Está en la página 1de 11

Building and Environment 54 (2012) 186e196

Contents lists available at SciVerse ScienceDirect

Building and Environment


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

Verication and validation of EnergyPlus phase change material model


for opaque wall assemblies
Paulo Cesar Tabares-Velasco*, Craig Christensen 1, Marcus Bianchi 2
National Renewable Energy Laboratory, 1617 Cole Blvd. MS 5202, Golden, CO 80401, USA

a r t i c l e i n f o a b s t r a c t

Article history: Phase change materials (PCMs) represent a technology that may reduce peak loads and HVAC energy
Received 8 December 2011 consumption in buildings. A few building energy simulation programs have the capability to simulate
Received in revised form PCMs, but their accuracy has not been completely tested. This study shows the procedure used to verify
16 February 2012
and validate the PCM model in EnergyPlus using a similar approach as dictated by ASHRAE Standard 140,
Accepted 18 February 2012
which consists of analytical verication, comparative testing, and empirical validation. This process was
valuable, as two bugs were identied and xed in the PCM model, and version 7.1 of EnergyPlus will have
Keywords:
a validated PCM model. Preliminary results using whole-building energy analysis show that careful
Phase change materials
Validation
analysis should be done when designing PCMs in homes, as their thermal performance depends on
Building energy simulation several variables such as PCM properties and location in the building envelope.
Building envelope 2012 Elsevier Ltd. All rights reserved.
PCM
Storage

1. Introduction impregnated PCM wallboards [9,10]. More recent studies of PCM


wallboards have analyzed building elements containing micro-
Phase change materials (PCMs) represent a technology that may encapsulated PCMs. One compared the performance of a detailed
reduce peak loads and HVAC energy consumption in buildings. model considering a solideliquid interface against a simpler model
PCM research has been broad on many heat transfer applications using equivalent heat capacity model. Overall, the equivalent heat
during recent decades, resulting in a considerable amount of capacity method performed better than experimental data [11].
literature about PCM properties, indoor temperature stabilization Another study found the optimum time and space steps for
potential, and peak load reduction potential. PCM-oriented a specic PCM using heat capacity method, DSC material property
research on buildings has investigated two primary applications: data, and an implicit nite difference model [12]. A follow-up study
passive and active building systems [1e5]. Previous PCM studies concluded that hysteresis should be considered in the numerical
have shown that important benets are related to thermal comfort, simulations to improve the overall accuracy of the model [13].
energy savings, and perhaps HVAC downsizing when thermal Numerical research on PCM-enhanced building enclosure systems
storage is added to buildings [6]. has followed a similar approach to wallboards, analyzing building
Early numerical studies analyzed PCM wallboards during late enclosure systems containing PCM between two layers of insu-
1970s and 1980s, mainly for PCM-impregnated wallboard appli- lation [14,15].
cations [7,8]. Later numerical studies used PCM thermal properties These studies focused on understanding the physics behind
obtained from differential scanning calorimetry (DSC) techniques PCMs. They validated models and investigated heat ux reduction
and from measurements of temperature proles across potential using special laboratory and small-scale test rooms. Other
studies have attempted to estimate potential energy savings
through building energy simulation (BES). Energy simulation
* Corresponding author. Tel.: 1 303 384 7591; fax: 1 303 384 7495. studies analyzing energy and peak load benets from PCMs have
E-mail addresses: paulo.tabares@nrel.gov (P.C. Tabares-Velasco), used commercial BES software such as CoDyBa [16], EnergyPlus
Craig.Christensen@nrel.gov (C. Christensen), Marcus.Bianchi@owenscorning.com [17], ESP-r [18e20], and TRNSYS [21e23]. The implemented models
(M. Bianchi).
1 varied from early PCM models [21,24], to empirical models using an
Tel.: 1 303 384 7510; fax: 1 303 384 7540.
2
Present address: Owens Corning Science Technology Center, 2790 Columbus equivalent heat transfer coefcient [22], to fully implemented
Road, Granville, OH 43023-1200, USA. Tel.: 1 720 258 6822. nite difference models [17,23] and control volume models [18].

0360-1323/$ e see front matter 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.buildenv.2012.02.019
P.C. Tabares-Velasco et al. / Building and Environment 54 (2012) 186e196 187

Predictions from these studies included no signicant energy where ki k(Tj1 i ) if thermal conductivity is variable,
benets [17]; improved thermal comfort and decreased peak load T temperature, i node being modeled, i1 adjacent node to
[21,24]; and 90% reduction of heating energy demand during the interior of construction, i1 adjacent node to exterior of
heating season [18]. Likewise, an energy simulation study focusing construction, j1 new time step, j previous time step, Dt time
on building enclosure systems predicted a 19e57% peak load step, Dx nite difference layer thickness, Cp specic heat of
reduction for an attic system consisting of a PCM sandwiched material, and r density of material.
between two conventional insulation layers [15]. In the CondFD algorithm, all elements are divided or discretized
In conclusion, PCMs have different benets depending on automatically using Eq. (2), which depends on a space discretiza-
quantity and types (phase change temperature, energy storage tion constant (c), the thermal diffusivity of the material (a), and the
capacity), locations (drywall, walls, attic, and oor) and climates time step. Users can leave the default space discretization value of 3
(heating and/or cooling performance). Therefore, there are clear (equivalent to a Fourier number (Fo) of 1/3) or input other values.
differences between the methods and results from previous r
research efforts as: (1) the reviewed building energy simulation p a$Dt
studies, except a TRNSYS model that works for certain exterior PCM
Dx c$a$Dt (2)
Fo
applications, did not perform comprehensive model PCM valida-
For the PCM algorithm, the CondFD method is coupled with an
tion [25] and (2) previous studies do not cover a wide range of PCM
enthalpyetemperature function (Eq. (3)) that the user inputs to
types, locations, and climates. Thus, there is a simulation and
account for enthalpy changes during phase change [17]. The
analysis gap with respect to PCM benets and modeling. The
enthalpyetemperature function is used to develop an equivalent
objective of this study is to verify, validate, and improve the Ener-
specic heat at each time step. The resulting model is a modied
gyPlus PCM model, which has not been fully validated. This
version of the enthalpy method [17].
procedure will be performed using analytical verication,
comparative verication, and empirical validation of three PCM h hT (3)
applications:

1) PCM distributed in drywall, hji  hij1


Cp* T j j1
(4)
2) PCM distributed in brous insulation, and Ti  Ti
3) Thin, concentrated PCM layers.
where h enthalpy.

