Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Available online at www.sciencedirect.

com

ScienceDirect
Solar Energy 133 (2016) 141–154
www.elsevier.com/locate/solener

CFD simulation of the thermal performance of an opaque


water wall system for Australian climate
Ting Wu, Chengwang Lei ⇑
School of Civil Engineering, The University of Sydney, Sydney, Australia

Received 15 November 2015; received in revised form 13 February 2016; accepted 4 April 2016

Communicated by: Associate Editor Yanjun Dai

Abstract

The thermal performance of an opaque water wall system is numerically investigated using the shear-stress transport (SST) k–x tur-
bulence model and the Discrete Ordinates radiation model for typical winter and summer climate conditions in Sydney, Australia. With a
periodic sol-air temperature specified on the outside Perspex panel, the energy performance of the water wall system is examined over a
range of water wall thicknesses. The computational fluid dynamics (CFD) model is compared with an experimentally validated transient
heat balance model (THBM), and a fair agreement between the CFD model and the THBM results is achieved. The present numerical
results indicate that the performance of the opaque water wall system is improved by increasing the thickness of the water column under
the winter climate condition in Sydney. It is also found that less supplementary energy is required in winter than that in summer in order
to maintain a comfortable interior temperature. Further, a comparison between the present water wall system and a conventional con-
crete wall system shows that the water wall system performs significantly better than the concrete wall system of the same thickness in the
winter climate of Sydney, whereas both the water wall and concrete wall systems have a similar performance in terms of energy savings in
the summer climate of Sydney.
Ó 2016 Elsevier Ltd. All rights reserved.

Keywords: Water wall; Passive solar energy; SST k–x turbulence model; Discrete Ordinates radiation model; Transient heat balance model

1. Introduction water walls. Both types of water wall systems have unique
advantages over other passive strategies as they have rela-
Over the last several decades, passive solar technologies tively large heat capacities and cost significantly less than
have attracted growing research interests due to increasing thermal energy storage walls using phase change materials
energy consumption by residential and commercial build- (PCMs). Further, heat stored in water may be redistributed
ings. In this regard, water wall offers an excellent solution by convection, and thus a water wall provides much
which can maintain thermal comfort in buildings while quicker heat exchange than a concrete or brick wall.
reducing energy consumption in low to medium latitude A large body of literature exists on analytical investiga-
countries. Based on the transmission of solar radiation in tions of the thermal performance of opaque water walls.
the system, water walls can be generally classified into two Balcomb and McFarland (1978) applied a simple empirical
categories, i.e. opaque water walls and semi-transparent method called Solar Load Ratio Method to estimate the
thermal performance of a Trombe wall and a water wall
with or without night-time insulation and with or without
⇑ Corresponding author. Fax: +61 2 9351 3343.
a reflector. Their results showed that the water wall
E-mail address: chengwang.lei@sydney.edu.au (C. Lei).

http://dx.doi.org/10.1016/j.solener.2016.04.001
0038-092X/Ó 2016 Elsevier Ltd. All rights reserved.
142 T. Wu, C. Lei / Solar Energy 133 (2016) 141–154

