You are on page 1of 12

Solar Energy 189 (2019) 45–56

Contents lists available at ScienceDirect

Solar Energy
journal homepage: www.elsevier.com/locate/solener

Numerical simulation and experimental validation of an outdoor-swimming- T


pool solar heating system in warm climates
S. Lugoa, L.I. Moralesb, R. Bestc, V.H. Gómezc, O. García-Valladaresc,

a
Posgrado en Ingeniería UNAM, Privada Xochicalco S/N, Temixco, Morelos C.P. 62580, Mexico
b
Facultad de Ciencias Química de Ingeniería, Universidad Autónoma del Estado de Morelos, Av. Universidad 1001, Col. Chamilpa, Cuernavaca, Morelos C.P. 62209,
Mexico1
c
Instituto de Energías Renovables-UNAM, Privada Xochicalco S/N, Temixco, Morelos C.P. 62580, Mexico

ARTICLE INFO ABSTRACT

Keywords: This paper presents a mathematical model developed in TRNSYS to simulate the performance of a solar heating
Solar energy system for an outdoor swimming pool in regions with a warm climate. For this purpose, a new type for TRNSYS
Outdoor swimming pool was developed. The model was validated using experimental data collected from an outdoor 53.8 m3 swimming
Experimental validation pool in Cuernavaca, Morelos, Mexico. The data used to confirm the model’s components and the full model were
TRNSYS
gathered from March 2016 to June 2017. This pool is located in a hotel surrounded by trees and vegetation that
Numerical model
Solar thermal systems
cause shading on the pool’s surface throughout the day, thus reducing the heat gain from direct solar radiation. A
shading factor equation for the shading over the pool was developed, introduced, and validated in the model to
consider variations in pool temperature. For evaluating the evaporative losses, six empirical correlations ob-
tained from the literature were tested. The model margin error was estimated at less than ± 2% (an average
of ± 0.41%) with a temperature differential of less than ± 0.5 °C (an average of ± 0.12 °C, root of the mean
quadratic error (RMSE) = 0.148 °C, mean bias error (MBE) = −0.058 °C, and coefficient of determination
(R2) = 0.9723) between the measured and simulated pool temperatures. Therefore, the model adequately re-
produced the pool’s temperature under different working conditions, and can be a valuable tool for generating a
technical and economic analysis of solar heating systems in outdoor pools for regions with similar climatic
conditions.

1. Introduction adequate solar radiation, and can achieve an expected investment re-
turn in less than 2 years in warm climates. SWHS can also be combined
Pool heating has been an issue for several centuries; Roman records with other technologies such as heat pumps and photovoltaic/thermal
indicate the use of solar water heating in public baths since 200 B.C. (PV/T) systems (Chow et al., 2012; Katsaprakaki, 2015) to maintain the
There is a constant requirement for comfortable temperatures in re- pool temperature at desired values, even in colder months or under
sidential and public pools, and amongst the most frequently used other conditions.
technologies in pool heating are water heaters that use fossil fuels such Since 1980, several theoretical models for outdoor swimming pools
as diesel, liquefied petroleum (LP) gas, and natural gas. However, the have been published (Govind and Sodha, 1983), and the most common
combustion of such fuels generates greenhouse emissions, which are a programs used are MATLAB’s Simulink and TRNSYS. The purpose of
critical factor affecting climate change. In addition, the cost of fossil these models is to predict the pool temperature under various operating
fuels have increased considerably in recent years. These factors have and meteorological conditions and obtain an economic analysis of the
created an opportunity for the use of renewable energy technology, return time of investment.
which aligns with the requirement of reducing costs, using energy more Govaer and Zarmi developed an analytical model to determine the
efficiently, and reducing CO2 emissions into the atmosphere (Govaer long-term thermal performance of an outdoor pool, and their results
and Zarmi, 1981; Singh et al., 1989; Molineaux et al., 1994; Barbato show that the model is applicable to indoor pools after making an
et al., 2018; Li et al., 2018). Solar water heating systems (SWHS) have adequate modification to the parameters (Govaer and Zarmi, 1981).
been proven to work very successfully for pool heating in places with Govind and Sodha presented an analysis of the heat transfer


Corresponding author.
E-mail address: ogv@ier.unam.mx (O. García-Valladares).
1
Posdoctoral stay at Instituto de Energías Renovables-UNAM.

https://doi.org/10.1016/j.solener.2019.07.041
Received 19 February 2019; Received in revised form 8 July 2019; Accepted 11 July 2019
Available online 18 July 2019
0038-092X/ © 2019 International Solar Energy Society. Published by Elsevier Ltd. All rights reserved.
S. Lugo, et al. Solar Energy 189 (2019) 45–56

Nomenclature emissivity (dimensionless)


humidity (%)
A surface area (m2) Stefan–Boltzmann constant (5.67 × 10 8W/m2K4 )
Ac solar collector array area (m2)
c constant from function f (t ) (h) Subscripts
COPr coefficient of performance ratio
Cp specific heat at constant pressure (J/kg K) a ambient
f factor (dimensionless) aux auxiliary
SF solar fraction (%) ave average
h enthalpy (kJ/kg) col solar collector
h̄ average convective heat transfer coefficient (W/m2 K) conv convective
H monthly average solar irradiation (MJ/m2 day) cond conduction
I solar irradiance (W/m2) exp experimental
Ic solar irradiance at collector plane (W/m2) eva evaporation
k thermal conductivity (W/m K) gain gain
L length (m) in inlet
m mass flow (kg/s) losses losses
N number of data points max maximum
P pressure (Pa) rad radiation
q conduction speed, (dimensionless) refill replacement water
Q heat flow or power (W) s shade
Q heat (J) sat saturation
T temperature (°C or K) sim simulated
t time (s) sol solar
V pool volume (m3) sky sky
X variable data soil ground
w wind speed (m/s) p pool
pump pump
Greek letters out outlet
v steam
pool water absorbance (dimensionless) w water
density (kg/m3)

