You are on page 1of 13

Energy and Buildings 107 (2015) 213225

Contents lists available at ScienceDirect

Energy and Buildings


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

A simplied PEM fuel cell model for building cogeneration


applications
Sang-Woo Ham, Su-Young Jo, Hye-Won Dong, Jae-Weon Jeong
Division of Architectural Engineering, College of Engineering, Hanyang University, 222 Wangsimni-Ro, Seungdong-Gu, Seoul 133-791, Republic of Korea

a r t i c l e

i n f o

Article history:
Received 19 May 2015
Received in revised form 20 July 2015
Accepted 10 August 2015
Available online 14 August 2015
Keywords:
PEM fuel cell
CHP
Building cogeneration

a b s t r a c t
A simplied model of a polymer electrolyte membrane (PEM) fuel cell has been suggested for simulating
packaged commercial fuel cell systems. PEM fuel cell systems are used in building cogeneration applications because of their high efciency, low transmission loss and pollution, exible scalability, and low
noise. In conventional cogeneration applications, optimization techniques are utilized to size a system
and determine proper operational strategies using simulations. To evaluate the performance of a building cogeneration system, a fuel cell model should be concise but accurate to allow its implementation in
a whole-building simulation program. Some existing models are appropriate for building applications,
but they have some limitations in modeling when a commercial packaged fuel cell system is used. To
overcome these problems, a simplied fuel cell model is suggested for commercial packaged fuel cells by
adding some new variables and validating through experimental and published data. This model is relatively simple compared to other models but can be easily utilized in some limited cases with performance
predictions.
2015 Elsevier B.V. All rights reserved.

1. Introduction
Combined heat and power (CHP), which simultaneously generates heat and power, has been extensively used recently owing to
its high efciency [1]. In addition, when CHP systems are used at
a micro-scale or in a distributed manner (i.e., decentralized energy
generation), their efciency in terms of primary energy is maximized owing to their low transmission and heat losses [24].
Recently, among many applicants in micro CHP systems, fuel cell
systems have proved to be attractive in many aspects, including
the facts that they have high electrical efciency because of direct
conversion from fuel to electricity, low pollutant emissions, exible
scalability, and production of low noise during generation [4].
Numerous recent studies [514] have focused on the building cogeneration applications of polymer electrolyte membrane
(PEM) fuel cells (also known as proton exchange membrane fuel
cells) because they have high power density (small stack size), high
cogeneration efciency (sum of heat and power), and fast startup
time owing to their low operating temperature [4,15,16].
In general building cogeneration applications, PEM fuel cell systems generate power for buildings, and the recovered heat is used
for domestic hot water or heating. However, in the summer, the

Corresponding author.
http://dx.doi.org/10.1016/j.enbuild.2015.08.023
0378-7788/ 2015 Elsevier B.V. All rights reserved.

surplus heat that is recovered is useless. Therefore, researchers


have studied ways to utilize the heat for cooling. This technique
is usually termed trigeneration or combined cooling, heating, and
power (CCHP) [17]. When fuel cells are used for cooling applications, the general approach is to use the recovered heat for
generators of absorption chillers [18,19]. However, the operating
stack temperature of a PEM fuel cell is 6080 C, which is not sufcient for the generation process in absorption chillers. Thus, some
researchers [20] have focused on the applicability of PEM fuel cells
for the regeneration process in liquid desiccant dehumidication
systems.
Another area of research applies PEM fuel cells to building
cogeneration centers based on the design and operating strategies
of the fuel cell system. The amount of electricity generated by the
fuel cell system and the heat recovered by it are not always equal
to the power and heat demand of the buildings it serves. For this
reason, a system will generally be connected to the grid or will have
batteries with thermal storage to handle load variations [21]. Even if
the system is connected to the grid and includes auxiliary heaters,
sizing and operating strategies of the system must be optimized.
Many studies have been conducted on these issues, and they have
typically employed optimization techniques based on operational
strategies of the system and target variables, such as the primary
energy or the life cycle cost [10,2225].
In general, the primary energy savings or life cycle cost is estimated based on the annual operation, and in this case, the general

214

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225

Nomenclature
a0 , a1 , a2 model coefcients
b0 , b1 , b2 model coefcients
BX
xed (bias) uncertainty
specic heat of water (kJ/kg K)
cp,w
LHVfuel lower heating value of fuel (kJ/kmol)
n
number of measured data
molar ow rate of fuel (kmol/s)
N fuel
p
number of predictors in the model
PAC,anc
parasitic AC power (kW)
PAC,net
net generated AC power (kW)
PAC,net,nom rated maximum power from manufacturer (kW)
PAC,net,rated rated power at part-load condition from manufacturer (kW)
PAC,PCU generated AC power after PCU (kW)
PAC,ref
generated AC power at reference condition (kW)
net generated DC power (kW)
PDC,net
P*
dimensionless power ()
P gen,start average power generation in startup mode (kW)
P use,start average power use in startup mode (kW)
PCU
power conditioning unit
PLR
part-load ratio ()
recovered heat (kW)
Q
Q fuel
caloric value of fuel (kW)
Q fuel,ref caloric value of fuel at reference condition (kW)
recovered heat at reference condition (kW)
Q ref
Q
dimensionless recovered heat ()

Q fuel
dimensionless caloric value of fuel ()
r0 , r1
model coefcients
random (precision) uncertainty
SX
Tw,in , Tw,out inlet/outlet temperature of stack cooling water
( C)
UX
overall uncertainty
V fuel
fuel volume ow rate (m3 /s)

V fuel,start average fuel volume ow rate in startup mode


Vm
V w
ymeasured
y measured
y predicted

(m3 /s)
molar volume of fuel (m3 /kmol)
water volume ow rate (m3 /s)
measured value
average value of measure data
predicted value by the model

Greek symbols
,
model coefcients
0 , 1 , 2 model coefcients
DC,net
electrical (DC) efciency of fuel cell
PCU
PCU efciency
density of water (kg/m3 )
w

method used estimates the expected consumption using computer


simulations based on the general power and heating load proles
[10,22,23,26,27]. In this case, one can notice that the modeling
of each component (e.g., fuel cell, thermal storage) is essential to
determine the accuracy of the analysis.
Many studies have dealt with the modeling of fuel cell systems. Components of the fuel cell are simulated individually using
dynamic simulation, because the goal of such models is to design
stacks and fuel cells systems [2830]. However, for building applications, the design of fuel cells is not typically an issue because
commercialized fuel cell systems are typically installed with preprogrammed control algorithms. In addition, the fuel cell system
is not the only building component that should be modeled. Thus,

in practice, the fuel cell model must be accurate but simple and
implementable in a whole-building simulation.
A few researchers have developed simplied PEM fuel cell models for building applications. However, these models have some
limitations in practice because they are based on simple linear
assumptions [22,31] or a stack polarization curve without balance of plant [10,16,3237]. In response, a simple and empirical
fuel cell model for building cogeneration simulation was developed in the Annex 42 project of the International Energy Agency
[38]. The model was basically developed for a solid-oxide fuel
cell system [39] but was extended to a PEM fuel cell system
[40]. The Annex 42 model is concise but accurate; however, it
has been noticed that a more simplied model is necessary for
practical implementations. Specically, in real applications, residential fuel cells are available off the shelf, and only limited
parameters can be modied in this case without disassembling the
package.
In this present research, a simplied PEM fuel cell model is
suggested based on the Annex 42 model. The model aims at a commercial PEM fuel cell system in which the Annex model is difcult to
implement. With seven measurable variables (e.g., design capacity,
AC power), the model predicts power, heat, and fuel consumption of
a fuel cell system by mapping performance curves through regressions. Although this model is not as specic as that in Annex 42, it
can be applicable to commercial fuel cells because it is manufactured with preprogrammed capacity controls (i.e., less performance
variations and part-load mode operation). First, in this research, a
simplied PEM fuel cell model based on other fuel cell models [40]
is derived, and then the proposed model is validated by experimental measurements. The model is also validated by measurement
data acquired from open literature.