achieved a higher monthly solar heating fraction (i.e. the was achieved. During January to February of 2013, an
percentage of the space heating load supplied by the pas- improved design of the passive heat pipe system was tested
sive solar system) than that of the Trombe wall. They fur- alongside the previous HBM and full-scale prototype by
ther carried out a parametric analysis of the annual energy Robinson and Sharp (2014). Significant improvement in
savings using the above-described Trombe wall and water increasing heat transfer to the classroom and reducing heat
wall (McFarland and Balcomb, 1979). The parameters con- losses was achieved by adding a copper absorber, a thicker
sidered in their analysis included the R-value of the night- insulation, a rubber adiabatic section and exposing one
time insulation, the wall absorptivity and emissivity, the condenser directly to the room air.
thermal storage capacity, and the additional building mass Whilst the thermal performance of water wall systems
etc. They found that the performance of the water wall was has been investigated extensively using the HBM approach,
enhanced by decreasing the R-value of the night-time insu- computational fluid dynamics (CFD) modelling of water
lation and the wall absorptivity and by increasing the wall wall systems has attracted little attention due to the high
emissivity, the thermal storage capacity and the additional cost involved in CFD. Among the existing CFD studies,
building mass. Karabay et al. (2013) studied the thermal performance
Nayak et al. (1983) compared the thermal performance of a concrete wall with embedded water pipes filled
of four typical passive heating concepts, namely a Trombe with constant-temperature water. The model was two-
wall, a water wall and a solarium under two different con- dimensional (2D) and the flow was assumed to be turbulent.
figurations, one with the glazing uncovered and the other The simulations were performed for steady state conditions
with the glazing covered by a moveable insulation during only. Moustafa and Aripin (2014) evaluated the thermal
off sunshine hours. They developed a heat balance model performance of a combined water wall and porous ceramic
(HBM), which is based on the energy conservation con- pipes system for evaporative cooling in a three-dimensional
cept, to establish an energy balance associated with con- (3D) model. In their numerical model, the flow was
ductive, convective and radiative heat transfer in order to assumed to be laminar, and the simulations were again
obtain surface and fluid (i.e. water and air) temperatures. performed for steady state conditions only.
Their calculations illustrated that the water wall system It is seen from the above literature survey that the HBM
resulted in less temperature fluctuation and a higher aver- has dominated the water wall research for over three
age heat flux than the Trombe wall when night-time insula- decades, whereas the application of the CFD approach to
tion was used, and a phase shift of almost 12 h was water wall research is very limited. Although CFD is a
observed for a 0.22-m thick concrete wall next to a 0.10- powerful numerical approach, which can resolve complex
m thick water wall. The HBM has been widely adopted fluid flow and heat and mass transfer processes and has
over the past three decades, for example, by Dutt et al. been widely adopted to solve various scientific and engi-
(1987), Kaushik and Kaul (1989), Sodha et al. (1992) and neering problems, its application to water wall research
Gupta and Tiwari (2002). has been limited to steady state conditions only. The pur-
Recently, the HBM has also been applied to investigate pose of the present investigation is to develop a transient
the thermal performance of water wall with PCMs. CFD model, accounting for time variations of solar radia-
Albanese et al. (2012) carried out a bench-scale experiment tion and ambient temperature, to evaluate the thermal per-
in order to verify the HBM for a passive solar space heat- formance of an opaque water wall system for typical
ing system utilising heat pipes to transfer latent heat to a climate conditions in Sydney, Australia, which are relevant
storage water tank inside a building. It was found that to similar climate conditions in other low to medium lati-
the solar heat pipe system gave a significantly higher solar tude countries. The performance of the water wall system
heating fraction than other passive technologies, especially will also be compared with that of a conventional concrete
in cold and cloudy climates, and a good agreement was wall model under both winter and summer climate
obtained between the HBM and the laboratory experiment. conditions.
In order to further validate the above-described HBM and
the bench-scale experiment, a full-scale prototype of the
heat pipe system was designed by Robinson et al. (2013) 2. Numerical model and tests
in a classroom at the University of Louisville during the
spring heating season of 2010. During that season, the 2.1. Model formulation
maximum hourly average heat gain was only 163 W/m2.
Their field results indicated that the thermal storage water Under consideration are a 2D opaque water wall and a
tank was heated to a sufficiently high temperature to supply 2D concrete wall, both being directly attached to an uncon-
heat to the classroom even during the coldest days of the trolled room and a controlled room (see Fig. 1). The water
season. It was reported that, over a long period (4 consec- wall consists of two Perspex panels and a water column
utive days) of low solar isolation, the average hourly heat (refer to Fig. 1a), whereas the concrete wall comprises a
delivery to the classroom remained positive and was always pure concrete slab (refer to Fig. 1b).
greater than 16.6 W/m2. Again, a good agreement between The outside Perspex panel of the water wall system is
the HBM prediction and field data of full-scale prototype painted black to absorb solar radiation. The uncontrolled
T. Wu, C. Lei / Solar Energy 133 (2016) 141–154 143