processes in the solar heating of a pool in two analytical models, one for value of ± 0.6 °C, and includes calculations for conductive heat losses.
an outdoor pool and another for a pool with a polyvinyl chloride (PVC) It takes into consideration that the total conductive losses can be
cover. The solar insolation and atmospheric air temperature are as- greater than 1% of the total losses, because the soil temperature is not
sumed to be periodic. The experimental water temperature in a pool in constant. The pool was used as a heat sink for air conditioners (Woolley
Victoria, Australia, showed close agreement with these theoretical et al., 2011).
calculations (Govind and Sodha, 1983). Cunio and Sproul investigated the theoretical and experimental
Hahne and Kluber’s model emphasises that losses due to heat con- performance of solar collectors for outdoor pool heating without
duction are equivalent to 1% of the total heat loss from pools, and they glazing and at low flow rates. They observed that the performance was
can thus be neglected. In contrast, evaporative heat losses can account heavily penalised until the flow drops below 50 l/min. Therefore, the
for up to 60% of the total pool heat loss. Furthermore, this model also efficiency for 60 l/min is approximately 15% of that obtained for 140 l/
takes into consideration that not all the solar radiation is absorbed by min. A simple cost analysis also showed that the energy savings result in
the pool surface. Therefore, it includes factors that represent the solar significant cost savings over a lifetime of 10 years (Cunio and Sproul,
radiation fraction absorbed by the pool surface as a function of the 2012).
incidence angle. However, this model has only been validated for the Zsembinszki et al. simulated the thermal behaviour of water in the
German climate with solar fractions below 50% (Hahne and Klüber’s, presence and absence of a PCM (phase-change material) storage system
1994). and studied the effects of PCM under comfort conditions. Two methods
Ahmad and Khane performed a gains and losses analysis for an of PCM used for heating are presented: the first comprised the use of the
outdoor pool. The results of their thermal analysis indicates that a heat in the sidewalls and bottom of the pool, and the second comprised
comfortable temperature can be maintained in the pool if a plastic the use of an external heat exchanger with a PCM. The results showed at
cover is used during the winter season (Ahmad and Khan, 2009). difference between the temperature values of the registered and simu-
Ruiz and Martínez used different correlations for the evaporative lated pool of less than 0.2 °C, and a deviation of 0.009. The im-
coefficient one of which was developed by Richter. Richter's correlation plementation of a PCM storage system in an outdoor pool can produce
had the lowest standard deviation value (0.036) for evaporative heat an increase in the water temperature of up to 2 °C even for a limited
losses as compared to other correlations. However, this model does not number of days which occurs during periods of adverse weather con-
take into consideration the losses due to water replacement and con- ditions such as low sunlight, low air temperatures, or high wind speeds.
ductive heat losses, and the experimental validation was only per- In contrast, when the weather is warm, and the pool water is suffi-
formed for three consecutive days. Further, the pool temperature was ciently warm such that it reaches a comfortable value, the PCMs have
the only parameter compared (Ruiz and Martínez, 2010). the effect of reducing the pool temperature by absorbing excess heat
Woolley et al. evaluated the effect of solar radiation on an outdoor (Zsembinszki et al., 2012).
pool without a solar heating system. This model demonstrated that the Zayed et al. presented that a significant improvement in energy of
equations used in previous models can be adjusted to obtain a deviation flate plate solar collectors can be obtained using carbon based

46
S. Lugo, et al. Solar Energy 189 (2019) 45–56

nanofluids compared to metal oxides nanofluids under the same con- without shading. It was found in the literature review that there are no
ditions (Zayed et al., 2019a) and they also presented a review of the models that take into consideration shading factors, which are com-
novel and most recent developments of PCM and cascade thermal sto- monly present in actual cases, and the majority of models have been
rage (multiple PCM in descending order of their melting temperatures) developed and validated for indoor and outdoor pools in cold climates
and their implementation in solar water collectors storage tanks (Zayed with low solar fractions and high return of investment periods. Few
et al., 2019b). works have been developed for countries with warm climates wherein
Santos et al. proposed a hybrid structure for simulating the ther- outdoor pools are very common, and they are used throughout the year.
modynamic behaviour of pools using neural computational models to For this reason, in these locations the return on investment is very in-
incorporate the climatic information of regions in Brazil. Although, the teresting to use solar technology in a massive way.
neural representation takes into consideration the latitude, longitude, In this work, a mathematical model was developed using TRNSYS,
and elevation, locations that are geographically distinct from the which has been validated using data collected from an outdoor pool
training data set cannot be represented, thus making the simulation located in a hotel in the city of Cuernavaca, Morelos, Mexico. This city
structure inconsistent (Santos et al., 2013). has a yearly solar radiation average of 5.3 kW h/m2day (19 MJ/m2day)
Buonomano et al. developed a model for two different cases: indoor with an average ambient temperature of 21.4 °C, data collected from
and outdoor pools heated using a PV/T system in which the electricity the National Meteorological Service, and in the months of April and
and water heating are produced simultaneously and a heat exchanger is May, the average maximum temperature can reach 28 °C. Owing to
used to heat the pool. This model makes use of seven correlations for optimal meteorological conditions, the use of solar energy systems has
estimating the evaporative coefficient and three for the convective been increasing in Mexico for the last few years.
coefficient. Furthermore, in addition to testing ASHRAE’s equation for The technology used in this hotel’s pool is unglazed, flat plate col-
total heat loss, it also takes into consideration the heat loss due to water lectors fabricated from high-density polypropylene with ultra-violet
replacement and conduction. An economic analysis was also performed. protection treatment. The pool’s temperatures and the inlet and outlet
However, this model has not been validated experimentally water temperature for the solar array were measured without the use of
(Buonomano et al., 2015). a thermal cover. As mentioned above, an additional factor was included
Kaci et al. developed a dynamic model for an outdoor pool heating to take into consideration the shading on the pool generated by the
while comparing two types of solar collector arrays: one with and an- dense trees and vegetation present at the hotel. The model used was
other without a glass cover. The results showed that the collector generated with TRNSYS version 16 (TRNSYS, 2005) and validated with
without a glass cover had an annual solar fraction of approximately data collected from March 2016 to June 2017.
47% against the collector with a cover with 27%. The economic as- The main contributions of this work are the development of a new
sessment indicates that the first system is profitable with a return period type (pool) of TRNSYS model in order to calculate the pool tempera-
of 9 years (Kaci et al., 2017). ture; a shading factor is included in this model, which has not been
Zuccari et al. calculated through an ad_hoc developed algorithm reported in other models of solar pool heating analysis and is useful for
(named EnerPool) the potential savings in terms of non-renewable pools with similar conditions and for the experimental validation of the
primary energy consumption that are achievable through energy effi- model. Further, a validation for each component of the system (and for
ciency actions involving heating, filtration, and water replacement for the system as a whole) is performed, which has also not been reported
indoor swimming pools in Italy. This work analyses in detail some in prior published models. Finally, with respect to the experimental
possible solutions for reducing the heating requirement, while facil- validation of the model, the validation for each component of the
itating high non-renewable primary energy savings (up to more than system including the solar collector array as well as the system as a
50%) at a low cost and with a payback time of less than 2 years by whole is performed and compared against the measured pool tem-
means of renewable sources (solar collectors, photovoltaic panels, and perature, which has also not been presented in previously published
pool covers), thus achieving economic and energy savings but with models. This facilitates the evaluation of the uncertainty of the solar
much higher initial costs. They found that 60% of the losses are due to collector array model, which is not restricted to the swimming pool
evaporation in this type of system. The case of the outdoor swimming model. The validated model can be used as a good tool for determining
pools has not been analysed or experimental validated (Zuccari et al., the thermal performance of outdoor pools in warm climates throughout
2017). the year.
Foncubierta et al. presented an experimental procedure at the la-
boratory scale that was developed to validate a new computational- 2. System description
fluid-dynamic-based (CFD-based) methodology for the estimation of the
water evaporative rate in indoor swimming pools. The comparison The thermal analysis was conducted on an outdoor swimming pool
between the simulated and experimental results shows that the mod- at “Las Quintas” hotel in the city of Cuernavaca, Morelos, Mexico, lo-
elling strategy proposed is a promising tool, with average relative errors cated at 18.9192°N 99.2181°W and an altitude of 1510 m above sea
of 9% for the typically mixed convection flows in indoor swimming level. The pool surface area is 44.4 m2, and it has a volume of 53.8 m3.
pools (Foncubierta et al., 2018). Owing to its location, this pool is partially shaded throughout the day.
With the information supplied by theoretical models, it is possible to Buildings and vegetation generate this shading around the pool, which
determine the energy savings generated by the solar system during its results in a decrease in the energy gained by the pool from direct solar
operation. According to different authors, the important variable in radiation (Fig. 1).
these models are the evaporative heat losses as these have a great im- This factor is taken into consideration in the proposed model be-
pact on the pool’s total energy balance. Therefore, some authors have cause it has a relevant effect on the temperature of the pool.
developed semi-empirical equations for estimating the evaporative The pool is heated by a solar thermal system comprised of a 2.18 kW
losses under various conditions such as in use, without use, indoor, and pump and 12 unglazed polypropilene flate plate solar collectors. The
outdoor pools (Smith et al., 1994; Shah, 2003, 2012). system configuration comprises two arrays connected in series, wherein
The location is another main factor affecting the pool temperature, the first array consists of five collectors in parallel, and the second
considering factors such as microclimate, constructions, and shadings. consists of seven collectors in parallel. The total collector area is
For example, a pool located in an open space will be affected by eva- 45.6 m2 and is located on top of one of the buildings close to the pool, as
porative losses due to air currents that have an impact on the heat shown in Fig. 2. The solar heating system has an angle of 35° SE and an
transfer and greater radiative losses to the sky; in contrast, a shaded inclination of 11.5°. It is considered that the surface of the installed
pool will have a lower heat gain from direct solar radiation than a pool collector field (45.6 m2) is similar to the pool surface (44.4 m2) as per a