2. Simplied PEM fuel cell model


A simplied fuel cell model based on a previous fuel cell model
is introduced [40]. The suggested model is based on steady-state
operation and empirical data. Although there is a wide range of
dynamic PEM fuel cell models, we concluded that the dynamic
models are not suitable for analyzing the annual performance and
economic feasibility of fuel cells in building applications because
they are intended for stack design [28,29] or require a signicant
amount of performance information and calibration for each component and their control algorithms in a fuel cell system [30,4144].
That level of modeling for a fuel cell system is excessively time
consuming considering that it is a small part of a building, and
a viable option is to use a simple empirical model because the
system exhibits relatively steady performance. Moreover, the system often cannot be investigated at the component level for some
commercial fuel cell packages. In contrast, the simplied fuel cell
model based on empirical data is suitable for building applications
because it can be easily coded into existing building simulation
programs, such as EnergyPlus [45] and TRNSYS [46]. Dynamic performance characteristics of the plant within a simulation time step
(1560 min) are negligible in the context of annual energy simulations. Some studies have introduced simplied PEM fuel cell
models for building cogeneration applications, but we found that
these models are unsuitable for a building cogeneration application
because they are either too simple [22,31] or limited to a polarization curve [16,3236]. Whereas polarization curve modeling is
accurate, it requires time domain-based simulation platforms and
stack-level modeling [30,41], which requires too much effort for
one component (fuel cell system) in a whole building. Moreover,
it is difcult to model a stacks transient operation without disassembling the factory-assembled commercial fuel cell, which can
yield malfunctions [27]. Thus, in building applications, the general

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225

approach is to use performance-based modeling because commercial fuel cell systems are manufactured and preprogrammed in a
factory [27,39,47].

the fuel (=) is almost constant:


N fuel =

2.2. Model derivation


(a) AC power. In the Annex 42 PEM fuel cell model [38,40], the
authors controlled two variables: AC power and the temperature
of the stack cooling water. As shown in Eq. (1), the generated net
AC power (PAC,net ) can be estimated by subtracting the parasitic AC
power (PAC,anc ) consumed by ancillary devices from the total AC
power (PAC,PCU ), which is the AC power inverted by the power conditioning unit (PCU). The inverted total AC power can be expressed
simply as the product of the net generated DC power (PDC,net ) and
PCU efciency (PCU ), as shown by the following equation:
PAC,net = PAC,PCU PAC,anc = PDC,net PCU PAC,anc PDC,net

PDC,net
DC,net LHVfuel
=

2.1. Motivations
Our proposed model is based on the quasi-dynamic PEM fuel cell
model developed as part of Annex 42 by the International Energy
Agency [38,40]. This model is a simple but accurate empirical model
whose purpose is to serve as a general model to be integrated into
and interoperable among various building energy simulation programs, thereby enabling overall performance evaluation of building
cogeneration systems. Although this model is relatively concise
and can be integrated into whole-building simulation programs,
we found that some commercial fuel cell systems have limitations
with regard to this model.
In a commercial fuel cell system, all components of the system
are assembled into one package including balance of plant (e.g.,
reformer), and it is impossible to disassemble the whole package
and model each component owing to safety and security concerns.
Moreover, the system has preprogrammed capacity control algorithms because the use of dynamic capacity control on the fuel cell
in order to meet load changes is unnecessary in stationary applications. This is because the overriding design criteria in stationary
fuel cells are durability and stable operation [4850]. Thus, regardless of the fuel cells operational strategy (e.g., electricity-led or
heat-led operation), auxiliary heaters and thermal storage, as well
as electricity storage such as capacitors and batteries, are utilized
for handling variations in electricity and heat demand [21,51,52].
This is very common in any CHP design [15,32,53]. For this reason, a
typical capacity control algorithm for a commercial fuel cell aims to
regulate the current density [54] to meet the required constant AC
power by using a DC-DC converter and a DC-AC inverter [51,55,56].
Thus, many commercial fuel cells have preprogrammed capacity control algorithms (i.e., preset turn-down or part-load ratios).
Finally, the power and heat generation of commercial fuel cell packages sometimes differs from the design capacity (in laboratory)
after installation. Therefore, the simplied model aims at commercial fuel systems in which application of the Annex model is limited
and new parameters of part-load ratio and AC power output are
used.

(1)

In the Annex 42 model, the measured PDC,net and fuel consumption N fuel are used to estimate electrical efciency (DC,net , see Eq.
(2)). In addition, the model suggested that DC,net can be correlated
with a second-order polynomial in terms of PDC,net . However, when
we used a previously reported dataset [40], we found that 1 and 2
have relatively small values compared to 0 , and the relationship
between PDC,net and N fuel was almost linear. According to this correlation, we can conclude that the fuel consumption is a rst-order
function with respect to PDC,net because the lower heating value of

215

PDC,net
PDC,net
1

LHVfuel 0 + 1 PDC,net + 2 P 2
0
DC,net

(2)

In addition, the Annex 42 model suggested that PAC,anc is linearly


related to N fuel ; based on this relationship and the form of Eq. (2),
we can conclude that PAC,anc and N fuel are rst-order functions of
PDC,net . Thus, from Eq. (1), we can conclude that PAC,net is also a rstorder function with respect to PDC,net , assuming that PCU is almost
constant (1.5% variations were previously reported [40]).
(b) Part-load ratio and dimensionless power model. Commercial fuel cells are controlled by means of a part-load ratio
preprogrammed based on testing conducted by the manufacturer.
According to the literature [54,56], a PEM fuel cell controller modulates the current density to meet the desired AC load, even if the
polarization characteristics and stack performance vary in response
to changes in the operational characteristics (e.g., stack temperature or operating time) [5658]. Therefore, the preprogrammed
part-load ratio (Eq. (3)) was chosen as a parameter of the simplied
model because the AC output of a fuel cell is almost constant regardless of the operating point of current density and the resulting fuel
consumption:
PLR =

PAC,net,rated
PAC,net,nom

(3)

In practice, after installation, it is possible that the AC output


would differ from the rated AC output reported by the manufacturer. However, according to the literature [59,60] and our
experiments, the AC output remains nearly constant with time,
despite such a discrepancy. Thus, it is more accurate to use dimensionless AC power that is based on the reference AC power after
installation.
In this research, the dimensionless power used (Eq. (4)) is based
on the measured power considering the performance degradation
after installation. The reference condition will be addressed in Section 2.2(c). To reect the use of a preprogrammed controller, PAC,net
is written as a function of part-load ratio (PLR), as shown in Eq.
(5). This equation uses a second-order polynomial because the AC
output resulting from the preprogrammed part-load ratio is not
always linear owing to differences in the PCU efciency and parasitic power (Eq. (1)) under different part-load ratio conditions and
performance degradation:
P =

PAC,net
PAC,ref

(4)

P = a0 + a1 PLR + a2 PLR2

(5)

(c) Dimensionless heat recovery model. Unlike the case of a


portable fuel cell application, in a building application, both power
and heat are important. Thus, predicting the recovered heat is
somewhat important in the simplied model. In the same way as
that adopted for power, we used dimensionless heat in this model
(Eq. (7)), and the recovered heat (Q ) can be estimated based on
measured data using Eq. (6). The dimensionless heat (Q ) is the
ratio of the recovered heat in the present condition (Q ) to that in
reference condition (Q ref ):

Q = cp,w w V w Tw,out Tw,in


Q
Q =
Qref

(6)
(7)

The original Annex 42 model [38] predicted the recovered heat


as a function of stack temperature, but ones ability to measure

216

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225

Ideal thermal cell voltage (V)

Cell voltage (V)

1.2

Heat
0.8

Operating point
0.4

Power

0
0

0.4

0.8

1.2

1.6

Current density(A/cm2)
Fig. 1. General polarization curve of a single cell.

the stack temperature can be limited in practice [56]. Moreover,


the model must include a detailed representation of each subcomponent in the cooling systemthat is, the heat exchanger, the
air-cooler, and the pump. Owing to these limitations, Johnson et al.
[40] proposed a simplied heat recovery model including DC power
and stack cooling water temperature, the latter of which represents
the stack temperature. Although this model is accurate but concise,
it requires measurements for various points, which could be limited
for some other PEM fuel cells where collecting diverse data is inaccessible [60]. Specically, replacing the DC power term with the
part-load ratio degrades the accuracy of the model.
Therefore, we modied the heat recovery model as shown in Eq.
(8). Heat generation depends on the operating current density, and
the operating point is modulated to meet the required AC demand
[54,56], causing heat generation to uctuate (Fig. 1). However, the
ideal thermal voltage (i.e., the theoretical chemical energy) does
not change much, regardless of the operating point. Thus, it is more
reliable to predict the power and heat performance simultaneously
when predicting the generated heat (Fig. 1). We also found, through
a detailed inspection of previously reported data [40,59,60], that
Q varies slightly with Tw,in , so we added a linear term () in Eq.
(8). This dependency might be attributed to the slight performance
changes in the polarization characteristic when the stack cooling
water temperature (Tw,in ) is changed [56]:
P Q =

