2.000 99.7 1.0
and F12 stream (1243.0 W), which is about 1278.0 W in total inlet
Air
A1 0.02 25.0 1.0 exergy flow; the outlet exergy flow only consists of Ex stream (255.9 W).
A2 0.02 35.3 1.1 Correspondingly, the exergy destruction (326.27 W) and the thermal
A3 1.1862 25.0 1.0 (175.81 W)/work exergy (520 W) can be calculated from Eqs. (40)–(44),
A4 1.1862 36.4 1.1 resulting in a total outlet exergy flow of about 1277.98 W. Contrasting
A5 1.1862 50.0 1.1
with the full inlet exergy flow (1278.0 W) and the total outlet exergy
Combustion
C1 1.585 996.0 1.0 flow (1277.98 W), two flow results are basically identical, indicating a
C2 1.585 920.2 1.0 good balance of the exergy of the component and the accuracy of the
C3 1.585 903.7 1.0 simulation model.
C4 1.585 856.1 1.0
C5 1.585 502.1 1.0
C6 1.585 493.6 1.0 5. Conclusion
C7 1.585 200.0 1.0
This work proposes and models a LT-PEMFC system integrated with
SMR. The main significant conclusions are summarized as follows:
exergy efficiency (66.34%) of the combustor also cements that. Due to
the significant temperature difference between the inlet and outlet of the a) With S/C of 3.0, SMR reaction temperature of 700 ◦ C, and the PEMFC
HC, the lowest exergy efficiency (41.94%) and the relative exergy hydrogen utilization of 0.8, the electrical power achieves 520 W,
destruction ratio reaches 19.53% (185.65 W). delivering energy and exergy efficiencies as high as 47.4% and
Fig. 17 presents the exergy destruction and the relative exergy 45.7%.
destruction ratio of each component to get a better and more precise
understanding of the exergy destruction difference. Figure (a) shows

Fig. 18. Exergy flow diagram (Sanky diagram) of the integrated system.