47
S. Lugo, et al. Solar Energy 189 (2019) 45–56

direction and speed, ambient temperature, humidity, and precipitation;


the characteristics of all the data acquisition devices are included in
Table 1.
The solar heating system pump is controlled (switch on and switch
off of the pump) by a differential temperature control between the
outlet section of the collector array and pool temperature; it switches on
with a temperature difference of 4 °C and switches off when this tem-
perature difference falls below 2 °C. The pump also switches off when
the pool temperature is greater than 31 °C.
The solar system and monitoring diagram are presented in Fig. 3.
This system has been in operation since July 31st, 2015, and has pro-
duced 194393 MJ and mitigated 18.29 Ton of CO2, which translates
into savings of 6250 kg of LP gas savings of 4254 USD as reported on
April 2nd, 2019 (Las Quintas, 2019).
Table 2 presents technical information and the characteristics of the
solar collectors used in the system. The efficiency curve for the solar
collector clearly shows how the efficiency is affected by wind speed (w )
in the case of an unglazed flat plate collector.
Fig. 1. Aerial view of the pool.

3. Mathematical model in TRNSYS

3.1. Collectors

The mathematical model for solar collectors was developed using


the dynamic simulation software TRNSYS version 16. The mathematical
model for solar collectors corresponds to Type 1 for a flat plate solar
collector with a quadratic efficiency curve.
On applying the thermal efficiency equation and using the average
water temperature in the collector and climate data (wind speed, solar
irradiance, and ambient temperature), it is possible to obtain the effi-
ciency value for a given instant within a time interval and, therefore,
obtain the collector-array useful heat and the outlet temperature for
that time interval.
Fig. 2. Solar collector installation. Efficiency values obtained using the ISO 9806:2013 standard
(Table 2) were modified using correction factors that apply when: (a)
rule of thumb practice used by some installation technicians that as- the mass flow used differs from the mass flow reported during testing
sumes that the thermal losses in the pool are equivalent to the energy for the efficiency curve; (b) more than one collector is connected in
gain by the solar collector array for most of the year in warm climates. series; and (c) the solar incidence is not perpendicular to the collector.
The solar installation of this pool was not sized by the authors of this
work. This system was equipped with a mass flow rate sensor and 3.2. Pool
several temperature sensors, which send a signal to a monitoring system
that can be viewed in real time via a web page (Las Quintas, 2019). For The mathematical model for an outdoor pool comprises a balance of
obtaining the atmospheric data, a meteorological station was installed energy gains and losses, as shown in Eq. (1). In Table 3, various cor-
to measure the values of the variables such as solar radiation, wind relations for heat gains and losses reported by different authors are
presented.

Table 1
Data acquisition instrumentation.
Pool measuring instruments

Variable Sensor Characteristics

Temperature PT-1000 Accuracy: ± 0.3 °C


Measurement range: −50 °C-750 °C
Mass flow Magnetic flowmeter Accuracy: ± 0.5%
Measurement range: 0–830 kg/min

Weather station measuring instruments

Temperature and humidity Temperature and humidity transducer Temperature accuracy: ± 0.5 °C
Measurement range: −20-60 °C
Humidity accuracy: ± 0.3%
Measurement range: 0–100%
Solar irradiance Spectral pyranometer Measurement range: 0–2000 W/m2
Sensibility: 5-20 µV/W/m2
Wind speed Anemometer Accuracy: ± 5%
Measurement range: 0.5–89.0 m/s
Rainfall Rainfall sensor Accuracy: ± 4%
Measurement range: 0–999.8 mm/m2

48
S. Lugo, et al. Solar Energy 189 (2019) 45–56

Fig. 3. Solar system and monitoring schematic.

Table 2 A new type (Pool) of model for TRNSYS version 16 was developed in
Characteristics of solar collectors. order to calculate the pool temperature. The parameters and input and
Parameter Specification output data are presented in Table 4.

Collector type Unglazed flat plate 3.3. Pool solar heating system
Material Polypropylene
Area [m2] 3.65
Fluid capacity [litres] 11.54 The model of the complete heating system was developed in
Test conditions mass flow [kg/h m2] 288 TRNSYS; the solar collector and pool model were incorporated to op-
Optical efficiency [–] 0.9248–0.0512w erate in conjunction while simulating the performance of the entire
First order thermal losses coefficient [W/m2 K] 15 + 6.3w
system.
Fig. 4 shows the layout of the complete system for simulating the
complete performance of an outdoor pool with a solar heating system.
The pool energy balance is
In the modelling, Type 9 was included to input the experimental data
dTp Qgain Qlosses from the meteorological station obtained during the test time interval in
Vp w Cpw =
dt dt dt (1) order to calculate the theoretical pool temperatures from the model and
The heat gain is compare them to the experimental data.
Type 9 can be replaced with Type 109 for inputting the meteor-
Qgain = Qsol + Qcol + Qaux (2) ological data of the region for a full typical year in the TMY format in
The heat losses are order to perform a feasibility and profitability analysis of the installa-
tion for a complete operational year. The TMY file was generated using
Qlosses = Qeva + Qrad + Qconv + Qcond + Qrefill (3) the Meteonorm 7.0 software and the climate data for Cuernavaca,
In the pool’s energy balance Eq. (1), Vp represents the pool volume, Morelos, obtained from the UNAM database (ESOLMET-IER, 2018).
w is the pool water density, and Cpw is the specific heat at the constant
pressure of the pool's water, the heat gains due to direct solar radiation 3.4. Piping heat losses
on the pool’s surface are represented by Qsol , gains at the solar collector
array are Qcol , and auxiliary heat is denoted by Qaux , heat losses due to The piping heat loss modelling was based on Type 204 previously
evaporation are Qeva , radiative losses are Qrad with used by the authors for another case study (Lugo et al., 2019). Never-
Tsky = 0.0552(Ta + 273.15)1.5 273.15(oC), convective losses are Qconv , theless, the piping heat losses were considered irrelevant owing to the
conductive losses are Qcond , and losses due to water replacement are flow rate (100–180 l/min), piping distance (10 m), piping diameter
Qrefill . (0.0381 m), piping material (PVC), and operating conditions, which in
In this model, the conductive losses (Qcond ), replacement water this case amounted to losses lower than 0.1 °C.
(Qrefill ), and pump power were considered to be negligible (Buonomano
et al., 2015) given that their magnitudes were considerably small as 4. Experimental validation
compared to the other coefficients.
The pool model developed in this article is based on the equations The experimental data obtained from March 2016 to June 2017
presented by Ruiz and Martínez (2010), except for the evaporation were analysed to validate the model, and the representative days of this
correlation. The evaporative losses presented in Table 3 were con- period are presented in this section. To validate the complete model, it
sidered in order to determine the one that best fit the experimental was essential to validate each component separately against the ex-
data. perimental data. Further, while considering the necessary input data for