PAC,net + Q

= r0 + r1 PLR(Tw,in + Tw,in )
PAC,ref + Q ref

(8)

According to the literature [40] and our experiment, we noticed


that the amount of recovered heat varies with the stack cooling

water temperature. However, within the range of 1040 C, only


small differences were observed. Additionally, in the cold-start
mode, electricity generation began when the temperature of stack
cooling water reached values in the range of 4050 C. For this reason, the reference condition was set to a part-load ratio of 100% and
a stack cooling water temperature of 40 C.
Previous experiments illustrated the impact of outdoor air and
stack coolant ow rate on fuel cell performance [59,60]. The generated power is affected by neither the ambient air conditions
(temperature and humidity) nor the ow rate because the fuel cell
system is internally controlled to provide stable electricity. From
the data in [59], the coolant ow and ambient temperature can
indirectly affect heat generation in fuel cells when there are special
conditions such as cold ambient air temperature or low coolant
ow. When the ambient air is cold, the recovered heat could be
lost while owing through the pipe and into thermal storage. In
other words, the performance variation due to cold air is presumably attributed to poor insulation [59]. Moreover, when the coolant
ow is small and not sufcient to cool the stack, an auxiliary independent cooler operates, resulting in a reduction of recovered heat
[59].
However, we also noticed that these conditions rarely exist in
real operation because the internal temperature of the fuel cell
engine remained somewhat constant for a year by generating heat
in the reformer. Moreover, generally, commercial fuel cell coolant
is designed to provide sufcient cooling with sufcient ow. Therefore, the general model [38] shows that the recovered heat can be
expressed as stack (or coolant) temperature. Thus, in this research,
we decided not to include the ambient temperature and coolant
ow in the model.
2.3. Model structure and fuel consumption
Fig. 2 shows a schematic diagram of our fuel cell system, which
was purchased from a vendor. The balance of plant differs slightly
from vendor to vendor, but the general conguration is almost
identical. A commercial fuel cell comprises various subcomponents,
including the fuel processing module (reformer), air supply module, PCU, and heat recovery module (water tank) [21]. The role of
the reformer is to convert fuel (e.g., methane gas, natural gas, or
liqueed petroleum gas) to hydrogen gas, because procurement of
pure hydrogen as a fuel is rare in practice. For conditioning, conversion, and purication of the fuel, resources such as water, air, or part
of the fuel are used in the reformer. Heat and electricity are generated in the fuel cell stack through electrochemical reactions, and
the generated DC electricity is converted to AC power in the PCU.
To ensure sufcient reactions, hydrogen stoichiometry is generally
maintained above than 1.1, resulting in unconsumed hydrogen fuel

Fig. 2. Schematic diagram of a commercial fuel cell system.

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225

217

LHVfuel can be calculated, measured, or quoted as reported by the


fuels local distributor:

=
Q fuel

Q fuel
Qfuel,ref

(9)

= b0 + b1 PLR + b2 PLR2
Q fuel

(10)

V
LHVfuel
Q fuel = fuel
Vm

(11)

3. Model validation
Fig. 3. Structure of the simplied PEM fuel cell model.

in the stacks [52]. One approach to reduce this wasted unconsumed


fuel is to reuse it as a fuel for combustion in the reformer [52]. Thus,
the afterburner and anode tail gas oxidizer are not included in this
model. Finally, heat is recovered through the use of cooling water
and a water tank. For safety, the system includes an air cooler, which
is operated to cool the water tank when the recovered heat exceeds
the heat demand (i.e., temperature of the water tank).
As mentioned above, it is impractical to model all heat, mass,
and the electrical transport phenomena of the stack and other subcomponents when analyzing a fuel cell for building cogeneration
applications. In addition, when the fuel cell system installed, only
limited data can be measured without disassembling the system.
Therefore, we simplied the complex system by collecting easily
accessible and necessary data for building applications. The black
box model comprises 10 parameters: seven inputs (part-load ratio,
net generated AC power, fuel consumption, fuel heating value, inlet
and outlet temperatures, and volume ow rate of stack cooling
water), and three outputs (generated power, generated heat, and
fuel consumption) (Fig. 3).
In this model, the input parameters are PLR and Tw,in . PLR is a
value preprogrammed by the manufacturer, such as full-load and
half-load (Eq. (3)). By changing PLR, a user can modulate the fuel cell
power as needed. The variable Tw,in is the temperature of water that
is used to cool the stack and other components. In fact, it is more
convenient to measure the water temperature of the demand side
than that of the cooling water (Fig. 2). However, the storage tank
interposed between them prevents the temperature of the water
on the demand side from immediately affecting the performance
of the fuel cell. Thus, in the simplied model, the cooling water
temperature indicates the temperature of the stack cooling water
(Cooling water in in Fig. 2), which directly affects the stack temperature. With this modeling strategy, one can simply incorporate
the fuel cell model with the existing thermal storage model [37] for
whole-building simulation.
The output parameters of this model are generated power,
recovered heat, and fuel consumption. In practice, only these three
variables are useful in analyzing the building cogeneration system,
so we disregarded all other possible output parameters.
Based on Section 2.2(a), we consider the relationship of PAC,net
and N fuel to be almost linear. Furthermore, in Eq. (5), the variable
P*, which is proportional to PAC,net , is expressed as a second-order
polynomial with respect to PLR. Thus, N fuel can be correlated with a
second-order polynomial with respect to PLR (Eq. (10)). However,
the lower heating value of fuel (LHVfuel ) varies with the type of
fuel and the operating temperature and pressure, so the use of the
total caloric value of consumed fuel (Q fuel ), rather than N fuel , is more
accurate in general applications. Like power and heat, we also use a
dimensionless fuel heating value, as shown in Eq. (9). This dimensionless fuel heat is expressed in Eq. (10). As shown in Eq. (11),
the fuel heating value can be estimated by multiplying LHVfuel by
the volume ow rate of the fuel (V fuel ) measured onsite, whereas

The main application of this model is to test the performance


of installed commercial fuel cells; in this research, the suggested
model was tested and validated experimentally. As mentioned earlier, commercial fuel cells are tested in the laboratory, and the actual
performance of each fuel cell after installation can differ from its
performance in the laboratory. Based on our experimental results,
we demonstrate that the proposed model can be applicable in real
settings.
3.1. Fuel cell system
The PEM fuel cell system was installed in a small chamber on the
top of a four-story factory building located in Incheon, South Korea.
The fuel cell was fueled with natural gas, and all its system components (reformer, PCU, stack, and hot water storage) were assembled
in one cabinet (Figs. 2 and 4b). The purpose of the system was to
generate power and heat for a liquid desiccant and evaporative
cooling-assisted 100% outdoor air system (LD-IDECOAS) [20,61,62].
The LD-IDECOAS system required heat to regenerate the desiccant
solution, and electricity was used for other components such as
fans and pumps. In addition, the system was connected to the grid,
so the stack generated constant AC power according to the preset
part-load ratio. When more electricity was generated than was consumed by the LD-IDECOAS, the surplus electricity was sent to the
grid. Hence, there was no electrical storage. However, the system
was designed based on electricity, so a thermal storage tank and
an air cooler were installed to control the temperature of the stack
cooling water (Fig. 2). Table 1 summarizes other details relevant to
the fuel cell system.
3.2. Experiments
In this study, only seven variables needed to be measured during the experiments: part-load ratio, net generated AC power, fuel
consumption, fuel heating value, inlet and outlet temperature, and
volume ow rate of stack cooling water. Table 2 lists these measured variables and their corresponding sensors.
The experiments were divided into three parts in this study. The
rst part is to obtain steady-state data to calibrate the model suggested in this paper. As with general fuel cell systems for stationary
applications, the part-load ratio is the only controllable variable for
our fuel cell system, and stack cooling water temperature can be
controlled using an external chiller and heater. Thus, the part-load
ratio and stack cooling water temperature were controlled in the
experiments (Table 4). The other part is to monitor the performance
variations in the fuel cell with changes in outdoor air conditions.
Outdoor conditions are impossible to control in practice, so the
performance of fuel cells was intermittently recorded from June to
December. Then, the collected data were statistically analyzed with
respect to the outdoor air conditions. As mentioned in Section 2.2,
the commercial fuel cell system internally controls stack incoming
water and gas conditions, and no signicant changes were observed
in its performance. The nal step is to monitor the transient characteristics of the fuel cell system when it is in startup, shutoff, and

218

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225

Fig. 4. Test facility.