3. Verication and validation


2. EnergyPlus PCM model
Several studies have included validation exercises to assess the
EnergyPlus PCM algorithm uses a one-dimensional conduction credibility of PCM models [9e11,13,14]. Before validation, most
nite difference (CondFD) solution algorithm.3 Recently, CondFD obtained the thermophysical material properties of the samples
was validated against multiple test suites [26]. This step was through laboratory testing. However, there are differences between
necessary before any serious PCM validation efforts could be made, the validation procedures followed in each study; approaches range
as the PCM model relies on CondFD. Several xes were made during from qualitatively comparing surface temperature graphs
the verication process. [9e11,13,14], heat uxes [14,27], or room air temperatures [28,29],
The PCM model considered for verication and validation is to validations that contained quantication of error caused by
a modied version of EnergyPlus 6.0.0.037 that includes all xes spatial and time discretization [12] and quantication of an average
from the CondFD verication [26]. This version will be referred to as error [30].
v6 in this publication. The new and improved PCM/CondFD This study follows a similar approach to ASHRAE Standard 140,
model will be referred to as v7.1, as time constraints prevented as the entire verication and validation scheme is divided into
the xes, which will be described later, from being included in the analytical verication, comparative testing, and empirical valida-
EnergyPlus version 7 release. tion [31]. Thus, the PCM model was veried using (1) an analytical
The CondFD algorithm in EnergyPlus uses an implicit nite solution from the Stefan Problem as generalized by Franz Neumann
difference scheme, where the user can select Crank-Nicholson or [32], (2) a numerical solution of a sine wave problem using a veri-
fully implicit. Eq. (1) shows the calculation method for the fully ed PCM model from Heating 7.3 [33], and (3) experimental data
implicit scheme inside a homogeneous material. from a hot box apparatus [34]. Moreover, the PCM model is tested
    for building envelope applications only, as heat diffusion and
j1 j1 j1 j1
Ti
j1 j
 Ti Ti1  Ti Ti1  Ti storage in the walls and attic were considered in the verication
Cp rDX kW kE (1)
Dt DX DX and validation. This follows the previous approach used to verify
and validate the EnergyPlus CondFD algorithm [26]. It is important
where,
to mention that: (1) the PCM model was not fully validated earlier,
  and (2) the validation and verication process is not calibrating or
j1 j1
ki1 ki evaluating the model, because the model input would not be
kW
2 changed or tuned as a result of the verication. As a result, the PCM
verication and validation process found stability problems with
 
j1
ki1 ki
j1 the v6 PCM model that were xed using an automatic under-
kE relaxation factor.
2
3.1. Analytical verication: Stefan Problem

3
The default conduction transfer function (CTF) solution algorithm in EnergyPlus The rst verication test was the analytical solution to the Ste-
cannot simulate materials with variable properties such as PCMs. fan Problem, which originally posed the problem for the thickness
188 P.C. Tabares-Velasco et al. / Building and Environment 54 (2012) 186e196

of polar ice. The solution was developed by Neumann for a semi-


innite wall. The assumptions are: (1) the wall is initially at
uniform temperature with PCM in liquid form above the melting
temperature (T(x,t) > Tmelting), (2) suddenly at time 0, the wall
outside surface temperature drops to 0  C (T(0,t) 0  C), (3) PCM
density is equal and constant in both the liquid and solid phases, (4)
PCM thermal conductivity and specic heat are constant in each
phase and can differ between the solid and liquid phases, and (5)
PCM has a xed melting temperature (no melting/freezing range)
[32].
The analytical solution was used to test the v6 PCM model for
the three PCM applications, assuming the properties and charac-
teristics shown in Table 1. Specic details include: (1) a zone with
one wall was used and (2) the walls were initially at a homogenous
temperature of 50  C and had a melting temperature range of Fig. 1. Outside surface heat ux calculated using analytical solution (Analytical) and
29.4e29.6  C to approximate the xed melting temperature PCM model v6 for three node spacing values (dx, dx/3, dx/9).
assumption in the Stefan Problem. The semi-innite wall was
simulated by a thick wall, so all surface temperatures were set in
EnergyPlus using the OtherSideCoefcient feature; the indoor air All three numerical and analytical solutions show similar results
temperature was set to the desired temperature and the inside and the PCM model behaves similarly between cases. Thus, detailed
convective heat coefcient was set to 1000 W/m2C. More details results are only shown here for the PCM-enhanced insulation.
about similar approaches are in the literature [26]. Figs. 1and 2 show results for three node spacing values: 0.015 m
As Table 1 shows, each PCM application was simulated in (the default, dx), 0.005 m (dx/3), and 0.00167 m (dx/9), which are
EnergyPlus as a wall with a thickness representative of typical obtained using Fourier numbers equal to 1/3, 3, and 26 respectively.
material congurations, with further details as follows: These numbers seem high and would have caused problems in an
explicit scheme. However, this study used the fully implicit scheme,
1. The distributed PCM in the insulation represents a 0.105 m which can work with higher Fourier numbers, and steadily decayed
layer of insulation that could be installed in a wall cavity or to convergence [35,36]. Node spacing values refer to the distance
attic. In this study, the simulated wall was 1.5 m thick to meet between adjacent nodes that discretized the analyzed wall. More
the semi-innite wall condition required by the Neumann information about the discretization of walls in EnergyPlus can be
solution. However, only the rst 0.105 m was considered when found in the literature [26,37].
comparing simulations to the analytical solution, as this Fig. 1 shows how the v6 PCM model solution for the default
represents the thickness of a PCM-enhanced portion in a wall node space (dx) oscillates around the analytical solution: the
cavity. temperature from the PCM model suddenly increases, then exhibits
2. The distributed PCM in drywall represents a 0.02 m layer of almost constant or at behavior around the analytical solution. This
drywall. In this study, the simulated wall was 0.3 m thick to behavior was observed at the default node spacing value in each
meet the semi-innite wall condition required by the Neu- solution for the three PCM applications, and in each solution for the
mann solution. However, only the rst 0.02 m was considered three PCM applications, but with a lesser magnitude, in the other
when comparing simulations to the analytical solution, as this two node spacing cases (dx/3 and dx/9). The reason for this is
represents the PCM-enhanced portion of the wall. mainly numerical; the solution requires smaller node spacing than
3. The concentrated PCM represents a 0.005 m layer that could be the default node space in EnergyPlus to properly simulate heat
installed between two layers of insulation or between insu- wave propagation through the PCM with a xed melting temper-
lation and drywall. In this study, the simulated wall contained ature. In this case, the oscillations were almost eliminated with
only concentrated PCM and was 0.155 m thick to meet the node spacing nine times smaller than the default node spacing.
semi-innite wall condition required by the Neumann solution. This study also analyzed the model stability and convergence by
However, only the rst 0.005 m was considered when recording the number of iterations required for the PCM model to
comparing simulations to the analytical solution, as this converge. In the three discretization cases, the solution did not
represent the PCM-enhanced portion of the wall.