49
S. Lugo, et al. Solar Energy 189 (2019) 45–56

each component model, the solar collector model and outdoor pool
model were validated separately and then simulated in a coupled
1 outdoor and 1 indoor pool heated with PV/T solar configuration, to obtain the validation for the complete system.

Qlosses = Qrad + Qeva + Qconv + Qcond + Qrefill 4.1. Solar collector model

To validate the solar collector model, the following input data were

Pv, a ] dt
QHE1, QHE 2 are heat exchangers used: solar radiation on a horizontal plane, ambient temperature, wind
Qlosses speed, inlet temperature to collectors, mass flow rate, area, and col-

qksoil Ap (Tp Tsoil )


Qgain = Qsol + QHE1 + QHE 2

dt

) dt
Ta) dt
Buonomano et al. (2015)

lector array. The values compared were the outlet temperature from the

Tw )
Ta )

Ap w (T p4 Tsky
Qeva = Ap h eva [Pv, sat (Tp)

4
h eva = 0.036 + 0.025w
solar collector array as well as the energy gain by the pool (experi-
Q gain

2.8 + 3.0w
3.1 + 4.1w
dt

Qrefill = mw Cpw (Tp


Qconv = hconv Ap (Tp
Qlosses = 0.06Ap (Tp

mental and simulated), and the instantaneous thermal efficiency of the

No solar heating
=
hybrid system

Does not apply


solar collector array was calculated using the following equation.

Qsol = Ap Idt
dTp
w Cpw dt

2 Lc
1
{ Qsolar mCp (Tout Tin)

Qcond =
=

Qrad =
= =
h conv
Ic Ac + Qpump Ic Ac + Qpump (4)
Vp

where Ic is the solar irradiance at solar collector plane (W/m2), Ac is the


Qlosses = Qrad + Qeva + Qconv + Qcond
100-m2 outdoor pool with solar heating Outdoor pool without solar heating

Pv, sat (Tp ) Pv, amb

area of the solar collector array (m2), Qsolar is the power transfer to the
Pv, a ] dt

Tp Ta

water for the solar collector array (W), and Qpump is the pump power
Qlosses

qksoil Ap (Tp Tsoil )


dt

) dt

(W). The coefficient of performance ratio (COPr) is evaluated using Eq.


Qrad = Ap w (T p4 Tsky
Qeva = Ap h eva [Pv, sat (Tp)

(4), while considering IcAc = 0.


4
Woolley et al. (2011)

h eva = 0.036 + 0.025w


Q gain

The margin of error for each simulated data point was obtained
dt

Qconv = Qeva Rbowen

Cbowen a
po
p

No solar heating

using Eq. (5).


=

Does not apply


Not considered
Qsol = Ap Idt
dTp
dt
Qgain = Qsol

2 Lc
1
w Cpw

Xsim Xexp
Rbowen =

Qcond =

%Error = 100
X exp (5)
Vp

The variable X can be the temperature (°C) or energy (J).


In the same manner, the root of the mean quadratic error was cal-
culated using Eq. (6).
Pv, a ] dt
Qlosses
Qlosses = Qrad + Qeva + Qconv

dt

N
) dt

Xexp, i )2
Ta ) dt

(Xsim, i
Ruiz and Martínez (2010)

h eva = 0.0423 + 0.0565w 0.5

Tp)

i=1
RMSE =
Tsky
Qeva = Ap h eva [Pv, sat (Tp)

(6)
4

N
Q gain

Qcol = mcol Cpw (Tcol


dt

Qconv = hconv Ap (Tp


Qgain = Qcol + Qsol

h conv = 3.1 + 4.1w

where N represents the number of data points.


(T p4
=

Does not apply


Not considered

The mean bias error (MBE) was calculated using the following
Qsol = Ap Idt
dTp

w
Not relevant
dt

equation:
Qrad = Ap
w Cpw

N
(Xsim, i Xexp, i )
Vp

i=1
MBE =
N (7)
2
The coefficient of determination R was calculated using Eq. (8):
Evaluation of eight outdoor pools with solar heating and

= Reflectedsolar radiation factor

N
i=1
(Xsim, i X exp, i ) 2
R2 = 1 N
i=1
( Xexp, i Xexp, ave ) 2 (8)
Qlosses = Qrad + Qeva + Qconv + Qcond + Q w

Not relevant, less than 1% of total heat loss

Fig. 5 shows a comparison between the experimental outlet tem-


peratures from the solar collector array and the simulated temperatures
as well as the instantaneous thermal efficiency of the solar collector
Pv, a ] dt

array and the solar irradiance measured at the collector plane. It can be
Qlosses
dt
Summary of pool energy balances reported in the literature.

) dt
Ta ) dt
Hahne and Klüber (1994)

Tp)

observed that there is a good correlation between the simulated and


Qgain = Qcol + Qsol + Qaux

Tp)
Tw )
Tsky
Qeva = Ap h eva [Pv, sat (Tp)

experimental data. The average deviation between the simulated and


Q gain

h eva = 5.058 + 6.69w

) Idt

Qaux = mw Cpw (Taux


Qcol = mcol Cpw (Tcol
dt

Qconv = hconv Ap (Tp

Qrefill = mw Cpw (Tp

experimental outlet solar collector temperature was ± 0.17 °C, and the
h conv = 3.1 + 4.1w
auxiliary heating

(T p4
=

error was ± 0.53% (RMSE = 0.194 °C, MBE = −0.12 °C, and
dTp

w
dt

Direct solar gain on the pool surface Qsol = Ap (1

R2 = 0.9805). With respect to the difference between the numerical


Qrad = Ap
w Cpw

and experimental data for the useful energy gain of the solar systems (in
this case between 31.37 and 8.53 kW), the average deviation was ±
Vp

5.67% (RMSE = 1.48 kW, MBE = −0.94 kW, and R2 = 0.9461). The
average mass flow rate was 6549 kg/h, the values of the instantaneous
thermal efficiency of the solar collector array Eq. (4) are between
52.9% and 75.6% with an average of 67.6%, and the coefficient of
performance ratios (COPr) are between 4.5 and 14.4 with an average of
10.6, with a solar irradiance at the collector plane between 331.6 W/m2
Auxiliary heat gain
Collector heat gain
System description

Evaporative losses

Conductive losses
Convective losses

and 866.5 W/m2 with an average of 696.9 W/m2.