Table 1
Fuel cell system summary.
Location
Type
Rated AC power
Rated heat
Balance of plant
Fuel
Capacity control

Table 4
Investigated test sets.

Incheon, South Korea


Polymer electrolyte membrane (PEM) fuel cell
10 kW (electrical efciency: 34.4% (LHV))
13.94 kW (thermal efciency: 49.5% (LHV))
Reformer, PCU, hot-water storage
Natural gas (LNG)
100% and 60% AC power (PLR = 1, 0.6) (grid-connected)

Temperature of inlet stack cooling water (Tw,in )


15 C

PLR
0.6
1

20 C



25 C



30 C



35 C



40 C



interval were used for the xed (bias) uncertainty (BX ) and the
random (precision) uncertainty (SX ), respectively:

Table 2
Measured variables and corresponding sensors used.

UX =

Measured variable

Sensor

Specications and
accuracy

Net generated AC power


(PAC,net )

1% of measure
value

Inlet stack cooling water


temperature (Tw,in )

Digital electricity
meter
(LD3410DR-040, LSIS)
T-type thermocouple
(MV2000, Yokogawa)

Outlet stack cooling water


temperature (Tw,out )

T-type thermocouple
(MV2000, Yokogawa)

Stack cooling water ow


(V w )

Ultrasonic ow meter
(TFM 100, Taehung
M&C)
Gas meter
(G2.5L, Daesung)

Gas ow rate (V Fuel )

LHV of fuel (LHVFuel )

Obtained from
supplier
(http://www.kogas.or.kr)

200 to 400 C
(0.15% + 0.7 C) of
measure value
200 to 400 C
(0.15% + 0.7 C) of
measure value
1% of measure
value
0.0254 m3 /h
1.5% of measure
value
Monthly average
LHV (kJ/m3 )
2% of measured
value

Table 3
Overall uncertainty of variables.
PAC,net
Q
Q Fuel
Tw,in
Tw,out
V w
Min 0.06 kW 0.19 kW 0.33 kW 0.72 C 0.73 C 0.0025 l/s
Max 0.15 kW 1.48 kW 2.00 kW 0.89 C 1.04 C 0.0028 l/s

V Fuel
0.0005 m3 /min
0.0025 m3 /min

part-load change modes. The impact of dynamic characteristics on


overall performance is not signicant, but some of discussions will
be presented in Section 3.6.
The experimental uncertainties (Table 3) were analyzed using
the root-sum-squares (RSS) method [63]. The overall uncertainties
(UX ) were estimated through Eq. (12), and sensor accuracy and
standard deviations of measurements with a 95% condential

BX2 + tSX

2



n
2
1  1 
SX = 
Xi X
n

n1

(12)

(13)

i=1

3.3. Validation process


The validation process was divided into two steps. First, the
performance of fuel cells was measured in the cases where the
controllable variables (i.e., PLR and Tw,in ) were set to their maximum, minimum, and median values, for estimation of the empirical
coefcients of the model (training variables), as indicated in Eqs. (5),
(8), and (10) (marked by in Table 4). After that, additional tests
were conducted using other randomly chosen conditions (marked
by  in Table 4) (i.e., testing variables), and the results were compared to the values predicted by the model during the rst step. To
determine the steady-state performance, each test was conducted
for one hour, and average values over periods of 30 min were estimated in the validation process [40]. In the rst step, six total tests
were conducted (including three different water temperatures and
two part-load ratios), followed by an additional 13 tests.
The fuel cell used in this research had two capacity control
options (two specic part-load ratio values; namely, 1 and 0.6). The
capacity was based on AC power, and the coincidental and nominal powers were 10 kW and 6 kW, respectively. In practice, we had
limited control over the water temperature, which ranged from 15
to 40 C. In theory, a PEM fuel cell can be operated with stack temperatures of 80100 C; however, for safety reasons, it is usually
operated in the range of 6080 C [56]. Nevertheless, in our fuel
cell, the temperature of the inlet stack cooling water was limited
to 45 C by the manufacturer to protect the water lter. Temperature in excess of 45 C caused the preprogrammed controller to
operate the air cooler (Fig. 2). According to the manufacturer, the
lter used for purifying the stack cooling water would be damaged

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225

219

Fig. 5. Performance map of tested fuel cell. (For interpretation of the references to colour in this gure citation, the reader is referred to the web version of this article.)

by water at temperatures higher than 45 C. This was attributed to


the use of open-loop stack cooling water. In other words, because
the water in the thermal storage tank was directly introduced into
the stack, the lter should provide high-quality purication performance to remove ions and contaminants from the cooling water
because these substances could cause short circuits, and ultimately
lead to stack damage [52]. For this reason, the controllable range of
the stack cooling water temperature was limited to 1540 C.
In this research, the accuracy of the model was evaluated
through two metrics: the coefcient of variation for the root mean
square error (CV(RMSE)) and the normalized mean bias error (MBE).
These two metrics are typically used to evaluate the accuracy of
a calibrated energy model and measured data in whole-building
simulations. They are also used for the component models in simulations [64]. CV(RMSE) (Eq. (14)) is used to evaluate the degree of
data scatter. Values less than 5% are considered to be acceptable
[64]. MBE (Eq. (15)) quanties the bias of the model. Correspondingly, positive values indicate underestimation, and negative values
indicate overestimation:

CV (RMSE) =

i=1 
n

ymeasured,i y predicted,i

y measured

np

2
100 [%]
(14)

i=1
MBE =

ymeasured,i y predicted,i
y measured (n p)

100 [%]

(15)

3.4. Model validation on experimental results


The test results and model coefcients are summarized in Fig. 5
and Table 5. The model coefcients were estimated by regression
using six boundary points within the controllable range, marked
with lled blue dots in the gure. The other 13 tests for validation
are marked with hollow blue dots.
All test results were compared to the values predicted by
the model. Although the values were predicted using only the

Table 5
Model coefcients of tested fuel cell.
Reference
conditions
Equations

Power:
8.72 kW

Power (Eq.
(5))
Heat + power
(Eq. (8))

P* = 0.2367 + 0.7575PLR
R2 = 0.9817
P Q = 0.1572 +
1.493PLR(0.0008274Tw,in
0.0809
+ Tw,in
)

Fuel (Eq.
(10))
Range of
model

R2 = 0.9992

Q fuel
=
0.1938 + 1.1913PLR
2
R = 0.9982
PLR: 0.6, 1

Heat:
7.59 kW

Fuel:
40.6 kW

Tw,in : 15.140.3 C

part-load ratio, the predicted AC power and fuel consumption values agreed well with the measured values (Figs. 6 and 7). The main
reason is that the algorithm for stack capacity control is based on AC
power, and the fuel consumption actually depends on the stacks
fuel demand. As a result, the model yielded CV(RMSE) values of
2.6% and 1.89% for power and fuel, respectively, and an MBE value
of nearly zero for both of these parameters.
The predicted and measured values of recovered heat were
compared (Fig. 8). The heat recovery model agreed well with the
measured data, and all tested data fell within the 10% error bound.
The error of the heat recovery model was noticeably larger than
that of the other variablesnamely, power and fuel. In fact, the
heat recovery model predicted the sum of power and heat, not the
heat itself, and thus the predicted heat included the error of the
power model. Furthermore, the uncertainties associated with the
measurements of heat were noticed to be relatively large. The main
reason for this is the poor sensor accuracy of the water temperature. Because the heat generation was estimated based on Eq. (6),
the errors of each variable were propagated with changes in temperature, leading to large uncertainties. In spite of the cycling and
large uncertainty, we observed that the 1 h average of produced

220

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225


Table 6
Model coefcients.

10
CV(RMSE) = 2.6%
MBE = 0%

PAC,net,measured [kW]

5% error bound

Reference
conditions
Equations

Power: 0.9745 kW

Power (Eq.
(5))

P* = 0.082 + 1.2194PLR
0.221PLR2
R2 = 0.9988
P Q = 0.0125 +
0.6907PLR(0.01519Tw,in +
0.1953
)
Tw,in

Heat + power
(Eq. (8))

Fuel (Eq.
(10))

0
0

10

Range of
model

Heat:
0.1538 kW

R2 = 0.9981

Q fuel
= 0.1456 +
0.6162PLR + 0.2348PLR2
2
R = 0.9993
PLR: 0.251

Fuel:
0.2941 kW

Tw,in : 16.159.1 C

PAC,net,predicted [kW]
1

Fig. 6. Power model validation with experimental data.