Fig. 1. Schematic of two-dimensional (a) opaque water wall and (b) concrete wall models.

rooms are filled with air and the dimensions of the on the external surface of the outside Perspex panel or
uncontrolled rooms are fixed at D = H = 3 m, which concrete panel in the present study. The calculation of
resemble a typical residential space. The controlled rooms the sol-air temperature, which is the same as that applied
are kept at a constant temperature to resemble a heated in Yumrutasß et al. (2007) and Ruivo et al. (2013), is as
space in winter or an air-conditioned room in summer. follows:
The surface between the controlled and uncontrolled
aIðtÞ  eDR
rooms is referred to as Controlled Surface CS (refer to T sol-air ðtÞ ¼ T a ðtÞ þ ð1Þ
h0
Fig. 1), where the heat flux will be evaluated. The thickness
of the Perspex panel (dp) is kept constant, whereas the where T sol-air ðtÞ and T a ðtÞ are instantaneous sol-air temper-
thickness of the water column (dw) is varied in this study ature and ambient temperature respectively at the time
in order to investigate its effect on the thermal performance instant t; I(t) is the incident total solar radiation on the
of the water wall system. The thickness of the concrete slab external surface of the outside Perspex panel or concrete
(dc) is the same as that of the water column. In this study, panel; a and e are the absorptivity and emissivity of the
the computational domain covers the water wall or con- external surface; h0 is the heat transfer coefficient between
crete wall and the uncontrolled room only. the external surface and the ambient; and DR is the differ-
ence between the long-wave radiation from the sky incident
2.2. Boundary conditions on the external surface and the radiation emitted by the
external surface. In practice, DR = 0 for vertical surfaces
In order to account for the effect of solar radiation on (Yumrutasß et al., 2007). Here, the external heat transfer
the outside Perspex panel or concrete panel, sol-air temper- coefficient h0 is evaluated using the following expression
ature (refer to O’Callaghan and Probert, 1977) is specified (ASHRAE, 1975):
144 T. Wu, C. Lei / Solar Energy 133 (2016) 141–154

h0 ¼ 18:6V 0:605
loc ð2Þ radiation, and the temperatures of the outside Perspex
panel, the water column and the room air may decrease
where Vloc is given by: dramatically due to the drop of the outside temperature.