Radiative losses
Energy balance

Refill losses
Reference

4.2. Outdoor pool model


Table 3

Climate variables. To validate the model, the experimental data

50
S. Lugo, et al. Solar Energy 189 (2019) 45–56

Table 4
Parameters of the new type of model for TRNSYS.
Parameters Input Output

Initial temperature of pool water Ambient temperature (Ta ) Pool temperature (Tp )
.
Pool surface area ( Ap ) Relative humidity ( ) Power gain from auxiliary heat (Qaux )
.
Pool volume (Vp ) Wind speed (w ) Power loss by evaporation (Qeva )
.
Height at which wind speed measurement is obtained Solar irradiance (I ) Power loss by convection (Qconv )
.
Mass flow rate supply by solar collectors (mcol ) Power loss by radiation (Qrad )
.
Temperature supply by solar collectors (Tcol ) Power gain by solar collectors (Qcol )
.
Power gain by direct solar radiation (Qsol )

was measured from March 2016 to June 2017. Corrections to this ex-
perimental data were made to take into consideration the meteor-
ological station being 7 m above the pool level, and it was necessary to
obtain the humidity, wind speed, and ambient temperature data at the
pool level for implementing the model.
Solar contribution. As mentioned in previous sections of this work,
a shading factor ( fs ) was considered in the model to adjust the direct
solar contribution (Qsol ). The pool is in a section of the hotel with
considerable shading from buildings and vegetation. The amount of
shading varies depending on the time of day and year; for example, in
spring and autumn, the shading begins at 12:00 h, and by 15:20 h, the
pool is completely shaded, while in summer and winter, the shading
starts at 13:00 h, and by 16:20 h, the pool is completely shaded.
The shading factor ( fs ) was introduced into the solar gain equation
as shown in the following equation.

Qsol = (1 fs ) Ap Idt (9)

The shading factor ( fs ) takes a value from 0 to 1 during the shaded Fig. 5. Solar irradiance, solar collector array efficiency, and comparison be-
period. However, the shade varies according to the season of the year as tween simulated and experimental data for the outlet solar collector tempera-
mentioned above. Therefore, the value of fs in Eq. (9) is estimated ac- ture for October 14th, 2016.
cording to Table 5.
The function f (t ) is given by Eq. (10). (10).
Fig. 7 shows the experimental pool temperature profile (T2) ac-
f (t ) = 0.3(t c) (10)
cording to Fig. 3, and the simulated pool temperatures using several
where t is the solar time in decimal form (for example, 13:20 would be semi-empirical equations for the evaporative coefficient in the pool for
13.333 h), and c is a constant equal to 12 h. April 25th, 2017. The error analysis between the experimental results
Fig. 6 shows the map of the pool shade obtained according to Eq. and the different empirical correlations used for the evaporation losses

Fig. 4. TRNSYS layout for simulating the outdoor pool performance with a solar heating system.

51
S. Lugo, et al. Solar Energy 189 (2019) 45–56

Table 5 Table 6
Values of fs according to the season of the year. Comparison of average errors of pool temperature for the different evaporative
losses equations against experimental data.
fs Spring–autumn Summer–winter
Evaporative losses Error (%) ΔT (°C) RMSE (°C) MBE (°C) R2
0 t<c t < (c + 1) equation
f(t) c ≤ t ≤ (c + 3.333) (c + 1) ≤ t ≤ (c + 4.333)
1 t > (c + 3.333) t > (c + 4.333) Richter (1979) 0.10 0.03 0.0436 −0.0018 0.9979
ISO, McMillan (1971), 0.20 0.07 0.0843 −0.0636 0.9920
ISO (1995)
Smith et al. (1994) 0.48 0.16 0.1868 −0.1608 0.9607
Rohwer (1931) 0.94 0.31 0.3573 −0.3147 0.8563
ASHRAE (2003) 1.02 0.34 0.3893 −0.3434 0.8294

The solar heating system was in operation from 10:00 h to 17:00 h, local
time. The performance for simulated temperatures and experimental
data can be observed with an average margin of error of ± 0.14%
(RMSE = 0.05 °C, MBE = −0.02 °C, and R2 = 0.9974), and a tem-
perature difference error of ± 0.04 °C. If the shading factor was not
taken into consideration in the numerical model (see Tsim_without
shadow in Fig. 8), the margin of error between the simulated and ex-
perimental data was ± 0.83% (RMSE = 0.31 °C, MBE = +0.21 °C, and
R2 = 0.9124) with a temperature difference of ± 0.26 °C. The margin of
error indicates that the model accurately reproduces the values of the
pool temperature and the importance of considering the pool shading
factor. It can also be confirmed that the model accurately simulates the
pool data for days with clear skies and high levels of solar irradiance.
The highest irradiance value for this specific day was 923.3 W/m2, with
an average irradiance of 732.5 W/m2 and an average ambient tem-
perature of 33.5 °C. The average and final value of pool temperature are
30.4 °C and 31.8 °C, the instantaneous thermal efficiency of the solar
Fig. 6. Map of pool shades in percentage.
collector array is 60.1% (COPr = 7.7) with a maximum value of 81.4%
(COPr = 13.0). If the shading factor was not taken into consideration in
are shown in order of priority in Table 6. The Richter’s correlation the numerical model, the average and final value of pool temperature
(Richter, 1979) provides the best adjustments for obtaining real values are 31.0 °C and 32.8 °C, the instantaneous thermal efficiency of the
with an error of ± 0.10%, RMSE = 0.0436 °C, and MBE = −0.0018 °C solar collector array is 57.6% (COPr = 7.4) with a maximum value of
with a temperature difference of ± 0.03 °C. This is followed by the 75.0% (COPr = 12.3). As shown in Fig. 8, there is a decline in the solar
correlation used by McMillan (1971) and ISO (1995) with an error irradiance after 15:00 h; nevertheless, this does not affect the behaviour
of ± 0.20%, and RMSE = 0.0843 °C with a temperature difference of the pool temperature because the pool is completely shaded by this
of ± 0.07 °C. After these are the correlations of Smith et al. (1994), time.
Rohwer (1931), and ASHRAE (2003). Fig. 9 shows a temperature comparison for April 25th, 2017; for this
Fig. 8 shows the results obtained for May 19th, 2017. For this day, date, a complete energy balance for the collector system was calculated
the same shading factor was used as that used in the month of April. along with the pool heat gains and heat losses. The margin of error
between the simulated and experimental data was ± 0.11%
(RMSE = 0.044 °C, MBE = +0.035 °C, and R2 = 0.9978) with a tem-
perature difference of ± 0.035 °C, an average instantaneous thermal

Fig. 8. Pool temperature and solar collector array efficiency comparison for the
Fig. 7. Pool temperature profiles obtained using various empirical correlations model with and without shading factor with solar irradiance for May 19th,
for the evaporative losses against experimental data. 2017.

52
S. Lugo, et al. Solar Energy 189 (2019) 45–56

reached during the day time.