Table 1
PCM strategies tested in analytical verication.

Properties PCM- PCM-Drywall Concentrated


insulation thin PCM layer
% by weight of 20% 30% 100%
microencapsulated
PCM in wall layer
Equivalent latent heat of wall 34 33 80
layer (kJ/kg)
Thermal conductivity (W/m K) 0.03365 0.2300 0.1600
Density (kg/m3) 29 1000 850
Specic heat (J/kg K) 960 1400 2500
Thickness (m) 0.105 0.020 0.005
Default node spacing 0.015 0.005 0.005 Fig. 2. Node temperature 0.105 m from outside surface calculated using analytical
solution (Analytical) and PCM model v6 for three node spacing values (dx, dx/3, dx/9).
P.C. Tabares-Velasco et al. / Building and Environment 54 (2012) 186e196 189

converge when a node entered or left the melting range. Thus, after
CondFD passed the maximum allowable iterations, the solution did
not meet the convergence criteria. Therefore, the PCM model used
the values from the last iterations and moved to the next time step,
even though it did not converge. This problem did not cause any
signicant accuracy issues for the three cases tested, because of the
small time step used, but did increase the run time signicantly,
taking 6, 15, and 66 min to simulate the wall for 2 days in the dx, dx/
3, dx/9 scenarios, respectively.
The convergence problem was solved by adding an automatic
and dynamic under-relaxation factor after a determined number of
iterations. This new under-relaxation factor did not affect the
accuracy of the PCM model and reduced the run time 60e90%
relative to the v6 model. It is important to mention that the
semi-innite wall problem stated by the Neumann solution is Fig. 3. Calculated PCM-Insulation middle node temperature (Tmid) using comparative
demanding and challenging from a numerical point of view. A very software (H73), EnergyPlus PCM model with default node spacing (E), smaller node
space (Edx/3), 2- and 4-min time steps (E2 min, E4 min) and without PCM
close match was obtained only with Heating 7.3 using a time step of
(ENoPCM).
0.01 s and node spacing roughly 0.01 times the default size in
EnergyPlus. These values are stricter than those recommended in
previous studies [25]; they are likely overly strict and impractical  Indoor convective heat transfer coefcient: hi 5 W/m2 K
for modeling more realistic PCM-enhanced wall congurations in  Outside air temperature: 25  C at t 0; Varying sinusoidal over
annual building simulations. As a result, the PCM model must be 24 h with a range 10e40  C
veried with more realistic wall and boundary conditions. There is  Outdoor convective heat transfer coefcient: hout 20 W/m2 K
no analytical solution for this case, so another veried numerical  Wall initial conditions: homogenous temperature of 25  C
model needs to be used to conduct comparative testing.  Ideal/pure PCM with a xed melting temperature

3.2. Comparative testing against Heating 7.3 The two wall applications veried in this test are shown in
Table 2. This test evaluates only two applications, because no
The analytical verication enabled us to detect run time issues signicant information was obtained from conducting three tests in
and identify node spacing requirements for an extreme circum- the analytical test. In addition, the concentrated thin PCM layer
stance with a large and sudden temperature drop. The last test is application will be tested with experimental data in Section 3.3.
useful as a quality assurance tool but not as a performance-based Fig. 3 through Fig. 5 shows the results of the comparative veri-
tool. Thus, the next step in the verication process was compara- cation for the distributed PCM in insulation. All gures show
tive testing relative to the ideal PCM model in Heating 7.3. This is calculated values using Heating 7.3 (H73), EnergyPlus v7.1 PCM
a multidimensional, nite difference heat conduction program that model with default space discretization and 1-min time step (E),
can simulate materials with variable thermal properties and other EnergyPlus v7.1 with no PCM and 1-min time step (ENoPCM),
features [33]. This software has been used before in multidimen- EnergyPlus v7.1 PCM model with smaller node spacing (Edx/3)
sional heat transfer problems with and without PCMs [14,38]. It and 1-min time step, EnergyPlus v7.1 PCM model using a 2-min
was also tested with the Neumann solution, which proved its time step (E2 min) and EnergyPlus v7.1 PCM model using a 4-min
accuracy at small time steps (0.01 s). The comparative verication time step (E4 min). Simulations with 2- and 4-min time steps
consisted of a 24-h sine temperature wave test with amplitude of used the default space discretization values. All EnergyPlus simu-
30  C representing outside temperature variation during a day, and lations were performed using the updated model (v7.1). Fig. 3
a 1-dimensional wall consisting of 1 cm wood, 10 cm ber insu- through Fig. 5 shows that discrepancies increase substantially
lation, and 1.5 cm drywall. This test represents a more realistic wall when time steps of 4 min (E4 min) or longer are used. Thus, all
conguration and boundary conditions than the analytical solution. simulations should use time steps shorter than 4 min.
Model space discretization in Heating 7.3 and boundary conditions
are:

 0.01 s time steps


 Node spacing 0.01 times the EnergyPlus default value (see
Table 2)
 Indoor air temperature: 25  C t  0

Table 2
PCMs tested in comparative verication.

Properties PCM-insulation PCM-Drywall


% by weight of 20 30
microencapsulated PCM
in wall layer
Equivalent latent heat (kJ/kg) 34 33
Melting temperature ( C) 26 25.05
Thermal conductivity (W/m K) 0.03365 0.1600
Density (kg/m3) 50 950
Specic heat (J/kg K) 960 840
Thickness (m) 0.100 0.015 Fig. 4. Inside surface temperature (Tin) calculated using comparative software (H73),
Default node spacing 0.015 0.005 EnergyPlus PCM model with default node spacing (E), smaller node space (Edx/3),
2- and 4-min time steps (E2 min, E4 min) and without PCM (ENoPCM).
190 P.C. Tabares-Velasco et al. / Building and Environment 54 (2012) 186e196

Table 4
Differences in net heat gain over 12 h between Heating 7.3 and other models.