CV(RMSE) = 1.43%
MBE = 0%

50

40

Qfuel,measured [kW]

PAC,net,measured [kW]

0.8
CV(RMSE) = 1.89%
MBE = 0.32%

30
5% error bound

20

0.6

0.4
5% error bound

0.2

0
0

10

0.2

0.4

0.6

0.8

PAC,net,predicted [kW]
0
0

10

20

30

40

50

Fig. 9. Power model validation with experimental data from [40].

Qfuel,predicted [kW]
3

Fig. 7. Fuel model validation with experimental data.

CV(RMSE) = 1.08%
MBE = 0.01%

2.5

10

Qmeasured [kW]

Qfuel,measured [kW]

CV(RMSE) = 5.12%
MBE = -0.2%

1.5
1

5% error bound

0.5

10% error bound

0
0

0.5

1.5

2.5

Qfuel,predicted [kW]
0
0

10

Fig. 10. Fuel model validation with experimental data from [40].

Qpredicted [kW]
Fig. 8. Heat recovery model validation with experimental data.

heat was relatively constant. The CV(RMSE) value of this model was
5.12%, and this value was within the acceptable range [64].
3.5. Model validation on literature data
The simplied model was also validated based on data from the
literature [40]. The coefcients of the simplied model (Eqs. (5), (8)

and (10)) were estimated using multivariate regression [65]. Table 6


presents the regression results. The reference condition is the fuel
cells performance at PLR = 1 (100%) and Tw,in = 40 C, as mentioned
in Section 2.2(c). Correspondingly, the measured part-load ratios
were 0.25, 0.5, 0.75, and 1 (i.e., 25%, 50%, 75%, and 100%).
Figs. 9 and 10 compare the measured and predicted results for
power and fuel consumption, respectively. Although the models
predict the generated power and the consumed fuel based on a
xed part-load ratio, predictions from both models differed slightly
from the measured data. This was attributed to the facts that the

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225


Table 7
Summary of transient variables in the fuel cell model.

2
CV(RMSE) = 3.2%
MBE = 0.01%

CV(RMSE) = 3.2%
MBE = 0.01%

1.6

Qmeasured [kW]

221

1.2

Mode

Duration
time

Values

Description

Start-up

3h

P use,start : 2.4 kW

Electricity
consumption
Electricity generation
in the fuel cell

0.5 h

Start-up

0.8
5% error bound

P gen,start : 5.3 kW

Start-up

3h

V fuel,start : 0.00018055 m3 /s

Shut-off

2h

Gas consumption in
start-up mode
At least 2 h is required
to start after shut-off

0.4

0
0

0.4

0.8

1.2

1.6

Qpredicted [kW]
Fig. 11. Heat recovery model validation with experimental data from [40].

AC power is generated according to the preprogrammed partload ratio, and the fuel is consumed accordingly. As a result, the
CV(RMSE) values were 1.43% and 1.08% for power and fuel modeling, respectively, which were within the acceptable range of less
than 5% [64]. In addition, the MBEs were almost 0% for both models.
These results also indicate that it is acceptable to directly predict
AC power rather than the DC power.
Fig. 11 compares the measured and predicted values of the heat
recovery model. Unlike the power and fuel models, the recovered
heat was affected by the temperature of the stack cooling water,
and scattered data were observed. The heat recovery model was
quite accurate and yielded an acceptable CV(RMSE) value of 3.2%
(<5%) and an MBE of nearly 0%.

that, it generated approximately 2.65 kWh of electricity and consumed 0.85 m3 of fuel for 30 min. Not shown in Fig. 12, the system
also consumed electricity because it did not generate electricity
during this period. In both modes, the system consumed 2.4 kW on
average. These values are summarized in Table 7.
The operational characteristics of the system in part-load
change mode are shown in Fig. 13. As shown in Fig. 13a, 610 kW
change mode takes 20 min, and 106 kW change mode takes 5 min.
In addition, shutoff mode takes less than 10 min to stop consuming
fuel and generating electricity. However, the manufacturer prevents the fuel cell from being turned on again for 2 hours from
shutoff mode for safety reasons (cool-down time). Both part-load
change and shut-off modes are completed in 1020 min. The experimental results of startup, part-load change, and shutoff modes are
similar to the data in [60].
The default time steps for whole-building simulation programs
such as EnergyPlus [45] and TRNSYS [37] are 15 and 60 min, respectively. Thus, only generation/consumption of electricity and fuel in
startup mode must be considered in annual simulation. However,
the time required by shutoff mode must be considered even though
there is no fuel consumption.

3.6. Transient characteristics


4. Application and discussion

10

4.1. Implementation example


In this section, an example application of the proposed model
will be presented for guidance. After presenting the process of
model implementation in a whole-building simulation domain,
simple calculation results will be discussed.
Fig. 14 shows the overall process of model implementation. First,
after calibrating the fuel cell model, the whole-building simulation
program calculates the buildings energy use in the current time
step, the information related to electricity and heating demand, and

0.06

10

0.04
6

4
Warm up period

0.02

8:26

8:47

9:08

9:29

9:50

10:11

10:32

10:53

Electricity (generation) [kW]

0.07
Electricity
Fuel

Fuel (consumption) [Nm3/min]

Electricity (generation) [kW]

Electricity
Fuel

0.06

8
0.05
6

0.04
0.03

4
Warm-up period

0.02
2
0.01
0

5:46

6:07

6:28

6:49

7:10

Time

Time

(a) 6 kW start

(b) 10 kW start
Fig. 12. Transient characteristic start mode.

7:31

7:52

0
8:13

Fuel (consumption) [Nm3/min]

Although stationary fuel cell systems are generally designed for


steady operation, transient operational characteristics in startup,
part-load change, and shutoff modes must be considered to analyze
the annual performance of the whole building. Fig. 12 shows the
operational characteristics of our fuel cell systems during startup
mode. According to the manufacturer, the warm-up period to start
the system takes 3 h because the reformer requires more than 2 h to
sufciently increase its temperature to make pure hydrogen while
consuming fuel. The test results showed that 2.53 h were needed
for startup mode. For 2.5 h, the system consumed approximately
1.1 m3 of fuel (natural gas) without electricity generation. After

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225

0.06

0.06
Electricity
Fuel

0.045
0.04
7
0.035
Electricity
Fuel

13:03

13:09

13:15

13:21

0.03

Electricity (generation) [kW]

Electricity (generation) [kW]

0.05
8

Fuel (consumption) [Nm3/min]

0.055
Mode change (6 to 10 kW)

0.055
0.05

8
0.045

Mode change (10 to 6 kW)

0.04
7
0.035
0.03

0.025
13:27

15:00

15:06

Time

15:12

15:18

Fuel (consumption) [Nm3/min]

222

0.025
15:24

Time

(a) 6 kW to 10 kW

(b) 10 kW to 6 kW
Fig. 13. Transient characteristic part-load change mode.

Whole building simulation domain

Required electricity
Required heating energy
Hot water temperature
Hot water flow rate
Time step information

Building
surface
model

Building
system
model

Building
airflow
model

Net electricity
Net fuel consumption
Hot water temperature

Fuel cell model


Electricity
model

Heat
recovery
model

Auxiliary device
Fuel
model

In fuel cell:
Generated electricity
Generated heat
Fuel consumption

Fuel cell
operational schedule
or
control algorithm

Thermal storage
constraints for safe
operation in fuel cell
(e.g., min and max
temperature in
thermal storage)

Thermal or electric
storage model
Thermal
Storage
energy
balance

Fuel cell
Auxiliary
cooler

Auxiliary
boiler

Net electricity
Net heating energy
Temperature change in storage
Electricity changes in battery

Electricity
battery
energy
balance

Fig. 14. Model implementation schematic diagram.

hot water information (if heating is implemented by hot water circulation with a fuel cell system). The fuel cell model then estimates
electricity and heating energy generation in the current time step
by using Eqs. (4)(11) based on the fuel cells control algorithm.
In general, schedule-based fuel cell operation (grid-connected) is
used in stationary applications for reliable operation. In most cases,
there are mismatches in electricity and heating demand between
the building and the fuel cell system. Thus, models for the electricity battery (if it exists) and thermal storage are necessary. In
these models, the net required electricity from the grid and heating
energy from the auxiliary boiler are estimated. Moreover, the temperature changes in the thermal storage must be calculated because
the performance of the fuel cell is affected by this factor. Finally,
there might be minimum or maximum constraints for hot water
temperature (stack coolant) in the fuel cell system. If a fuel cell auxiliary cooler or heater is installed for these reasons, the energy used
in these devices is added to the net electricity and heating energy
demand, and some information such as hot water temperature is
transferred to the next time step.