0:5 m=s for V 10 < 2 m=s When the sun rises in the next day, the outside Perspex
V loc ¼ ð3Þ panel is being heated again, and a new thermal cycle starts.
0:25V 10 else
The heat transfer processes involved in the concrete wall
Here V10 is the wind speed measured at a weather station system are similar to those in the water wall system except
10 m above the ground (m/s). for that the water wall is replaced by an opaque concrete
In order to simulate the effect of the diurnal temperature wall. Clearly convection takes place in both the water col-
variation for typical Sydney climate, a sinusoidal function umn and the air space (the uncontrolled room) in the water
of the ambient temperature is specified as: wall system, but occurs only in the air space in the concrete
wall system.
DT
T a ðtÞ ¼ T 0 þ sin½2pðt  tlag Þ=P  ð4Þ The convective flows inside the present 2D water wall
2
model and concrete wall model can be characterized by
where P is the period of the thermal cycle, which is 24 h in the Rayleigh number, the Prandtl number and the aspect
this study; T0 is the mean ambient temperature over one ratio, which are defined as follows:
thermal cycle; DT is the difference between the maximum
gbf DTH 3 mf H
and minimum temperatures over one thermal cycle; and tlag Raf ¼ ; Prf ¼ ; Af ¼ ð6Þ
is the time lag of the ambient temperature change relative mf j f jf df
to the change of solar radiation, which is set to 2 h in this where the subscript f denotes the type of fluid (i.e. ‘w’ for
study, indicating that the daily maximum ambient temper- water and ‘a’ for air); g is the acceleration due to gravity;
ature appears 2 h after the incident solar radiation peaks. mf, jf and bf are the kinematic viscosity, thermal diffusivity
In real life situations, the daily minimum ambient temper- and thermal expansion coefficient of the working fluid at
ature usually occurs at approximately one hour before sun- the reference temperature T0; df is the horizontal extent
rise, which does vary slightly throughout the year. In order of the working fluid column (dw for water and D for air);
to simplify the numerical model, a fixed time lag between and DT is the temperature difference between the maximum
the ambient temperature variation and the solar radiation and minimum sol-air temperatures over a diurnal cycle.
is assumed. It is anticipated that the relatively small varia- Here the nominal Rayleigh numbers are fixed at
tion of the occurring time of the daily minimum ambient Raw = 2.1  1013 for water and Raa = 1.52  1012 for air,
temperature relative to solar radiation will not significantly and the water and air flows are assumed to be turbulent.
affect the numerical results presented here and the major The Prandtl numbers are fixed at Prw = 7 for water and
conclusions drawn upon them. Pra = 0.71 for air respectively. The aspect ratio of the room
The incident total solar radiation on the outside surface is fixed at Aa = 1, whereas the aspect ratio of the water col-
is calculated as follows: umn Aw is a control parameter in the present study.
(  
I max sinð2pt=P Þ for ðm  1ÞP < t 6 m  12 P
IðtÞ ¼   2.3. Governing equations
0 for m  12 P < t 6 mP
ð5Þ The buoyancy-induced turbulent water and air flows are
governed by the following unsteady Reynolds Averaged
where Imax is the maximum solar radiation over one ther- Navier–Stokes and energy equations with Boussinesq
mal cycle; and m is the sequence number of the thermal assumption:
cycles.
@ui
Apart from the external surface, Surface CS is kept at a ¼0 ð7Þ
constant temperature, whereas the top and bottom surfaces @xi
   
are assumed adiabatic. All the other internal surfaces are @ui @ðui uj Þ 1 @p 1 @ @ui @uj 0 0
þ ¼ þ l þ  qf u i u j
coupled between solid and fluid. @t @xj qf @xi qf @xj f @xj @xi
Under the above boundary conditions, the temperature
 gi bf ðT  T 0 Þ ð8Þ
of the outside Perspex panel in the water wall system rises " #
during the day and heat is transferred to the adjacent water @T @ðuj T Þ 1 @ kf @T 0 0
column due to heat conduction through the Perspex panel. þ ¼  qf u i T ð9Þ
@t @xj qf @xj C pf @xj
A convective flow is then induced by buoyancy in the water
column, and heat is quickly transferred to the inside Per- where xi and xj are the Cartesian coordinates in the i and j
spex panel in addition to that stored in the water column. directions (i, j = 1 and 2 corresponding to the x and y
Subsequently, heat is further transferred from the inside directions respectively); t is the time; p is the pressure; T
Perspex panel to the adjacent air. As a consequence, the and T 0 are the mean and fluctuating temperatures; ui and
uncontrolled room air is heated up. At night, the heat uj are the mean velocity components in the i and j direc-
transfer process may be reversed due to the absence of solar tions; u0 i and u0 j are the corresponding fluctuating velocity
T. Wu, C. Lei / Solar Energy 133 (2016) 141–154 145