Case Outside Inside Outside Inside


heat gain (kJ) heat gain (kJ) discrepancy discrepancy
with H7.3 (%) with H7.3 (%)
H73 216 43 e e
E 213 43 1.3 1.5
Edx/3 215 43 0.6 0.7
E2 min 214 46 0.9 7.8
E4 min 297 18 37.6 58.5
E noPCM 127 115 41.3 169.8

The PCM model for the case with PCM distributed in the drywall
Fig. 5. Inside heat ux (Qin) calculated using comparative software (H73), EnergyPlus has performance similar to the one shown in Table 4. Table 5 shows
PCM model with default node spacing (E), smaller node space (Edx/3), 2- and 4-min that all cases agree, having an almost zero heat ux coming into the
time steps (E2 min, E4 min) and without PCM (ENoPCM).
inside wall over the 24-h test.

Fig. 3 also shows the node temperature in the middle of the 3.3. Experimental: DuPont hot box experiment
insulation layer. All node locations for the middle node in the
insulation are in the middle of the insulation except the nodes for The last step in the verication and validation of the PCM model
the 2-min (E2 min) and 4-min (E4 min) time steps. These were was empirical validation. This step compared the PCM model with
linearly interpolated, as the automatic node spacing in EnergyPlus experimental data from the literature [34,39] for the concentrated
did not generate a node in the middle of the insulation. Figs. 3 and 4 PCM application. Fig. 6 is a schematic of the wall tested with the
show that the node temperature calculated by the EnergyPlus PCM PCM layer representing the DuPont Energain PCM product with
model oscillates around the temperature calculated by Heating 7.3. a melting temperature range centered around 21.7  C, a latent heat
This is the same behavior observed in the analytical test. The of 70 kJ/kg, and a variable thermal conductivity. Properties for this
oscillations attenuated and disappeared once smaller node spacing PCM are shown in the Appendix [39]. PCM properties in the
was used, which in this case was about 1/3 smaller than the default experimental study were provided by the manufacturer; thus,
size. Fig. 3 shows that during approximately the rst 4 h the middle there were no degree of freedom to calibrate EnergyPlus results
wall node temperature with PCM remained at almost constant to match measured values. The tests were performed in a hot box
temperature around the melting temperature (26  C). Once all the apparatus, where initially the cold box side was kept at 20  C and
PCM melted, the temperature started to increase as no more latent the hot side was kept at 20  C. At t 0, a heater in the hot box
storage was available, delaying the peak temperature by almost 2 h. started heating the air temperature inside the hot box for 7 h
Overall, the benets of PCM in this particular example are reected (heating stage). The nal inside wall temperature reached 24  C.
in Figs. 4 and 5, where the peak inside surface temperature and After that, the heater was turned off and the hot box slowly cooled
heat ux are attenuated by 0.5  C and 50%, respectively, and the to the initial temperature, 20  C (cooling stage) [34,39].
peak time is delayed almost 4 h. Initial testing of the PCM model, in this case with variable
Figs. 4 and 5 and show the impact distributing PCM in the thermal conductivity revealed another bug that occurred when
insulation has on the inner surface temperature and inside heat these two features were simulated at the same time. The thermal
ux for this example. The PCM model correctly calculates peak load conductivityetemperature array had larger dimensions than the
reduction and shift. Despite the small oscillations around the values allowable inputs for users. Thus, the variable thermal conductivity
calculated from Heating 7.3, the PCM model with default space was lling the empty values with temporary data from the PCM
discretization obtains values close to Heating 7.3 (see Table 3). enthalpy array. This bug was xed by sizing the variable thermal
Moreover, using smaller node spacing improves the agreement conductivity array correctly. Fig. 7 through Fig. 10 compares the
with Heating 7.3 and using a 2-min time step still has close results for the xed version to the original, which is marked Ev6.
agreement with Heating 7.3. Figs. 7e10 relate to the measurement points shown in Fig. 6. Results
Finally, Table 4 summarizes the discrepancies in the rst 12-h from the empirical validation are shown in Fig. 7 through Fig. 10 for
cycle using the same caption labels as in Fig. 3. The results agree experimental data (Experiment) with the uncertainty bars, Ener-
with Table 3: the default space discretization (E) and smaller node gyPlus v7.1 PCM model with variable thermal conductivity (E),
space (Edx/3) models with PCM agree with Heating 7.3 very EnergyPlus v7.1 PCM model without PCM (EnoPCM), previous
closely. The 2-min time step (E2 min) has looser agreement but is version of EnergyPlus (v6), and EnergyPlus v7.1 PCM model with
still close compared to the case without PCM which is not the case variable thermal conductivity (E) and 4-min time step. The PCM
when using the 4-min time step. model was further evaluated using a 2-min time step and a smaller
node space (Edx/3). These are not shown in Fig. 7 through Fig. 10

Table 3
Root Mean Square Error (RMSE) compared to Heating 7.3 values for PCM distributed Table 5
in Insulation. RMSE for PCM distributed in drywall compared to Heating 7.3.

Tin ( C) Tmin ( C) Qin (W/m2) Qout (W/m2) Tin ( C) Tmin ( C) Qin (W/m2) Qout (W/m2)
ENoPCM 0.45290 2.6470 2.2940 4.370 ENoPCM 5.6180 3.066 4.610 1.801
E 0.01919 0.1384 0.1406 1.885 E 0.0254 0.138 0.019 1.800
Edx/3 0.01564 0.0741 0.0771 0.357 Edx/3 0.0255 0.114 0.017 0.347
E2 min 0.03715 0.2180 0.2698 0.717 E2 min 0.0257 0.015 0.193 1.777
E4 min 0.30770 2.3040 1.4590 2.844 E4 min 0.0340 0.753 0.197 4.394
P.C. Tabares-Velasco et al. / Building and Environment 54 (2012) 186e196 191

Fig. 6. Tested wall conguration with dimensions (in cm) and location of temperature
(Temp) and heat ux measurements (HF).