13
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

b) With increased H2 utilization μH2 from 0.5 to 1.0, the H2 supply rate References
decreases from 0.498 to 0.249 mol/min. The overall efficiency is
increased simultaneously, e.g., the energy efficiency raises from [1] Wang Z, Mao J, He Z, Liang F. Fuzzy control based on IQPSO in proton-exchange
membrane fuel-cell temperature system. J Energy Eng 2020;146(5):
42.0% to 49.2%, and the exergy efficiency boosts from 40.4% to 040200441–514.
47.5%. [2] Bird TJ, Jain N. Dynamic modeling and validation of a micro-combined heat and
c) In the temperature range from 640 to 670 ◦ C, the overall energy power system with integrated thermal energy storage. Appl Energy 2020;271:
114955.
efficiency reduces from 48.1% to 44.5%, and the exergy efficiency [3] Liu G, Qin Y, Wang J, Liu C, Yin Y, Zhao J, et al. Thermodynamic modeling and
falls from 46.3% to 42.9%; however, the energy efficiency improves analysis of a novel PEMFC-ORC combined power system. Energy Convers Manag
from 44.5% to 48.5%, and exergy efficiency raises from 42.9% to 2020;217:112998.
[4] Xu H, Song H, Xu C, Wu X, Yousefi N. Exergy analysis and optimization of a HT-
46.8% in the temperature range from 670 to 750 ◦ C, respectively. PEMFC using developed Manta Ray Foraging Optimization Algorithm. Int J
d) When current density increases from 0.1 to 1.6 A/cm2, overall en­ Hydrogen Energy 2020;45(55):30932–41.
ergy efficiency decreases from 55.6% to 27.1%, and exergy efficiency [5] Hajmohammadi MR, Aghajannezhad P, Abolhassani SS, Parsaee M. An integrated
system of zinc oxide solar panels, fuel cells, and hydrogen storage for heating and
decreases from 53.6% to 26.2%.
cooling applications. Int J Hydrogen Energy 2017;42(31):19683–94.
e) The thermodynamic analysis shows that the PEMFC stack is the [6] Bandlamudi V, Bujlo P, Sita C, Pasupathi S. Study on electrode carbon corrosion of
primary source of the system exergy destruction (326.27 W) with high temperature proton exchange membrane fuel cell. Mater Today: Proc 2018;5
34.33% of the relative exergy destruction ratio and 74.47% exergy (4):10602–10.
[7] Qin Y, Du Q, Fan M, Chang Y, Yin Y. Study on the operating pressure effect on the
efficiency. performance of a proton exchange membrane fuel cell power system. Energy
Convers Manag 2017;142:357–65.
CRediT authorship contribution statement [8] Rosli RE, Sulong AB, Daud WRW, Zulkifley MA, Husaini T, Rosli MI, et al. A review
of high-temperature proton exchange membrane fuel cell (HT-PEMFC) system. Int
J Hydrogen Energy 2017;42(14):9293–314.
Zaixing Wang: Writing – original draft, Conceptualization. Junkui [9] Jannelli E, Minutillo M, Perna A. Analyzing microcogeneration systems based on
Mao: Supervision, Funding acquisition, Resources. Zhenzong He: Data LT-PEMFC and HT-PEMFC by energy balances. Appl Energy 2013;108(1):82–91.
[10] Xing S, Zhao C, Ban S, Liu Y, Wang H. Thermodynamic performance analysis of the
curation. Fengli Liang: Validation. influence of multi-factor coupling on the methanol steam reforming reaction. Int J
Hydrogen Energy 2020;45(11):7015–24.
Declaration of Competing Interest [11] Ribeirinha P, Alves I, Vázquez FV, Schuller G, Boaventura M, Mendes A. Heat
integration of methanol steam reformer with a high-temperature polymeric
electrolyte membrane fuel cell. Energy 2017;120(3):468–77.
The authors declare that they have no known competing financial [12] Shafiee A, Arab M, Lai Z, Liu Z, Abbas A. Modelling and sequential simulation of
interests or personal relationships that could have appeared to influence multi-tubular metallic membrane and techno-economics of a hydrogen production
process employing thin-layer membrane reactor. Int J Hydrogen Energy 2016;41
the work reported in this paper.
(42):19081–97.
[13] Katikaneni SP, Al-Muhaish F, Harale A, Pham TV. On-site hydrogen production
Acknowledgments from transportation fuels: an overview and techno-economic assessment. Int J
Hydrogen Energy 2014;39(9):4331–50.
[14] Do H-Y, Kim C-H, Han J-Y, Kim H-S, Ryi S-K. Low-temperature proton-exchange
The support of this work by the Defense Industrial Technology membrane fuel cell-grade hydrogen production by membrane reformer equipped
Development Program (No: JCKY2018605B006), National Natural Sci­ with Pd-composite membrane and methanation catalyst on permeation stream.
ence Foundation of China (No. 51806103), and Aeronautical Science J Membrane Sci 2021;634:119373. https://doi.org/10.1016/j.
memsci.2021.119373.
Foundation of China (No. 201928052002) are gratefully acknowledged. [15] Minutillo M, Perna A, Sorce A. Green hydrogen production plants via biogas steam
In addition, an extraordinary acknowledgment is made to the editors and autothermal reforming processes: energy and exergy analyses. Appl Energy
and referees who make critical comments to improve this paper. 2020;277:115452.
[16] Ben-Mansour R, Abuelyamen A, Habib MA. CFD modeling of hydrogen separation
through Pd-based membrane. Int J Hydrogen Energy 2020;45(43):23006–19.
Appendix [17] Serra M, Ocampo-Martinez C, Li M, Llorca J. Model predictive control for ethanol
steam reformers with membrane separation. Int J Hydrogen Energy 2017;42(4):
1949–61.
The modeling equations for the major components used in Aspen Plus [18] Shagdar E, Lougou BG, Shuai Y, Ganbold E, Chinonso OP, Tan H. Process analysis
are list in Table A1. of solar steam reforming of methane for producing low-carbon hydrogen. RSC Adv
2020;10(21):12582–97.
[19] Fausto G, Luca P, Angelo B. A simulation study of the steam reforming of methane
in a dense tubular membrane reactor. Int J Hydrogen Energy 2004;29(6):611–7.
[20] Maqbool F, Abbas SZ, Ramirez-Solis S, Dupont V, Mahmud T. Modelling of one-
dimensional heterogeneous catalytic steam methane reforming over various
catalysts in an adiabatic packed bed reactor. Int J Hydrogen Energy 2021;46(7):
5112–30.
Table A1 [21] Oliveira ELG, Grande CA, Rodrigues AE. Steam methane reforming in a Ni/Al2O3
The modeling equations for the major components. catalyst: Kinetics and diffusional limitations in extrudates. Can J Chem Eng 2009;
87(6):945–56.
Components Modeling equations Description [22] Xu J, Froment GF. Methane steam reforming, methanation and water-gas shift: I.
Intrinsic kineticsAIChE J 1989;35(1):88–96.
Heat exchangers Q = U⋅A⋅LMTD U: the heat transfer
[23] Arzamendi G, Diéguez PM, Montes M, Odriozola JA, Sousa-Aguiar EF, Gandía LM.
coefficient, W/m2 K
Methane steam reforming in a microchannel reactor for GTL intensification: a
A: the heat exchange area, m2 computational fluid dynamics simulation study. Chem Eng J 2009;154(1-3):
LTMD: the log-mean 168–73.
temperature difference [24] De Falco M, Piemonte V, Di Paola L, Basile A. Methane membrane steam reforming:
Heat duty assessment. Int J Hydrogen Energy 2014;39(9):4761–70.
Compressor [( )k − 1 ] k: the ratio of the specific
pin k − 1 heats [25] Ye G, Xie D, Qiao W, Grace JR, Lim CJ. Modeling of fluidized bed membrane
Wt = m⋅ηt ⋅cp⋅Tin
pout ηt : polytropic efficiency reactors for hydrogen production from steam methane reforming with Aspen Plus.
Int J Hydrogen Energy 2009;34(11):4755–62.
Pump mW1 ⋅hW1 + Wp = mW2 ⋅hW2 Wp : power consumption
[26] Sayar A, Eskin N. Experimental and theoretical analysis of a natural gas fuel
Pd-M mF9 ⋅hF9 = mF11 ⋅hF11 + mF10 ⋅hF10 – processor. Int J Hydrogen Energy 2021;46(2):1569–82.
Combustor Gibbs model Refer to Section 2.3 [27] Hussain MM, Baschuk JJ, Li X, Dincer I. Thermodynamic analysis of a PEM fuel cell
power system. Int J Therm Sci 2005;44(9):903–11.
Mixer mF2 ⋅hF2 + mF3 ⋅hF3 = mF6 ⋅hF6 –
[28] Cai Y, Wang W-W, Wang L, Liu D, Zhao F-Y. A proton exchange membrane fuel
cell-compound thermoelectric system: bidirectional modeling and energy
conversion potentials. Energy Convers Manag 2020;207:112517.