4.3. Solar heating system for the outdoor pool

Figs. 9 and 10 show the results for the simulated pool temperatures
against the measured pool temperatures; with the previous results
presented, both the models were validated separately, and the pool
temperatures for both the linked models will be presented in the fol-
lowing figures.
Once the separate models were validated, the complete model was
assembled to validate its performance (solar collector model and pool
heat gains and losses model) to obtain a complete model that simulates
an outdoor pool that is heated using a solar thermal system.
Fig. 11 shows a comparison between the simulated temperatures
and experimental data for April 25th, 2017. There is a very close ap-
proximation between the model and the measured results, and the
margin of error is ± 0.24% (RMSE = 0.099 °C, MBE = −0.029 °C, and
Fig. 9. Pool temperature comparison, solar collector array efficiency, and solar
R2 = 0.9888), which is equivalent to a ± 0.08 °C temperature differ-
irradiance for April 25th, 2017. ential. The solar irradiance and efficiency of the solar collector array for
this day can be observed in Fig. 9.
Fig. 12 shows the comparison between the model simulation and
measured data for 47 h for 3 consecutive days, while considering the
initial pool temperature for day one and the climate data for each time
interval. An average deviation of ± 0.73% (RMSE = 0.246 °C,
MBE = −0.061 °C, and R2 = 0.9223) is observed between the simu-
lated results and measured data, which translates into a ± 0.20 °C
temperature difference. The solar irradiance and efficiency of the solar
collector array for these days can be observed in Fig. 10.
It is important to mention that several simulations for other days
during different seasons were performed (four in spring, one in
summer, four in autumn, and one in winter), and the average error for
these data was ± 0.41% (RMSE = 0.148 °C, MBE = −0.058 °C, and
R2 = 0.9723), which is equivalent to a temperature difference
of ± 0.12 °C.

5. Case study

After a complete evaluation of the pool simulation model and ver-


Fig. 10. Pool temperature comparison, solar collector array efficiency, and
ifying its reliability, the operation of the pool was analysed for an entire
solar irradiance for June 14th–16th, 2017.
year, for which the monthly climate data are presented in Table 7.
An analysis was performed for the two study cases: the first case did
efficiency of the solar collector array of 56.25% (COPr = 9.7) with a
maximum value of 85.82% (COPr = 15.8), an average reported irra-
diance of 733.1 W/m2 with a maximum and minimum value of
1014.3 W/m2 and 429.3 W/m2, respectively, and an average ambient
temperature of 31.1 °C. This confirms a good performance of the pool
model as compared with the experimental data.
Once the model was adjusted to simulate the pool temperature
during the day, simulations were run for 47 h while taking into con-
sideration nocturnal heat losses, and the dates used were June 14th to
16th, 2017. For these dates and the shading factor used in the model,
the shading started at 13:00 h because summer begins on June 20th.
Fig. 10 shows the results for this simulation: the margin of error for the
pool temperature was ± 0.65% (RMSE = 0.214 °C, MBE = −0.129 °C,
and R2 = 0.9411) with a temperature difference of ± 0.18 °C. In this
figure, the thermal efficiency of the solar collector array is obtained for
June 15th, the solar irradiance is obtained for the partially cloudy days
of June 14th and 15th, and the pump is switched off on June 14th. A
decline in the temperature can be observed to start at 15:20 h because
of the pool shading factor, and the maximum pool temperature mea-
sured was 28.6 °C, which is within the acceptable pool temperatures for
the hotel. Owing to cloudy skies, the temperatures did not reach as high
as those in April or May (maximum temperature measured was 34.1 °C),
and it can also be observed that the nocturnal temperature losses were
approximately 2 °C in June with reference to the highest temperature Fig. 11. Comparison between pool temperature of complete model simulation
and measured data for April 25th, 2017.

53
S. Lugo, et al. Solar Energy 189 (2019) 45–56

Fig. 12. Pool temperature comparison for the complete model simulation
versus measured data for June 14th–16th, 2017. Fig. 13. Average pool temperatures during operation hours during different
seasons.

Table 7
Monthly climate data for the typical meteorological year in Cuernavaca, 5.1. Solar heating system without auxiliary power
Morelos (ESOLMET-IER, 2018).
Table 8 shows the monthly results obtained by the model for each
Month Ta Tmin Tmax w H
component throughout a typical year without auxiliary power and
(°C) (°C) (°C) (%) (m/s) (MJ/m2 day)
without the use of a thermal pool cover.
January 19.4 12.9 26 52.9 3.15 16.45 Fig. 13 shows the average pool temperature for all the operation
February 20.55 13.65 27.35 48.4 3.1 19.21 days during the hours of pool operation from 10:00 to 18:00 h for the
March 22.55 15.4 29.75 44.2 3.02 22.00
different seasons; spring and summer are evidently the best seasons for
April 23.95 16.95 30.95 46.2 2.92 22.73
May 24.25 17.85 30.75 56.6 2.39 19.48
pool operation as the temperatures reached were greater than 25 °C.
June 22.45 17.1 27.85 71.8 2.21 19.23 The main advantage of installing solar systems is determined based
July 21.6 16.2 26.95 68.9 2.36 22.19 on an analysis of the cost and return on investment. The following as-
August 21.45 16.15 26.8 68.6 2.24 20.06 sumptions were made for this analysis:
September 20.95 15.95 25.95 73.6 2.17 15.40

• Analysis is performed for 15 years of the useful life of the solar


October 20.95 15.3 26.6 67.8 2.5 18.71
November 20.6 14.35 26.8 61 2.91 17.07
December 19.8 13.4 26.15 56.3 3.17 15.06 system.
• Input economics assumptions (see Table 9).
• Calorific power of the LP gas is 24.12 MJ/lt LP gas
Average 21.54 15.43 27.66 59.69 2.68 18.97

(INECC-SEMARNAT, 2014).
Table 8 • Emission factor is 1.58 kg CO /lt LP gas (INECC-SEMARNAT, 2014).
2
Monthly heat gains and losses without auxiliary power and thermal pool cover. • Efficiency of the auxiliary LP gas heater is 82% at sea level and 70%
Month Qeva Qrad Qconv Qsol Qcol Qaux SF
in Temixco, Morelos (altitude of 1510 m above sea level).
GJ GJ GJ GJ GJ GJ %
For this case, the heat gains are estimated at 336 GJ (see Table 8).
January −14.16 −11.66 −1.51 6.12 20.87 0 100 The energy supplied by the solar collector system was 242.5 GJ
February −14.73 −10.47 −1.47 6.42 20.86 0 100
(67.4 MW h) per year, which is equivalent to the energy obtained from
March −20.76 −11.25 −1.71 9.38 25.29 0 100
April −21.60 −11.34 −2.08 11.22 23.09 0 100 14363 litres of LP gas or 7182 USD. On taking into consideration the
May −17.39 −10.86 −1.49 9.32 20.62 0 100 cost variables (Table 9), we obtain a 5-month payback period for the
June −12.28 −11.60 −2.02 8.41 17.31 0 100 solar heating system with an internal rate of return of 286% and a net
July −13.82 −12.23 −2.12 8.18 20.91 0 100
present value in 15 years of 294586 USD and an annual avoided
August −12.75 −11.97 −1.93 7.06 19.98 0 100
September −9.40 −11.39 −1.62 5.81 15.34 0 100
emissions of 22.69 Ton of CO2.
October −13.31 −12.26 −2.09 8.58 20.03 0 100
November −13.04 −11.74 −1.76 −7.32 19.00 0 100
December −12.87 −11.54 −1.35 6.45 19.19 0 100 5.2. Solar heating system with auxiliary power
TOTAL −176.11 −138.30 −21.13 94.26 242.51 0 100
In the second case study, an auxiliary LP gas water heating system
was considered to supplement the solar collector system to maintain the
not consider the use of auxiliary power for heating, while the second pool temperature at 28 °C during the operation hours throughout the
case considered the use of an LP gas heater to supply the energy re- year. The system operates without a thermal pool cover.
quired to maintain a minimum temperature of 28 °C in the pool during Table 10 shows the results for this analysis, where it is evident that
sunny hours throughout the year. the solar fraction is reduced owing to a higher operating temperature,
thus reducing the collector efficiency. This also results in an additional