Fig. 8. Point 5 temperature obtained from experimental data (Experiment), Ener-


to avoid overpopulating the graphs. No new results were observed
gyPlus v7.1 PCM model with variable thermal conductivity (E), EnergyPlus v7.1 PCM
besides those that were already found in Section 3.2 for the 2 and 4 model without PCM (EnoPCM), previous version of EnergyPlus (v6).
min time steps. The 2-min time steps and smaller node space
behaved similarly to the 1-min time step with default space dis-
cretization. In contrast, using a time step of 4 min introduces
show heat ux at measurement Points 1 and 5, respectively,
important errors in all gures that are consistent with the results
using the same legend as in Fig. 7. All models in Fig. 9 started with
shown in the analytical and numerical cases. All cases with
similar heat uxes and followed the same trend, but are outside the
different time steps used the same Fourier number, which makes
experimental uncertainty, a behavior that was also observed
the node space increase (Eq. (2)). This problem with the 4-min time
previously using WUFI and COMSOL [34], even when a smaller
step will be further analyzed in the future.
node space was used. However, measuring heat ux in this type of
Figs. 7 and 8 show the temperatures at Points 3 and 5. In both
interface is very challenging because radiation and convective
gures, Ev7.1 with PCMs agrees well with the experimental data
uxes are present in the air-wall surface interface. This might not
for the rst 8 h (heating stage). However, all models disagree with
be correctly simulated in EnergyPlus because the surface temper-
the experimental data after 8 h during the cooling stage of the
ature was input as a boundary condition.
problem. The disagreement shown is mainly due to the PCM
In contrast with Fig. 9, Fig. 10 shows that most models results
hysteresis, as the current PCM model in EnergyPlus does not
are within the uncertainty range of the measured heat ux in Point
simulate hysteresis. Thus, only the enthalpyetemperature infor-
5, located between the PCM and mineral wood during the heating
mation for the heating mode was input. This is also consistent with
process. As with the results shown in Figs. 7 and 8 the disagreement
the calculated values from Haavi et al., which simulated the same
increases once the PCM starts to cool down and hysteresis occurs. In
problem using PCM models in COMSOL and WUFI [34].
this case, the differences in the cooling range (starting around
Figs. 7 and 8 show that Ev6 completely disagrees with the data
Time 8 h) are on the order of 10e15%. In all cases, the previous
for roughly half the test because of the bug related to variable
PCM model completely missed the behavior of the PCM because of
thermal conductivity. The EnergyPlus simulation without PCM
the bug with variable thermal conductivity.
shows a peak reduction and time shift of roughly 0.5  C and 30 min,
Overall, this comparison to empirical data was important
respectively. These differences are small and caused mainly by the
because it used a real PCM product with variable thermal
slight excitation in the experiments (4  C).
conductivity to test the PCM model. Combined, the results from
In addition to temperature data used for empirical validation,
Sections 3.1e3.3 indicate that the v7.1 PCM model is performing
the experiments measured heat ux at two points. Figs. 9 and 10
well for PCM wall applications, except when the materials present

Fig. 7. Point 3 temperature obtained from experimental data (Experiment), EnergyPlus Fig. 9. Point 1 heat ux obtained from: experimental data (Experiment), EnergyPlus
v7.1 PCM model with variable thermal conductivity (E), EnergyPlus v7.1 PCM model v7.1 PCM model with variable thermal conductivity (E), EnergyPlus v7.1 PCM model
without PCM (EnoPCM), previous version of EnergyPlus with PCM (v6). without PCM (EnoPCM), previous version of EnergyPlus (v6).
192 P.C. Tabares-Velasco et al. / Building and Environment 54 (2012) 186e196

Table 7
PCMs used in modied ASHRAE Standard 140, Case 600.

Properties PCM- PCM-Drywall Concentrated


insulation thin PCM layer
% weight of 20% 30% 100%
microencapsulated
PCM in wall layer
Equivalent latent heat 34 33 130
(kJ/kg)
Melting temperature 28.5e30.5  C 22.5e24.5  C 28.5e30.5  C
range
Thermal conductivity 0.03365 0.23 0.16
(W/m K)
Density (kg/m3) 29 1000 850
Specic heat (J/kg K) 960 1400 2500
Location Walls and roof Walls Walls and roof
(5 cm insulation
0.5 cm PCM 5 cm
insulation)
Fig. 10. Point 5 heat ux obtained from experimental data (Experiment), EnergyPlus
v7.1 PCM model with variable thermal conductivity (E), EnergyPlus v7.1 PCM model
without PCM (EnoPCM), and previous version of EnergyPlus (v6).

building has interior loads equal to 200 W (60% radiative, 40%


convective) and a highly insulated slab to essentially eliminate
strong hysteresis. The scope of this verication and the hysteresis
thermal ground coupling. The inltration was set to 0.5 air changes
limitation should be kept in mind when conducting whole-building
per hour. The building mechanical system is 100% efcient with
simulations using EnergyPlus, such as those presented in the next
100% convective air, no duct losses, and no capacity limitation. The
section.
thermostat is set with dead band so heating takes place for
temperatures below 20  C and cooling for temperatures above
4. Annual whole-building performance evaluation of veried 27  C. Insulation properties have been slightly modied to incor-
PCM model porate PCMs (Table 6 and 7). In contrast with the PCMs modeled in
Sections 3.1e3.3, all PCMs in Table 7 are assumed to have a 2  C
After verication, the ideal step is to assess the performance of melting temperature range.
the PCM model with a simple and a more complicated building and The modied building was simulated using Phoenix TMY3
mechanical system. The model was evaluated in Phoenix using weather conditions with the three PCM applications described in
a lightweight building described in case 600 of ASHRAE Standard Table 7. The same building was also analyzed without PCMs using
140 [31] as well as a more realistic house using BEopt [40] with CondFD, and then using conduction transfer functions under
EnergyPlus as the underlying simulation engine. The main purpose different time steps and smaller node spacing for comparative and
of this initial evaluation was to test and demonstrate the simulation time-dependency analysis. Fig. 11 shows the run time for all
capabilities of the PCM model in the context of whole buildings, not simulated cases using EnergyPlus v6 (Ev6), v7.1 (Ev7.1), v7.1
to draw conclusions about the performance of specic PCM tech- with a smaller node space (Ev7.1 dx/3), v7.1 with a 2-min time
nologies. Future research will include detailed parametric analyses step (Ev7.1 2 min), and v7.1 with a 4-min time step (Ev7.1 4 min).
where different PCM melting temperatures and enthalpies are For each setting, the following simulations were performed: CTFs
analyzed. without PCM (CTF), CondFD without PCM (CondFD), CondFD with
PCM distributed in insulation (Insulation), CondFD with PCM
4.1. Annual performance model evaluation: modied ASHRAE between two insulation layers (Ins/PCM/Ins) and, CondFD with
Standard 140, Case 600 PCM distributed in drywall (Drywall). There are several important
aspects to highlight here:
The Case 600 model was selected because it is a simple, well-
referenced, and well-understood building that has been simu-
lated by several major simulation engines [31]. The basic test
building is a lightweight, rectangular, single-zone building, with
dimensions of 8 m wide by 6 m long by 2.7 m high, no interior
partitions, and a total window area of 12 m2 on the south wall. The

Table 6
Modied Case 600 materials properties.