Fig. 15 shows an implementation of the model with hourly time


steps. The example is simply estimated based on the load prole
of a residential house presented in [30]. It is assumed that one fuel
cell system modeled in this study handles three residential houses
in [30]. That kind of load prole can be estimated in a wholebuilding simulation domain as presented in Fig. 14. In Fig. 15a, the
positive y-axis shows the electricity demand, and the negative part
represents the electricity generation in the fuel cell system. In this
example, a simple schedule-based control algorithm is used for the
fuel cell system (gray polygon in Fig. 15a). The time 0:003:00 is
the startup period, 3:0015:00 is 6-kW mode, and the remainder
is 10-kW mode. The blue polygon represents electricity demand in
the building, and the orange one represents the demand in the fuel
cell system in startup mode. The yellow polygon represents the
net electricity demand in the building (i.e., positive: buy electricity
from the grid; negative: sell electricity to the grid).
Fig. 15b is a daily heating demand prole based on the control
algorithm in the electricity example. The blue, yellow, and gray
polygons are heating demand in the building, net required heating

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225

223

10

Electricity demand [kW]

8
6
4
2
0
-2

9 10 11 12 13 14 15 16 17 18 19 20 21 22 23

-4
-6
-8
Start-up mode

-10

6 kW mode

10 kW mode

Time [hr]
Electricity demand (building) [kW]

Electricity generation (fuel cell) [kW]

Electricity from grid [kW]

Electricity demand (fuel cell) [kW]

41

40

39

38
37

0
-2

9 10 11 12 13 14 15 16 17 18 19 20 21 22 23

36

-4

35

-6

34

-8

33

-10

Storage temperature [C]

Heating demand [kW]

(a) Daily electricity demand

32

Time [hr]
Heating demand [kW]

Fuel cell heating energy [kW]

Boiler heating energy [kW]

Storage temperature [C]

(b) Daily heating demand


Fig. 15. Implementation example. (For interpretation of the references to colour in this gure citation, the reader is referred to the web version of this article.)

energy in the boiler, and fuel cell heating energy, respectively. In


this example, fully mixed thermal storage is assumed, and the heating energy generation in the fuel cell system varies, although it is
very small because of the small temperature changes in the storage.
4.2. Discussions
In this study, a simplied fuel cell model is suggested for
packaged fuel cell systems, and the model is validated against
experimental results and published data. Although diverse fuel cell
models exist for building applications, this study specically aims at
the packed system. Table 8 summarizes the characteristics of the
diverse model. Some dynamic models based on electrochemistry
are useful for designing stack and fuel cell systems [30,41,44], but
they are complex and have some limitations with regard to modeling in practice. Although some building simulation programs such
as TRNSYS [37] and EnergyPlus (Annex 42 model) [40] provide fuel
cell models for analyzing building cogeneration, the TRNSYS model
is limited to the stack only. The Annex 42 model (EnergyPlus) is
powerful in terms of building applications and applicable in general. However, the models use might be limited in some packaged
fuel cell systems where component modeling is not accessible without disassembly. In other words, the suggested model specically
aims at this case, and the Annex model should be more appropriate
for the general case.

Table 8
Comparison of different fuel cell models for building cogeneration.
Model

Modeling method

Applications

Limitations

Dynamic
models
[30,41,44]

Physical model
based on
electrochemistry

Stack design and


dynamic
simulation

TRNSYS [37]

Annual simulation
and control

Annex 42 [40]

Physical model
based on
electrochemistry
Empirical model

Modeling
difculties and
interoperability
with building
simulation
program
Model is limited to
the stack

Present study

Empirical model

Annual simulation
and control for
general fuel cell
system
Annual simulation
and control for
commercial fuel
cell package

Steady-state
model, each
component should
be modeled
Steady-state model
and limited to the
commercial
package

In addition, there are a few limitations in the application of


this model to whole-building simulation. Although the model
coefcients in Eqs. (3)(11) in Tables 5 and 6 can be used for a
simulation as a default case, the coefcients must be calibrated
through experiments if the performance of the other fuel cell system must be analyzed. Furthermore, the performance of the model

224

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225

is valid only within experimental data because it is an empirical


model that does not guarantee extrapolation. Finally, the use of a
complex control algorithm for a fuel cell system is limited in this
model. Some advanced fuel cells used in automotive engineering
can change their electricity generation according to the electricity
demand. In this case, the fuel cell should be modeled by using other
dynamic models. However, in general, the fuel cell system used in
building applications is controlled based on a preset schedule, and
this kind of control algorithm is applied to this model.
5. Conclusion
In building cogeneration applications, the main issues for engineers are to determine the size of the fuel cell systems and establish
operating sequences because the cogeneration system cannot meet
both heat and power demands in real time. A general method to
overcome this limitation is to use an optimization technique developed based on simulations.
In this research, a simplied PEM fuel cell model, which can be
easily implemented in whole-building simulation programs, was
suggested for commercial fuel cell packages, and the model was
validated by experimental data and established research outcomes.
The model showed good agreement with measurements. In addition, a simple application of this model is presented.
Unlike other fuel cell models, this empirical model attempts to
predict the performance of the fuel system without solving complex
electrochemical reactions. Correspondingly, this allows the model
to be implemented in whole-building simulation programs.
Although some existing models can be used with building simulation programs, they have some limitations when the commercial
fuel cell system is packaged and hard or impossible to disassemble. Disassembling packaged commercial units is often difcult
or undesirable owing to safety and security concerns. This study
suggested a simpler fuel cell model for that case. For those cases,
this study utilizes new terms for part-load ratio and AC power to
model the performance, which are controlled variables in commercial fuel cell systems for capacity control. Consequently, the
model is composed of only seven variables, which can be measured without disassembly. In addition, dimensionless heat and
power variables were introduced in order to consider performance
degradation after installation.
In practice, the success of a fuel cell cogeneration system
depends on how the system is sized and operated. With proper
sizing and operation, a cogeneration system provides savings in
primary energy and cost. When detailed modeling is possible, the
Annex model [40] is the rst choice, but when it is not applicable,
the model developed herein can be used in practical applications
to establish sizing and operation strategies of fuel cell cogeneration
systems both before and after installation.
Acknowledgements
This work was supported by a National Research Foundation
of Korea (NRF) grant funded by the Korean government (No.
2015R1A2A1A05001726).
References
[1] ASHRAE, ASHRAE Handbook: HVAC Systems and Equipment, American
Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc., n.d.,
Atlanta, 2008.
[2] G. Angrisani, C. Roselli, M. Sasso, Distributed microtrigeneration systems,
Prog. Energy Combust. Sci. 38 (2012) 502521, http://dx.doi.org/10.1016/j.
pecs.2012.02.001.
[3] ENERGYSTAR, Source Energy ENERGY STAR PortfolioManager Technical
Reference, 2013, pp. 117.