54
S. Lugo, et al. Solar Energy 189 (2019) 45–56

Table 9
Input economic assumptions.
Variable Value

SWHS key on hand price 3223 USD (MÓDULO SOLAR, 2019)


LP gas price $9.55 MX/litre (∼0.50 USD/litre) (CRE, 2018)
Annual increment of LP gas (January 2017–January 2018) 19.5% (CRE, 2018)
Annual inflation 6.77% (Banco de México, 2017)

Table 10 an internal rate of return of 218%, and the net present value of 218,745
Monthly heat gains and losses with auxiliary power without thermal pool cover. USD are obtained for an operation period of 15 years.
Qeva Qrad Qconv Qsol Qcol Qaux SF
Fig. 14 shows the pool temperatures reached during operating hours
GJ GJ GJ GJ GJ GJ % with the use of the solar heating system and auxiliary power; the
minimum set point is constant, and the pool temperature can rise to
January −22.38 −14.89 −4.30 6.12 14.93 21.17 41.4 30 °C.
February −21.35 −12.93 −3.66 6.42 16.08 15.99 50.1
March −26.08 −13.10 −3.43 9.38 21.50 12.59 63.1
The use of a thermal pool cover during the night time can sig-
April −24.49 −12.30 −2.97 11.22 21.05 7.38 74.0 nificantly reduce the heat losses from the pool. A comparison made for
May −20.73 −11.99 −2.47 9.32 18.12 7.86 69.7 case study 1 reveals that the heat losses can be reduced by 15.2%
June −16.83 −13.27 −3.42 8.41 14.07 11.41 55.2 through the use of a thermal pool cover, which would increase the
July −18.23 −13.84 −3.44 8.18 17.69 10.38 63.0
minimum pool temperatures to 26 °C throughout the year during the
August −17.67 −13.92 −3.44 7.06 16.18 11.88 57.7
September −15.18 −13.76 −3.44 5.81 10.81 15.73 40.7 pool operating hours. This could also help to reduce the number of
October −18.55 −14.34 −3.67 8.58 16.20 12.32 56.8 required collectors or increase the pool temperature.
November −19.11 −14.14 −3.67 7.32 14.79 15.08 49.5
December −20.65 −14.68 −3.91 6.45 13.44 19.61 40.7

TOTAL −241.25 −163.18 −41.81 94.26 194.86 161.40 54.7 6. Conclusions

The objective of this work was to formulate a mathematical model


that could simulate pool temperatures. Therefore, the model could be
used to estimate the amount of energy required to maintain comfort
conditions in an outdoor, partly shaded pool with a solar heating
system. Based on the models found in the literature, a program was
developed on TRNSYS 16 to simulate the heat gains and losses from the
pool, in addition to modelling an array of solar collectors and the
amount of energy that it contributes.
To simulate the pool, it was necessary to program a new type of
model in TRNSYS. This model type integrates validated equations and
the shading factor for the swimming pools in warm climates. The new
model type can be used by other authors to simulate swimming pools
with similar characteristics to those described in this work.
As an integral part of the validation for this model, the experimental
data obtained from the pool’s monitoring system for the months of
March 2016 through June 2017 were used. The validation was per-
formed for each of the model’s components as well as for the complete
model, to ensure an adequate reproduction of the measured data. As
shading can be a relevant factor affecting outdoor pools owing to the
reduction in important heat gains from direct solar radiation at the
pool’s surface, an equation for the shading factor was developed ac-
cording to the profile of the shading at the pooĺs surface for each
season.
Fig. 14. Average pool temperatures during operation hours during different
The analysis of the different evaporation equations proposed
seasons with auxiliary power.
showed that Richteŕs correlation best fits the experimental data and is
followed by the correlation of McMillan and ISO.
heat loss of 450.7 GJ (125.2 MW h) and, therefore, more energy is re- For May 19th, 2017, the simulated and experimental pool tem-
quired to maintain the set temperature of 28 °C. The energy supplied by perature data have an average margin of error of ± 0.14%
the solar system is 195 GJ (54 MW h) and reaches a total solar fraction (RMSE = 0.05 °C, MBE = −0.02 °C, and R2 = 0.9974), and a tem-
of 54.7% (solar fraction is the amount of energy provided by the solar perature difference error of ± 0.04 °C. If the shading factor is not taken
technology divided by the total energy required) with a fuel savings of into consideration, the margin of error between the simulated and ex-
11541 litres of LP gas (5770 USD) and annual avoided emissions of perimental data was ± 0.83% (RMSE = 0.31 °C, MBE = +0.21 °C, and
18.24 Tons of CO2. The solar fraction of Table 10 is less than that ob- R2 = 0.9124) with a temperature difference of ± 0.26 °C. This beha-
tained in Table 8 owing to the higher pool temperature required and the viour is similar for other test days. The margin of error indicates that
use of auxiliary power in this case in order to reach the established pool the model accurately reproduces the pool temperature values and also
temperature around the year. indicates the importance of considering the pool shading factor.
The total energy supplied by the auxiliary power was 161.4 GJ The results of the validation of the solar collector array model
(44.8 MW h), which is equivalent to 9560 L of LP gas (5353 USD) and present an average deviation of ± 0.53% (RMSE = 0.194 °C,
annual avoided emissions of 15.10 Tons of CO2. On applying the same MBE = −0.12 °C, and R2 = 0.9805) in the collector outlet temperature
analysis as that presented in Section 5.1, a payback period of 6 months, simulation and a ± 5.67% margin of error (RMSE = 1.48 kW,