Material Thermal Thickness Density Specic


conductivity (m) (kg/m3) heat (J/kg K)
(W/m K)
Wall Plasterboard 0.1600 0.0120 950 840
Insulation 0.03365 0.0660 29 960
Wood siding 0.1400 0.0090 530 900
Roof Plasterboard 0.1600 0.0100 950 840 Fig. 11. Run time for lightweight construction using EnergyPlus: v6 (Ev6), v7.1
Insulation 0.03365 0.1118 29 960 (Ev7.1), v7.1 with a smaller node space (Ev7.1 dx/3), v7.1 with a 2-min (Ev7.1
Roof deck 0.1400 0.0190 530 900 2 min), and a 4-min time step (Ev7.1 4 min). The y-axis represents: CTFs without PCM
Floor Timber ooring 0.1600 0.0100 950 840 (CTF), CondFD without PCM (CondFD), CondFD with PCM distributed in insulation
Insulation 0.0400 0.2500 12 840 (Insulation), CondFD with PCM between two insulation layers (Ins/PCM/Ins), and
CondFD with PCM distributed in drywall (Drywall).
P.C. Tabares-Velasco et al. / Building and Environment 54 (2012) 186e196 193

Table 8
Thermal storage of composite PCMs for BEopt new house.

Properties PCM distributed PCM distributed Concentrated


insulation drywall PCM layer
Thickness wall 0.089 m 0.0127 m 0.005 m
Walls latent storage 87 kJ/m2 420 kJ/m2 550 kJ/m2
Amount of PCM 0.5 kg/m2 3.8 kg/m2 4.2 kg/m2

speed is a key concern. Although not shown in the paper, the same
conclusions were found with an annual heating energy difference,
but with higher differences with respect to the case without PCMs.
Energy Difference was used instead of Energy Savings in
Fig. 12 because the authors main objective in using this simple
Fig. 12. Annual cooling energy difference for lightweight construction using captions building was to test the CondFD and PCM models, not to assess the
as Fig. 11. energy benets of the PCM applications. Predicted benets may be
smaller when more realistic homes with higher internal gains and
inltration loads are modeled. Results should not be extrapolated
 Overall, CondFD takes about 2e3 times longer to run than CTF from Fig. 12 to state which PCM application is best, as more detailed
when PCMs are not simulated, analysis is needed to address that problem.
 If PCMs are simulated, the run time doubles in comparison with
the case without PCMs and using CondFD, 4.2. Annual performance model evaluation: BEopt house
 The new automatic under-relaxation described in Section 2
reduces run time by a factor up to 4e5, and A second house was analyzed in EnergyPlus using the same
 Using a node space size three times smaller than the default weather data. This is a two-story, slab-on-grade, 231 m2 (2500 ft2)
size increases run time by a factor of 2e3. house with three bedrooms, two bathrooms, an unconditioned
attic, garage and a more realistic HVAC. This EnergyPlus model was
Fig. 12 shows the annual cooling energy difference, using CTF generated using BEopt (see Fig. 13) according to the Building
without PCM as the baseline. Key highlights include: America 2010 Benchmark [41]. Thus, this is a realistic example of
a newly constructed house. The 2010 Benchmark represents
 CondFD agrees with CTF within less than 0.2% when no PCM is a house built using 2009 International Energy Conservation Code
present, (IECC) and federal appliance standards in effect as of January 1, 2010
 Decreasing the node space size by a factor of 3 does not cause home. More details about the house can be found in the literature
signicant differences, [42].
 Increasing the time step to 2 min does not cause signicant A few modications were made to the original EnergyPlus
differences, and model generated by BEopt to accommodate PCMs:
 Increasing the time step to 4 min or longer causes signicant
differences, and should not be used.  The house was created in BEopt and the input EnergyPlus le
(idf) was exported to EnergyPlus where PCMs were introduced,
Overall, using the default space discretization and time steps up  The properties of the insulation used in the walls and attic were
to 2e3 min yields very similar results. This information can inform modied to the properties for PCM-distributed insulation
many parametric and optimization analyses where computational shown in Table 7, and

Fig. 13. BEopt rendering of example house simulated with various PCM applications.
194 P.C. Tabares-Velasco et al. / Building and Environment 54 (2012) 186e196