14
Z. Wang et al. Energy Conversion and Management 246 (2021) 114685

[29] Pukrushpan JT, Peng H, Stefanopoulou AG. Control-oriented modeling and [38] WANG S, WANG S. Exergy analysis and optimization of methanol generating
analysis for automotive fuel cell systems. J Dyn Syst Measure Control 2004;126(1): hydrogen system for PEMFC. Int J Hydrogen Energy 2006;31(12):1747–55.
14–25. [39] Tippawan P, Im-orb K, Arpornwichanop A. Efficient heat allocation in the two-step
[30] Samsun RC, Pasel J, Janßen H, Lehnert W, Peters R, Stolten D. Design and test of a ethanol steam reforming and solid oxide fuel cell integrated process. Energy 2017;
5 kW high-temperature polymer electrolyte fuel cell system operated with diesel 133(6):545–56.
and kerosene. Appl Energy 2014;114:238–49. [40] Arsalis A. A comprehensive review of fuel cell-based micro-combined-heat-and-
[31] Jin Y, Rui Z, Tian Y, Lin Y, Li Y. Sequential simulation of dense oxygen permeation power systems. Renewable Sustainable Energy Rev 2019;105:391–414.
membrane reactor for hydrogen production from oxidative steam reforming of [41] Nalbant Y, Colpan CO, Devrim Y. Energy and exergy performance assessments of a
ethanol with ASPEN PLUS. Int J Hydrogen Energy 2010;35(13):6691–8. high temperature-proton exchange membrane fuel cell based integrated
[32] Zhang Y, Zhang S, Gossage JL, Lou HH, Benson TJ. Thermodynamic analyses of tri- cogeneration system. Int J Hydrogen Energy 2020;45(5):3584–94.
reforming reactions to produce syngas. Energy Fuels 2014;28(4):2717–26. [42] Laurencelle F, Chahine R, Hamelin J, Agbossou K, Fournier M, Bose TK, et al.
[33] Yilanci A, Dincer I, Ozturk H. Performance analysis of a PEM fuel cell unit in a Characterization of a Ballard MK5-E proton exchange membrane fuel cell stack.
solar–hydrogen system. Int J Hydrogen Energy 2008;33(24):7538–52. Fuel Cells 2001;1(1):66–71.
[34] Chahartaghi M, Kharkeshi BA. Performance analysis of a combined cooling, [43] Purnima P, Jayanti S. A high-efficiency, auto-thermal system for on-board
heating and power system with PEM fuel cell as a prime mover. Appl Therm Eng hydrogen production for low temperature PEM fuel cells using dual reforming of
2018;128:805–17. ethanol. Int J Hydrogen Energy 2016;41(31):13800–10.
[35] Liu G, Qin Y, Yin Y, Bian X, Kuang C. Thermodynamic modeling and exergy [44] Khila Z, Baccar I, Jemel I, Houas A, Hajjaji N. Energetic, exergetic and
analysis of proton exchange membrane fuel cell power system. Int J Hydrogen environmental life cycle assessment analyses as tools for optimization of hydrogen
Energy 2020;45(54):29799–811. production by autothermal reforming of bioethanol. Int J Hydrogen Energy 2016;
[36] Baniasadi E, Toghyani S, Afshari E. Exergetic and exergoeconomic evaluation of a 41(39):17723–39.
trigeneration system based on natural gas-PEM fuel cell. Int J Hydrogen Energy [45] Ye L, Jiao K, Du Q, Yin Y. Exergy analysis of high-temperature proton exchange
2017;42(8):5327–39. membrane fuel cell systems. Int J Hydrogen Energy 2015;12(9):917–29.
[37] Al-Sulaiman FA, Dincer I, Hamdullahpur F. Exergy analysis of an integrated solid
oxide fuel cell and organic Rankine cycle for cooling, heating and power
production. J Power Sources 2010;195(8):2346–54.

15

You might also like