components in the i and j directions; and qf, kf, Cpf and lf Since water is a participating media, the DO radiation
are the fluid density, thermal conductivity, specific heat and model is adopted. In the DO radiation model, the direc-
dynamic viscosity, respectively. tional variation of the radiative intensity is represented
by a discrete number of ordinates, and integrals over solid
2.4. Turbulence models angles are approximated by numerical quadrature
(Modest, 2013). The radiative transfer equation is solved
In order to close Eqs. (7)–(9), the Reynolds stresses for a finite number of discrete solid angles for as many
(qf u0i u0j ) and the turbulence heat fluxes (u0i T 0 ) must be transport equations as there are in an associated vector
modelled. Different turbulence models have different treat- direction at spatial location (x, y). The equation used in
ments of the Reynolds stresses and the turbulence heat the DO model is written as:
fluxes. In this study, the shear-stress transport (SST) k–x r  ðIð~
r;~ sÞ þ ða þ rs ÞIð~
sÞ~ r;~

model is adopted due to its accuracy for a related problem 4 Z 4p
rT rs
demonstrated in a previous investigation (Wu and Lei, ¼ an2 þ Ið~ s0 ÞUð~
r;~ s0 ÞdX0
s ~ ð15Þ
2015a). In the SST k–x model, the Reynolds stresses are p 4p 0
modelled through the Boussinesq approximation as: where ~r, ~ s0 are the position vector, the direction vec-
s, and ~
 
@ui @uj 2 tor and the scattering direction vector respectively; a and rs
qf u0i u0j ¼ lt þ  qf kdij ð10Þ
@xj @xi 3 are the absorption coefficient and the scattering coefficient;
n is the refractive index; r is the Stefan–Boltzmann con-
where lt is the turbulent eddy viscosity; k is the turbulent stant (5.669  108 W/m2-K4); I is the radiation intensity,
kinetic energy; and dij is the Kronecker delta (dij = 0 if which depends on the position (~ r) and direction (~ s); and
i – j and dij = 1 if i = j). The turbulent eddy viscosity is T, U, and X0 are the local temperature, the phase function
computed from: and the solid angle respectively.
qf k 1
lt ¼ h i ð11Þ
x max 1 ; SF 2 2.6. Numerical scheme

a a1 x

where x is the specific dissipation rate; a1 = 0.31; and a* is The governing equations including the two additional
a coefficient used to predict transition from laminar to tur- transport equations for the turbulence kinetic energy (k)
bulent flow (a* = 1 in fully turbulent flows) (Wilcox, 1998). and the specific dissipation rate (x) along with the specified
F2 is a coefficient computed from an additional equation boundary and initial conditions are solved using the CFD
and S is the modulus of the mean rate-of-strain tensor package ANSYS FLUENT 14.0, which is a finite-volume
which is defined as: based solver. The pressure-velocity coupling is carried out
pffiffiffiffiffiffiffiffiffiffiffiffiffi using the SIMPLE scheme. The advection terms in the gov-
S ¼ 2S ij S ij ð12Þ erning equations are discretised by a second-order upwind
where Sij is the mean rate of strain tensor and is given by: scheme and the diffusion terms are discretised using a
  second-order central-differencing scheme. A second-order
1 @ui @uj
S ij ¼ þ ð13Þ implicit time-marching scheme is adopted for the unsteady
2 @xj @xi term.
More details about the SST k–x model can be found in
Menter (1994). 2.7. Grid and time-step dependency tests
The turbulence heat fluxes are modelled as:
lt @T Grid and time-step dependency tests have been con-
u0i T 0 ¼  ð14Þ ducted based on a case with the thickness of the water col-
qf rT @xi
umn dw = 0.15 m under a constant cooling condition. In
where rT is the turbulent Prandtl number, which is equal to this case, Surface CS is maintained at a constant tempera-
0.85. In the SST k–x model, two additional transport equa- ture Th = 22 °C (corresponding to a heated space), and the
tions (one for the turbulence kinetic energy (k), and the outside Perspex panel is kept at a constant temperature
other for the specific dissipation rate (x)) are solved. Fur- Tp = 3 °C (corresponding to a cold weather condition in
ther details of the SST k–x model can be found in ANSYS winter). Initially, the fluid and solid temperatures in the
FLUENT User’s Guide (Fluent Inc., 2011). model are set to T 0 ¼ ðT h þ T p Þ=2 ¼ 12:5  C, and the water
and air are assumed to be stationary. For the grid depen-
2.5. Radiation model dency test, three different non-uniform meshes 60  78,
100  130 and 200  260 with a fixed time-step of 0.5 s
Among all the radiation models available in ANSYS are calculated for 12 h. Since the energy efficiency of the
FLUENT 14.0, the Discrete Ordinates (DO) radiation water wall system may be ascertained by the heat flux
model is the only model that can deal with radiation prob- through the Control Surface CS, this heat flux is calculated
lems with both participating and non-participating media. for comparison. The time series of the heat flux obtained
146 T. Wu, C. Lei / Solar Energy 133 (2016) 141–154