wall drywall (Drywall) and combination of RCRWall with PCM


distributed in Drywall (RCRWall-DW). The last PCM application was
chosen to evaluate the ability of the PCM model to simulate walls
with different PCMs in a single wall. The predicted annual electric
cooling consumption for the house without PCM was 5500 kWh,
less than half the 2009 average for homes in Arizona, which is
12,900 [43]. In all cases, the EnergyPlus run time was about 20 min
using a 2-min time step, about twice the run time of CTF. This
suggests that the CondFD and PCM models are working properly, as
previously shown with the modied ASHRAE Standard 140
building. Moreover, no instability problems were detected.
The predicted annual peak cooling electric load was 3.83 kW for
the house without PCM. Fig. 15 shows the predicted peak load
reduction for the PCM applications analyzed, which was about 4.3%
Fig. 14. Monthly cooling electric energy difference for Phoenix, Arizona: upper half
(Upper Attic) and lower half (Lower Attic) of the attic insulation with PCM, PCM for the best PCM analyzed in the cooling peak month of July.
concentrated at the middle of the attic insulation (RCRAttic), wall cavity with PCM Figs. 14 and 15 show very modest predicted energy benets.
distributed in insulation (Walls), PCM concentrated at the middle of the wall cavity However, predicted savings could be greater for a well-designed
insulation (RCRWall), PCM distributed drywall (Drywall) and combination of RCRWall PCM strategy that implements multiple technologies in a single
and PCM-Drywall (RCRWall-DW).
home and optimizes PCM properties and locations within the walls
or attic.
 The walls and attic were assumed to be studless and trussless.
This assumption was made to simplify the problem, as a rst 5. Conclusions
approximation as BEopt cannot model PCMs at this time.
The PCM model in EnergyPlus was veried and validated using
These assumptions resulted in a well-insulated building enve- an approach similar to ASHRAE Standard 140, which consists of
lope, as the equivalent R-values of the studless wall and trussless analytical verication, comparative testing, and empirical valida-
attic are RSI  2.7 (RIP  15.3) and RSI  6.7 (RIP  40), respectively. tion [31]. During this process, two bugs were identied and xed in
Thus, this case represents a home having walls with an R-value the PCM model. Version 8 of EnergyPlus will include the validated
close to a wall assembly consisting of RIP  19 batts, 5  15 cm PCM model and xes, which speed up the run time and allow for
(2  6 in) studs, 60 cm (24 in) apart from center having a RSI  2.9 simulation of PCMs with variable thermal conductivity.
(RIP  16.3). In contrast a typical wall consisting of RIP  13 batts, This study also identied a few key limitations of, and guidelines
5  10 cm (2  4 in) studs, 40 cm (16 in) apart from center have for using, the EnergyPlus PCM model:
a RSI  2 (RIP  11.4).
Table 8 shows storage capacity for the PCM assemblies in the  Time steps equal to or shorter than three minutes should be
walls using PCMs composite properties shown in Table 7. As with used,
the ASHRAE modied Case 600, the PCM distributed in insulation  Accuracy issues can arise when modeling PCMs with strong
has the lowest storage capacity due to the low density of the hysteresis,
insulation. Higher latent storage values could be achieved by  Default CondFD can be used with acceptable monthly and
increasing the percentage of PCM, the thickness of insulation layer annual results. However, if accurate hourly performance and
or the density of insulation. analysis is required, smaller node space (1/3 of the default
Fig. 14 shows the predicted monthly cooling electric energy value in EnergyPlus) should be used at the expense of longer
reduction for the following PCM applications: upper half of the attic run times.
insulation enhanced with PCM (Upper Attic), lower half of the attic
insulation enhanced with PCM (Lower Attic), PCM concentrated at The modied Case 600 with PCM proposed in this study could
the middle of the attic insulation (RCRAttic), wall cavity with PCM also be added into ASHRAE Standard 140 now that PCMs are
distributed in insulation, PCM concentrated at the middle of the becoming more common and other building energy simulation
wall cavity insulation (RCRWall), PCM distributed in the ceiling and programs have PCM models that might not have been validated.
Finally, this research opens multiple dimensions for future
research on PCMs, including the ability to condently investigate
different PCM material properties, congurations, and locations
within a house. Future research will include more detailed para-
metric analyses where PCM melting temperatures and enthalpies
will be analyzed for specic U.S. climate zones. Finally, integration
of the PCM model into BEopt would allow for the optimization of
PCMs using BEopts proven optimization schemes for new home
construction and existing home retrots.

Acknowledgements

This work was supported by the U.S. Department of Energy


under Contract No. DE-AC36-08-GO28308 with the National
Renewable Energy Laboratory. The authors would like to thank Ben
Fig. 15. Monthly peak cooling electric difference for Phoenix, Arizona using same Polly and Dane Christensen of NREL for their critical reviews and
captions as in Fig. 14. Brent Grifth of NREL for his technical support.
P.C. Tabares-Velasco et al. / Building and Environment 54 (2012) 186e196 195