55
S. Lugo, et al. Solar Energy 189 (2019) 45–56

MBE = −0.94 kW, and R2 = 0.9461) in the useful energy gain of the Govaer, D., Zarmi, Y., 1981. Analytical evaluation of direct solar heating of swimming
simulated solar collector array. This represents a temperature difference pools. Solar Energy 27 (6), 529–533.
Govind, Sodha, M.S., 1983. Thermal model of solar swimming pools. Energy Convers.
of ± 0.17 °C. The experimental instantaneous thermal efficiencies of Manage. 23, 171–175.
the solar collector array including the pump power are between 52.9% Hahne, E., Klüber, R., 1994. Monitoring and simulation of the thermal performance of
and 75.6% with an average of 67.6%. With respect to the complete solar heated outdoor swimming pools. Solar Energy 53, 9–19.
INECC-SEMARNAT, 2014. Factores de emisión para los diferentes tipos de combustibles
model (pool + solar heating system), a margin of error of ± 0.41% is fósiles y alternativos que se consumen en México. < https://www.gob.mx/cms/
reported in the simulated pool temperature (RMSE = 0.148 °C, uploads/attachment/file/110131/CGCCDBC_2014_FE_tipos_combustibles_fosiles.
MBE = −0.058 °C, and R2 = 0.9723), which represents a difference of pdf > .
ISO/TC 180/SC 4 N 140, 1995. Solar Energy – Heating Systems for Swimming Pools –
temperature of ± 0.12 °C. Therefore, it can be concluded that this Design and Installation.
model adequately simulates the behaviour of the pool temperature as Kaci, K., Merzouk, M., Kasbadji, Merzouk N., El Ganaoui, M., Sami, S., Djedjig, R., 2017.
well as the pool heat gains and losses. Dynamic simulation of hybrid-solar water heated olympic swimming pool. Energy
Proc. 139, 750–757.
Overall, from the results obtained on comparing this model with the
Katsaprakaki, D.A., 2015. Comparison of swimming pools alternative passive and active
experimental data, it can be concluded that this model can be used as a heating systems based on renewable energy sources in Southern Europe. Energy 81,
powerful tool for designing and optimising solar thermal systems for 738–753.
outdoor pool heating applications and for generating feasibility and Las Quintas, 2019. < http://fordecyt.ier.unam.mx:8080/HosteriaLasQuintas/ > .
Li, Y., Huang, G., Wu, H., Xu, T., 2018. Feasibility study of a PCM storage tank integrated
cost effectiveness proposals for the potential users of this technology. heating system for outdoor swimming pools during the winter season. J. Sci. Appl.
It has been proven that the use of a solar thermal collector for pool Therm. Eng. 134, 490–500.
heating applications is viable in Mexico both technically and econom- Lugo, S., García-Valladares, O., Best, R., Hernández, J., Hernández, F., 2019. Numerical
simulation and experimental validation of an evacuated solar collector heating
ically and realises adequate temperatures for pool use and returns on system with gas boiler backup for industrial process heating in warm climates.
investment of less than 1 year. Renew. Energy 139, 1120–1132.
McMillan, W., 1971. Heat dispersal – Lake Trawsfynydd cooling studies. In: Symposium
on Freshwater Biology and Electrical Power Generation. Part 1, pp. 41–80.
Acknowledgements MÓDULO SOLAR, 2019. < www.modulosolar.com > .
Molineaux, B., Lachal, B., Guisan, O., 1994. Thermal analysis of five outdoor swimming
This paper was partially supported by the projects CEMIESOL P09, pools heated by unglazed solar collectors. Solar Energy 53 (1), 21–26.
Richter, D., 1979. Temperatur-und Wärmehaushalt des thermisch belasteten Stechlin-
CEMIESOL P12, and PAPIIT IT102618. Laura Morales is thankful for und Nehmitzsees. Abhandlung des Meteorologischen Dienstes der DDR No. 123.
the SENER-CONACYT postdoctoral scholarship. The authors are Akademie-Verlag, Berlin.
thankful to Iris Santos and Rodrigo Cuevas for their support in the ex- Rohwer, C., 1931. Evaporation from Free Water Surfaces. United States Department of
Agriculture, Technical Bulletin No. 271.
perimental work.
Ruiz, E., Martínez, P.J., 2010. Analysis of an open-air swimming pool solar heating system
by using an experimentally validated TRNSYS model. Solar Energy 84, 116–123.
References Santos, E.T., Zárate, L.E., Pereira, E.M.D., 2013. Hybrid thermal model for swimming
pools based on artificial neural networks for southeast region of Brazil. Exp. Syst.
Appl. 40, 3106–3120.
Ahmad, I., Khan, N., 2009. Solar swimming pool heating in Pakistan. In: Proceedings of Shah, M.M., 2012. Improved method for calculating evaporation from indoor water pools.
ISES World Congress 2007, vol. I–V. Springer, Berlin Heidelberg, pp. 2033–2037. Energy Build. 049, 306–309.
ASHRAE, 2003. ASHRAE Applications Handbook. Shah, M.M., 2003. Prediction of evaporation from occupied indoor swimming pools.
Banco de México, 2017. Inflación. < http://www.anterior.banxico.org.mx/portal- Energy Build. 35, 707–213.
inflacion/index.html > . Singh, M., Tiwari, G.N., Yadav, Y.P., 1989. Solar energy utilization for heating of indoor
Barbato, M., Cirillo, L., Menditto, L., Moretti, R., Nardini, S., 2018. Feasibility study of a swimming pool. Energy Convers. Manage. 29 (4), 239–244.
geothermal energy system for indoor swimming pool in Campi Flegrei area. J Sci Smith, C., Löf, G., Jones, R., 1994. Measurement and analysis of evaporation from in-
Therm. Sci. Eng. Prog. 6, 421–425. active outdoor swimming pool. Solar Energy 53, 3–7.
Buonomano, A., De Luca, G., Figaj, R.D., Vanoli, L., 2015. Dynamic simulation and TRNSYS, 2005. TRNSYS Manual (Version 16). U.S.A. Solar Energy Laboratory, University
thermo-economic analysis of a PhotoVoltaic/Thermal collector heating system for an of Wisconsin-Madison, USA.
indoor–outdoor swimming pool. Energy Convers. Manage. 99, 176–192. Woolley, J., Harrington, C., Modera, M., 2011. Swimming pools as heat sinks for air
Chow, T.T., Bai, Y., Fong, K.F., Lin, Z., 2012. Analysis of a solar assisted heat pump system conditioners: model design and experimental validation for natural thermal behavior
for indoor swimming pool water and space heating. J. Sci. Appl. Energy 100, of the pool. Build. Environ. 46, 187–195.
309–317. Zayed, M.E., Zhao, J., Du, Y., Kabeel, A.E., Shalaby, S.M., 2019a. Factors affecting the
CRE, 2018. Energy Regulatory Commision. < https://www.gob.mx/cre/documentos/ thermal performance of the flat plate solar collector using nanofluids: a review. Solar
precios-al-publico-de-gas-lp-reportados-por-los-distribuidores > . Energy 182, 382–396.
Cunio, L.N., Sproul, A.B., 2012. Performance characterisation and energy savings of un- Zayed, M.E., Zhao, J., Elsheikh, A.H., Hammad, F.A., Ma, L., Du, Y., Kabeel, A.E., Shalaby,
covered swimming pool solar collectors under reduced flow rate conditions. Solar S.M., 2019b. Applications of cascaded phase change materials in solar water collector
Energy 86, 1511–1517. storage tanks: a review. Solar Energy Mater. Solar Cells 199, 24–49.
ESOLMET-IER, 2018. Estación Solarimétrica y Meteorológica del IER-UNAM. Consulted Zsembinszki, G., Farid, M.M., Cabeza, L.F., 2012. Analysis of implementing phase change
in June 2018. < http://esolmet.ier.unam.mx/index.html > . materials in open-air swimming pools. Solar Energy 86, 567–577.
Foncubierta, B., Lázquez, J.L., Maestre, I.R., González Gallero, F.J., Álvarez, Gómez P., Zuccari, F., Santiangeli, A., Orecchini, F., 2017. Energy analysis of swimming pools for
2018. Experimental test for the estimation of the evaporation rate in indoor swim- sports activities: cost effective solutions for efficiency improvement. Energy Proc.
ming pools: validation of a new CFD-based simulation methodology. Build. Environ. 126, 123–130.
138, 293–299.

56

You might also like