with the three meshes are shown in Fig. 2a. The negative Table 1
heat fluxes over the time period of 12 h indicate that heat Comparisons of the statistical results of the predicted heat flux through
Surface CS with different meshes and time-steps.
continuously flows from the controlled room to the uncon- qmax qref;max qqref
trolled room, which is expected. It can be seen in this figure Case no. Mesh Time-step (s) qref;max ð%Þ qref ð%Þ
that the results obtained with the two finer meshes are very 1 100  130 0.5 0 0
similar, whereas the result obtained with the coarsest mesh 2 60  78 0.5 2.72 2.46
shows a discernible variation from those obtained with the 3 200  260 0.5 0.19 0.42
4 100  130 0.25 0.10 0.01
finer meshes.
5 100  130 0.125 0.08 0.01
A quantitative comparison of the results obtained with
these three meshes is shown in Table 1. Here,
qref;max ¼ 35:336 W=m2 and qref ¼ 44:077 W=m2 are The comparison of Cases 1, 2 and 3 in Table 1 demon-
the maximum and average heat fluxes through Surface strates that the 100  130 mesh is capable of producing a
CS respectively over the time period of 12 h obtained in sufficiently accurate solution. The medium mesh,
the reference case (Case 1) with the 100  130 mesh and i.e. 100  130, is therefore adopted for subsequent calcula-
the 0.5 s time-step. qmax  qref;max represents the variation tions in consideration of computing costs.
of the maximum heat fluxes between a test case and the Following the above mesh dependence test, three time-
reference case, and q  qref is the variation of the average steps of 0.5 s, 0.25 s and 0.125 s respectively with the
heat fluxes between a test case and the reference case. 100  130 mesh are tested for the same water wall model.

Fig. 2. Time history of the predicted heat flux through Surface CS with (a) different meshes and (b) different time-steps.
T. Wu, C. Lei / Solar Energy 133 (2016) 141–154 147

The heat fluxes through the same monitoring surface which resemble a typical winter climate condition in Syd-
obtained with these three time-steps are shown in Fig. 2b. ney, Australia. The difference between the daily maximum
It is observed in this figure that the results obtained with and minimum ambient temperatures DTa is set to 15 °C,
the different time-steps are almost identical. A quantitative whereas the wind speed is assumed to be 3.944 m/s accord-
comparison of the heat fluxes calculated in Cases 1, 4 and 5 ing to the statistical climate data for Sydney in July. Sur-
with different time-steps is given in Table 1, which confirms face CS is maintained at Tc = 22 °C which resembles a
that the variations of the results are indeed negligible. In typical heated space in winter. The water and air within
consideration of the numerical accuracy and the computa- the computational domain is initially stationary with a uni-
tional time, the 100  130 mesh and the 0.5 s time-step are form temperature T0. The simulations of both the CFD
selected for the present model. Similarly, a mesh and time- model and the THBM have been conducted over seven
step dependency test for the concrete wall model shows thermal cycles (days), which begins at 6am in Day 1.
that a 100  130 mesh and a 0.5 s time-step provide suffi- Fig. 3a shows the time series of the averaged water and
cient spatial and temporal resolutions for the concrete wall room air temperatures in the 2D CFD model and the pre-
system, and thus are adopted in this study. dicted water and air temperatures by the one-dimensional
(1D) THBM. Here the average temperatures of the water
3. Results and discussions column and the room air in the CFD model are evaluated
as:
The following subsections present the numerical results Z H Z dp þdw
1
obtained from 2D unsteady turbulent simulations using Tw ¼ T w dxdy ð16Þ
the SST k–x model and the DO radiation model for the dw H 0 dp