Appendix [10] Kedl RJ. Conventional wallboard with latent heat storage for passive solar
applications; 1990. Medium: X; Size: Pages: (5 p.).
[11] Ahmad M, Bontemps A, Salle H, Quenard D. Experimental investigation and
PCM properties used in Experimental Validation computer simulation of thermal behaviour of wallboards containing a phase
change material. Energ Build 2006;38(4):357e66.
Thickness: 0.0053 m [12] Kuznik F, Virgone J, Roux J-J. Energetic efciency of room wall containing PCM
wallboard: a full-scale experimental investigation. Energ Build 2008;40(2):
Thermal conductivity: Solid (T < 21.6  C): 0.18 W/m K 148e56.
Liquid (T > 21.6  C): 0.14 W/m K [13] Kuznik F, Virgone J. Experimental investigation of wallboard containing phase
Density: 855 kg/m3 change material: data for validation of numerical modeling. Energ Build 2009;
41(5):561e70.
Specic heat: 2500 J/kg-K [14] Petrie TW, Childs KW, Childs PW, Christian JE, Shramo DJ. Thermal behavior of
mixtures of perlite and phase change material in simulated climate. Quebec
Enthalpy vs temperature data for DuPont Energain PCM used in City, USA: ASTM; 1997.
[15] Halford CK, Boehm RF. Modeling of phase change material peak load shifting.
Experimental Validation. Data obtained from differential scanning calo- Energ Build 2007;39(3):298e305.
rimeter (DSC) measurements with a heating rate of 0.05  C/min [39]. [16] Virgone, J, Nol, J, Reisdorf, R. Numerical study of the inuence of the thick-
ness and melting point on the effectiveness of phase change materials:
application to the renovation of a low inertia school, In: Eleventh interna-
tional IBPSA conference; 2009.
T ( C) H (J/kg  C) [17] Pedersen, CO. Advanced zone simulation in EnergyPlus: incorporation of
variable properties and phase change material (PCM) capability, In: Building
9.0 0.001 simulation 2007; 2007. Beijing, China.
7.0 5200 [18] Heim D, Clarke JA. Numerical modelling and thermal simulation of PCM-
5.0 10,800 gypsum composites with ESP-r. Energ Build 2004;36(8):795e805.
4.0 13,750 [19] Heim D. Isothermal storage of solar energy in building construction. Renew
3.0 16,850 Energy 2010;35(4):788e96.
2.0 20,350 [20] Schossig P, Henning HM, Gschwander S, Haussmann T. Micro-encapsulated
1.0 24,750 phase-change materials integrated into construction materials. Solar Energy
0.2 30,030 Mat Solar Cells 2005;89(2e3):297e306.
0.0 31,610 [21] Stovall TK, Tomlinson JJ. What are the potential benets of including latent
1.0 37,160 storage in common wallboard? J Solar Energy Eng, Trans ASME 1995;117(4):
2.0 40,510 318e25.
[22] Ibez M, et al. An approach to the simulation of PCMs in building applica-
2.5 42,160
tions using TRNSYS. Appl Therm Eng 2005;25(11e12):1796e807.
4.0 47,335
[23] Koschenz M, Lehmann B. Development of a thermally activated ceiling panel
5.0 50,885
with PCM for application in lightweight and retrotted buildings. Energ Build
7.5 60,135 2004;36(6):567e78.
10.0 70,010 [24] Tomlinson JJ, Heberle DD. Analysis of wallboard containing a phase change
12.5 81,010 material. Reno, NV, USA: IEEE; 1990.
15.0 93,760 [25] Kuznik F, Virgone J, Johannes K. Development and validation of a new TRNSYS
17.5 109,385 type for the simulation of external building walls containing PCM. Energ Build
20.0 129,635 2010;42(7):1004e9.
22.5 157,385 [26] Tabares-Velasco PC, Grifth B. Diagnostic test cases for verifying surface heat
23.5 170,985 transfer algorithms and boundary conditions in building energy simulation
24.0 177,535 programs. J Building Performance Simul; 2011:1e18. doi:10.1080/
25.0 186,185 19401493.2011.595501.
[27] Kosny J, Yarbrough D, Wilkes K, Leuthold D, Syad A. PCM-enhanced cellulose
26.0 191,185
insulation e Thermal mass in lightweight natural bers. In: 2006 ECOSTOCK
27.0 195,535
conference. Richard Stockton College of New Jersey: IEA, DOE; 2006.
28.0 199,485
[28] Lin K, Zhang Y, Xu X, Di H, Yang R, Qin P. Modeling and simulation of under-
29.0 203,135 oor electric heating system with shape-stabilized PCM plates. Building
30.0 206,335 Environ 2004;39(12):1427e34.
31.5 210,535 [29] Zhang YP, Lin KP, Yang R, Di HF, Jiang Y. Preparation, thermal performance
45.0 244,960 and application of shape-stabilized PCM in energy efcient buildings. Energ
80.0 332,460 Build 2006;38(10):1262e9.
[30] Huang MJ, Eames PC, Norton B. Thermal regulation of building-integrated
photovoltaics using phase change materials. Int J Heat Mass Transfer 2004;
References 47(12e13):2715e33.
[31] ASHRAE. ANSI/ASHRAE Standard 140-2004. In: Standard method of test for
[1] Khudhair AM, Farid MM. A review on energy conservation in building appli- the evaluation of building energy analysis computer programs. Atlanta, GA:
cations with thermal storage by latent heat using phase change materials. ASHRAE; 2004. p. 151.
Energy Convers Manag 2004;45(2):263e75. [32] Carslaw HS, Jaeger JC. Conduction of heat in solids. 2nd ed. Oxford University
[2] Tyagi VV, Buddhi D. PCM thermal storage in buildings: a state of art. Renew Press; 1959.
Sustainable Energy Rev 2007;11(6):1146e66. [33] Childs KW. Heating 7.2 users manual. Oak Ridge National Laboratory; 2005.
[3] Pasupathy A, Velraj R, Seeniraj RV. Phase change material-based building p. 242.
architecture for thermal management in residential and commercial estab- [34] Haavi T, Gustavsen A, Cao S, Uvslkk S and Jelle BP. Numerical simulations of
lishments. Renew Sustainable Energy Rev 2008;12(1):39e64. a well-insulated wall assembly with integrated phase change material panels
[4] Sharma A, Tyagi VV, Chen CR, Buddhi, D. Review on thermal energy storage with eComparison with hot box experiments, In: The international conference on
phase change materials and applications. Renew Sustainable Energy Rev 2009; sustainable systems and the environment; 2011. Sharjah, Sharjah, United
13(2):318e45. Arab Emirates.
[5] Wang X, Zhang YP, Xiao W, Zeng RL, Zhang QL, Di HF. Review on thermal [35] Hensen JLM, Nakhi A. Fourier and Biot numbers and the accuracy of
performance of phase change energy storage building envelope. Chinese Sci conduction modelling. In: BEP 94 conference facing the future.
Bull 2009;54(6):920e8. York: Building Environmental Performance Analysis Club (BEPAC);
[6] Zhu N, Ma Z, Wang S. Dynamic characteristics and energy performance of 1994.
buildings using phase change materials: a review. Energy Convers Manag [36] Waters JR, Wright AJ. Criteria for the distribution of nodes in multilayer
2009;50(12):3169e81. walls in nite-difference thermal modelling. Building Environ 1985;20(3):
[7] Drake JB. A study of the optima; transition temperature of PCM wallboard for 151e62.
solar energy storage. Oak Ridge National Laboratory; 1987. [37] EnergyPlus. EnergyPlus engineering reference: the reference to EnergyPlus
[8] Solomon AD. Design criteria in PCM wall thermal storage. Energy 1979;4(4): calculations. Ernest Orlando Lawrence Berkeley National Laboratory; 2011.
701e9. p. 1130.
[9] Athienitis AK, Liu C, Hawes D, Banu D, Feldman D. Investigation of the thermal [38] Kosny J, Christian JE. Thermal evaluation of several congurations of insu-
performance of a passive solar test-room with wall latent heat storage. lation and structural materials for some metal stud walls. Energ Build 1995;
Building Environ 1997;32(5):405e10. 22(2):157e63.
196 P.C. Tabares-Velasco et al. / Building and Environment 54 (2012) 186e196

[39] Cao S, et al. The effect of wall-integrated phase change material panels on the [41] Hendron R, Engebrecht C. Building America house simulation protocols.
indoor air and wall temperature e Hot box experiments, In: Zero emission Golden, CO: National Renewable Energy Laboratory; 2010. p. 80.
buildings e Proceedings of renewable energy conference 2010; 2010: [42] Casey S, Booten C. Energy savings measure packages: new homes. Golden, CO:
Trondheim, Norway. p. 15e26. National Renewable Energy Laboratory; 2011. p. 205.
[40] Christensen C, Horowitz S. Building energy optimization. Available from: [43] EIA. 2009 residential energy consumption survey [cited 2011, 10/21/2011];
http://beopt.nrel.gov/; 2011. Available from: http://205.254.135.24/consumption/residential/index.cfm; 2011.

También podría gustarte