[4] T. Elmer, M. Worall, S. Wu, S.B. Riffat, Fuel cell technology for domestic built
environment applications: state of-the-art review, Renew. Sustain. Energy
Rev. 42 (2015) 913931, http://dx.doi.org/10.1016/j.rser.2014.10.080.
[5] Y. Wang, K.S. Chen, J. Mishler, S.C. Cho, X.C. Adroher, A review of polymer
electrolyte membrane fuel cells: technology, applications, and needs on
fundamental research, Appl. Energy. 88 (2011) 9811007, http://dx.doi.org/
10.1016/j.apenergy.2010.09.030.
[6] J.J. Hwang, M.L. Zou, Development of a proton exchange membrane fuel cell
cogeneration system, J. Power Sources 195 (2010) 25792585, http://dx.doi.
org/10.1016/j.jpowsour.2009.10.087.
[7] K. Maeda, K. Masumoto, A. Hayano, A study on energy saving in residential
PEFC cogeneration systems, J. Power Sources 195 (2010) 37793784, http://
dx.doi.org/10.1016/j.jpowsour.2009.12.075.
[8] N. Briguglio, M. Ferraro, G. Brunaccini, V. Antonucci, Evaluation of a low
temperature fuel cell system for residential CHP, Int. J. Hydrogen Energy 36
(2011) 80238029, http://dx.doi.org/10.1016/j.ijhydene.2011.01.050.
[9] M. Radulescu, O. Lottin, M. Feidt, C. Lombard, D. Le Noc, S. Le Doze,
Experimental results with a natural gas cogeneration system using a polymer
exchange membrane fuel cell, J. Power Sources 159 (2006) 11421146, http://
dx.doi.org/10.1016/j.jpowsour.2005.11.037.
[10] M.B. Gunes, M.W. Ellis, Evaluation of energy, environmental, and economic
characteristics of fuel cell combined heat and power systems for residential
applications, J. Energy Resour. Technol. 125 (2003) 208220, http://dx.doi.
org/10.1115/1.1595112.
[11] D.D. Massie, D.D. Boettner, C.A. Massie, Residential experience with proton
exchange membrane fuel cell systems for combined heat and power, J. Fuel
Cell Sci. Technol. 2 (2005) 263, http://dx.doi.org/10.1115/1.2041668.
[12] H. Koyanagi, H. Hukao, Study on energy conservation and economical
condition of PEFC Apartment House, ASHRAE Trans. (2009) 629637,
http://scholar.google.com/scholar?hl=en&btnG=Search&q=intitle:Study+on
+Energy+Conservation+and+Economical+Condition+of+?+PEFC+Apartment
+House+?#0 (accessed 22.11.14).
[13] H. Aki, Y. Taniguchi, I. Tamura, A. Kegasa, H. Hayakawa, Y. Ishikawa, et al., Fuel
cells and energy networks of electricity, heat, and hydrogen: a demonstration
in hydrogen-fueled apartments, Int. J. Hydrogen Energy 37 (2012)
12041213, http://dx.doi.org/10.1016/j.ijhydene.2011.10.021.
[14] W. Bendaikha, S. Larbi, M. Bouziane, Feasibility study of hybrid fuel cell and
geothermal heat pump used for air conditioning in Algeria, Int. J. Hydrogen
Energy 36 (2011) 42534261, http://dx.doi.org/10.1016/j.ijhydene.2010.09.
058.
[15] A. Adam, E.S. Fraga, D.J.L. Brett, Options for residential building services
design using fuel cell based micro-CHP and the potential for heat integration,
Appl. Energy 138 (2014) 685694, http://dx.doi.org/10.1016/j.apenergy.2014.
11.005.
[16] E. Jannelli, M. Minutillo, A. Perna, Analyzing microcogeneration systems
based on LT-PEMFC and HT-PEMFC by energy balances, Appl. Energy 108
(2013) 8291, http://dx.doi.org/10.1016/j.apenergy.2013.02.067.
[17] M. Jradi, S. Riffat, Tri-generation systems: energy policies, prime movers,
cooling technologies, congurations and operation strategies, Renew. Sustain.
Energy Rev. 32 (2014) 396415, http://dx.doi.org/10.1016/j.rser.2014.01.039.
[18] I. Pilatowsky, R.J. Romero, C.A. Isaza, S.A. Gamboa, W. Rivera, P.J. Sebastian,
et al., Simulation of an air conditioning absorption refrigeration system in a
co-generation process combining a proton exchange membrane fuel cell, Int.
J. Hydrogen Energy 32 (2007) 31743182, http://dx.doi.org/10.1016/j.
ijhydene.2006.03.016.
[19] T.A.H. Ratlamwala, M.A. Gadalla, I. Dincer, Thermodynamic analyses of an
integrated PEMFCTEARS-geothermal system for sustainable buildings,
Energy Buildings 44 (2012) 7380, http://dx.doi.org/10.1016/j.enbuild.2011.
10.017.
[20] S.-Y. Jo, J.-Y. Park, J.-W. Jeong, Applicability of a fuel cell for primary energy
saving in a liquid desiccant and evaporative cooling assisted 100% outdoor air
system, in: 2nd Asia Conf. Int. Build. Perform. Simul. Assoc., 2014, pp.
870877.
[21] L. Jrissen, Residential energy supply: fuel cells, in: Encycl. Electrochem.
Power Sources, Elsevier, 2009, pp. 108123, http://dx.doi.org/10.1016/B978044452745-5.00346-4.
[22] S. Oh, K. Kim, S. Oh, H. Kwak, Optimal operation of a 1-kW PEMFC-based CHP
system for residential applications, Appl. Energy 95 (2012) 93101, http://dx.
doi.org/10.1016/j.apenergy.2012.02.019.
[23] A. Arsalis, M.P. Nielsen, S.K. Kr, Application of an improved operational
strategy on a PBI fuel cell-based residential system for Danish single-family
households, Appl. Therm. Eng. 50 (2013) 704713, http://dx.doi.org/10.1016/
j.applthermaleng.2012.07.025.
[24] K.A. Pruitt, R.J. Braun, A.M. Newman, Establishing conditions for the economic
viability of fuel cell-based, combined heat and power distributed generation
systems, Appl. Energy 111 (2013) 904920, http://dx.doi.org/10.1016/j.
apenergy.2013.06.025.
[25] W. Yang, Y. Zhao, V. Liso, N. Brandon, Optimal design and operation of a
syngas-fuelled SOFC micro-CHP system for residential applications in
different climate zones in China, Energy Buildings 80 (2014) 613622, http://
dx.doi.org/10.1016/j.enbuild.2014.05.015.
[26] V. Dorer, A. Weber, Energy and CO2 emissions performance assessment of
residential micro-cogeneration systems with dynamic whole-building
simulation programs, Energy Convers. Manage. 50 (2009) 648657, http://dx.
doi.org/10.1016/j.enconman.2008.10.012.

S.-W. Ham et al. / Energy and Buildings 107 (2015) 213225


[27] E. Entchev, L. Yang, M. Ghorab, E.J. Lee, Simulation of hybrid renewable
microgeneration systems in load sharing applications, Energy 50 (2013)
252261, http://dx.doi.org/10.1016/j.energy.2012.11.046.
[28] F. Musio, F. Tacchi, L. Omati, P. Gallo Stampino, G. Dotelli, S. Limonta, et al.,
PEMFC system simulation in MATLAB-Simulink environment, Int. J.
Hydrogen Energy 36 (2011) 80458052, http://dx.doi.org/10.1016/j.ijhydene.
2011.01.093.
[29] C.A. Ramos-Paja, R. Giral, L. Martinez-Salamero, J. Romano, A. Romero, G.
Spagnuolo, A PEM fuel-cell model featuring oxygen-excess-ratio estimation
and power-electronics interaction, Ind. Electron. IEEE Trans. 57 (2010)
19141924, http://dx.doi.org/10.1109/TIE.2009.2026363.
[30] L. Barelli, G. Bidini, F. Gallorini, A. Ottaviano, Dynamic analysis of
PEMFC-based CHP systems for domestic application, Appl. Energy 91 (2012)
1328, http://dx.doi.org/10.1016/j.apenergy.2011.09.008.
[31] M.S. Behzadi, M. Niasati, Comparative performance analysis of a hybrid
PV/FC/battery stand-alone system using different power management
strategies and sizing approaches, Int. J. Hydrogen Energy 1852 (2014) 111,
http://dx.doi.org/10.1016/j.ijhydene.2014.10.097.
[32] A. Ferguson, V. Ismet Ugursal, Fuel cell modelling for building cogeneration
applications, J. Power Sources 137 (2004) 3042, http://dx.doi.org/10.1016/j.
jpowsour.2004.05.021.
[33] P. Corbo, F. Migliardini, O. Veneri, Experimental analysis and management
issues of a hydrogen fuel cell system for stationary and mobile application,
Energy Convers. Manage. 48 (2007) 23652374, http://dx.doi.org/10.1016/j.
enconman.2007.03.009.
[34] A. Saadi, M. Becherif, A. Aboubou, M.Y. Ayad, Comparison of proton exchange
membrane fuel cell static models, Renew. Energy 56 (2013) 6471, http://dx.
doi.org/10.1016/j.renene.2012.10.012.
[35] A. Contreras, F. Posso, E. Guervos, Modelling and simulation of the utilization
of a PEM fuel cell in the rural sector of Venezuela, Appl. Energy 87 (2010)
13761385, http://dx.doi.org/10.1016/j.apenergy.2009.05.040.
[36] M.V. Moreira, G.E. da Silva, A practical model for evaluating the performance
of proton exchange membrane fuel cells, Renew. Energy 34 (2009)
17341741, http://dx.doi.org/10.1016/j.renene.2009.01.002.
[37] TRNSYS 17, Mathematical Reference, Solar Energy Laboratory, University of
Wisconsin-Madison, 2013.
[38] I. Beausoleil-Morrison, A. Ferguson, B. Grifth, N. Kelly, F. Marchal, A. Weber,
Specications for Modelling Fuel Cell and Combustion-Based Residential
Cogeneration Devices within Whole-Building Simulation Programs, A Report
of Subtask B of FC + COGEN-SIM, The Simulation of Building-Integrated Fuel
Cell and Other Cogeneration Systems: Annex 42 of the International Energy,
Annex 42 of the International Energy Agency Energy Conservation in
Buildings and Community System, 2007.
[39] I. Beausoleil-Morrison, A. Schatz, F. Marchal, A model for simulating the
thermal and electrical production of small-scale solid-oxide fuel cell
cogeneration systems within building simulation programs, HVAC&R Res. 12
(2006) 641667, http://dx.doi.org/10.1080/10789669.2006.10391199.
[40] G. Johnson, I. Beausoleil-Morrison, B. Strathearn, E. Thorsteinson, T.
Mackintosh, The calibration and validation of a model for simulating the
thermal and electrical performance of a 1 kWAC proton-exchange membrane
fuel-cell micro-cogeneration device, J. Power Sources 221 (2013) 435446,
http://dx.doi.org/10.1016/j.jpowsour.2012.08.035.
[41] S. Ahmed, D.D. Papadias, R.K. Ahluwalia, Conguring a fuel cell based
residential combined heat and power system, J. Power Sources 242 (2013)
884894, http://dx.doi.org/10.1016/j.jpowsour.2013.01.034.
[42] T. Guan, P. Alvfors, G. Lindbergh, Investigation of the prospect of energy
self-sufciency and technical performance of an integrated PEMFC (proton
exchange membrane fuel cell), dairy farm and biogas plant system, Appl.
Energy 130 (2014) 685691, http://dx.doi.org/10.1016/j.apenergy.2014.04.
043.
[43] M. Gandiglio, A. Lanzini, M. Santarelli, P. Leone, Design and optimization of a
proton exchange membrane fuel cell CHP system for residential use, Energy
Buildings 69 (2014) 381393, http://dx.doi.org/10.1016/j.enbuild.2013.11.
022.
[44] M.Y. El-Sharkh, A. Rahman, M.S. Alam, P.C. Byrne, A.A. Sakla, T. Thomas, A
dynamic model for a stand-alone PEM fuel cell power plant for residential

[45]
[46]
[47]

[48]

[49]

[50]

[51]
[52]

[53]
[54]

[55]
[56]

[57]

[58]

[59]

[60]

[61]

[62]

[63]
[64]

[65]

225

applications, J. Power Sources 138 (2004) 199204, http://dx.doi.org/10.1016/


j.jpowsour.2004.06.037.
DOE, EnergyPlus Version 8.1, 2013, http://apps1.eere.energy.gov/buildings/
energyplus/energyplus about.cfm
TRNSYS 17, Solar Energy Laboratory, University of Wisconsin-Madison, 2013.
L. Yang, E. Entchev, M. Ghorab, E.J. Lee, E.C. Kang, Energy and cost analyses of
a hybrid renewable microgeneration system serving multiple residential and
small ofce buildings, Appl. Therm. Eng. 65 (2014) 477486, http://dx.doi.org/
10.1016/j.applthermaleng.2014.01.049.
J. St.-Pierre, Overview performance and operational conditions, in: Encycl.
Electrochem. Power Sources, Elsevier, 2009, pp. 901911, http://dx.doi.org/
10.1016/B978-044452745-5.00233-1.
J.-J. Hwang, Transient efciency measurement of a combined heat and power
fuel cell generator, J. Power Sources 223 (2013) 325335, http://dx.doi.org/10.
1016/j.jpowsour.2012.09.086.
V. Boscaino, R. Miceli, G. Capponi, G. Ricco Galluzzo, A review of fuel cell
based hybrid power supply architectures and algorithms for household
appliances, Int. J. Hydrogen Energy 39 (2014) 11951209, http://dx.doi.org/
10.1016/j.ijhydene.2013.10.165.
J. Larminie, A. Dicks, Fuel Cell Systems Explained, 2nd ed., Wiley, 2003.
Z. Qi, Proton exchange membrane fuel cells-systems, in: Encycl. Electrochem.
Power Sources, Elsevier, 2009, pp. 890900, http://dx.doi.org/10.1016/B978044452745-5.00238-0.
W. Cho, K.-S. Lee, A simple sizing method for combined heat and power units,
Energy 65 (2014) 123133, http://dx.doi.org/10.1016/j.energy.2013.11.085.
Y. Tang, W. Yuan, M. Pan, Z. Li, G. Chen, Y. Li, Experimental investigation of
dynamic performance and transient responses of a kW-class PEM fuel cell
stack under various load changes, Appl. Energy 87 (2010) 14101417, http://
dx.doi.org/10.1016/j.apenergy.2009.08.047.
EG&G Technical Services, Inc., Fuel Cell Handbook, 7th ed., U.S. Department of
Energy, 2004.
J.-J. Hwang, Thermal control and performance assessment of a proton
exchanger membrane fuel cell generator, Appl. Energy 108 (2013) 184193,
http://dx.doi.org/10.1016/j.apenergy.2013.03.025.
Z. Qi, H. Tang, Q. Guo, B. Du, Investigation on saw-tooth behavior of PEM
fuel cell performance during shutdown and restart cycles, J. Power Sources
161 (2006) 864871, http://dx.doi.org/10.1016/j.jpowsour.2006.04.137.
A.J.L. Verhage, J.F. Coolegem, M.J.J. Mulder, M.H. Yildirim, F.A. de Bruijn,
30,000 h operation of a 70 kW stationary PEM fuel cell system using hydrogen
from a chlorine factory, Int. J. Hydrogen Energy 38 (2013) 47144724, http://
dx.doi.org/10.1016/j.ijhydene.2013.01.152.
M.W. Davis, A.H. Fanney, M.J. LaBarre, K.R. Henderson, B.P. Dougherty,
Parameters affecting the performance of a residential-scale stationary fuel
cell system, J. Fuel Cell Sci. Technol. 4 (2006) 109115, http://dx.doi.org/10.
1115/1.2713767.
U. Arndt, I. Beausoleil-Morrison, M. Davis, W. Dhaeseleer, V. Dorer, E. Entchev,
et al., Experimental Investigation of Residential Cogeneration Devices and
Calibration of Annex 42 Models, A Report of Subtask B of FC + COGEN-SIM, The
Simulation of Building-Integrated Fuel Cell and Other Cogeneration Systems:
Annex 42 of the International Energy, Annex 42 of the International Energy
Agency Energy Conservation in Buildings and Community System, 2007.
M.-H. Kim, J.-S. Park, J.-W. Jeong, Energy saving potential of liquid desiccant in
evaporative-cooling-assisted 100% outdoor air system, Energy 59 (2013)
726736, http://dx.doi.org/10.1016/j.energy.2013.07.018.
M.-H. Kim, J.-Y. Park, M.-K. Sung, A.-S. Choi, J.-W. Jeong, Annual operating
energy savings of liquid desiccant and evaporative-cooling-assisted 100%
outdoor air system, Energy Buildings 76 (2014) 538550, http://dx.doi.org/10.
1016/j.enbuild.2014.03.006.
R.S. Figliola, D.E. Beasley, Theory and Design for Mechanical Measurements,
5th ed., John Wiley & Sons, Inc., Hoboken, NJ, 2011.
M. Hydeman, N. Webb, P. Sreedharan, S. Blanc, Development and testing of a
reformulated regression-based electric chiller model, in: ASHRAE Trans.,
2002, pp. 11181127, http://www.scopus.com/inward/record.url?eid=2-s2.00036459461&partnerID=tZOtx3y1
MathWorks, MATLAB, 2014.

You might also like