above-described water wall and concrete wall models with Z H Z 2dp þdw þD
1
periodic sol-air temperature variations on the external sur- Ta ¼ T a dxdy ð17Þ
DH 0 2dp þdw
faces. In Section 3.1, the present CFD model of a water
wall system is compared with an experimentally validated It is seen in Fig. 3a that a generally good agreement
transient heat balance model (THBM), and the thermal between the CFD model and the THBM is achieved for
stratification in the water column and in the uncontrolled the predicted water and air temperatures, with the differ-
room air is demonstrated. In Section 3.2, the thermal per- ence in the predicted air temperatures much less than that
formance of the water wall system is described in terms of of the water temperatures. It is also clear in Fig. 3a that the
the water and uncontrolled room air temperatures as well water temperature fluctuates over an approximate range of
as the heat fluxes through Surface CS. A parametric study 20–30 °C, which is significantly greater than that of the air
is also presented to illustrate the effect of the thickness of temperature. The daily maximum temperatures in the
the water column on the thermal stratifications in water water and air generally occur before 6 pm, whereas the
and air, the temperature fluctuation in the uncontrolled daily minimum temperatures in the water and air generally
room and the supplemental energy consumption of the occur after 6 am.
controlled room. In Section 3.3, quantitative analyses are Since both the CFD and the THBM models start with
performed to investigate the thermal response to different stationary fluids and a uniform temperature in the water
sol-air temperature variations, which resemble typical win- wall system, start-up effects of these models are evident in
ter and summer climate conditions. A comparison is also Fig. 3a. However, quasi-steady states are established in
made between the water wall model and a concrete wall the CFD and THBM models after several thermal cycles.
model to illustrate the advantages of the water wall system. In order to quantify the start-up effect of the individual
models, the time series of the daily maximum water and
3.1. Validation of the CFD model against the THBM air temperatures are plotted in Fig. 3b. It is clear in this fig-
ure that both the daily maximum water and air tempera-
The above-described CFD model for the opaque water tures converge quickly as the number of thermal cycle
wall system is compared with a transient heat balance increases, with the CFD model converging slightly more
model (Wu and Lei, 2015b), which has been validated pre- quickly than the THBM. After 6 full thermal cycles, the
viously against experimental measurements. In the THBM, variations of both the water and air temperatures become
the time variations of internal convective heat transfer less than 0.5% in both models. It is therefore assumed that
coefficients, the radiation emitted by external surfaces a quasi-steady state has been established in the water wall
and the internal surface radiation exchanges are all system after 6 thermal cycles.
accounted for. Details of the THBM and its validation Fig. 3c further displays the quantitative comparison of
can be found in Wu and Lei (2015b). the maximum daily water and uncontrolled room air tem-
The comparison of the CFD model and the THBM is peratures predicted by the CFD model and the THBM. It
carried out with the thickness of the water column dw fixed is seen in this figure that the differences between the CFD
at 0.15 m. The mean daily ambient temperature T0 and the model and the THBM in the predicted maximum water
maximum daily solar radiation of the outside Perspex and air temperatures both reduce with the increase of the
panel Imax are fixed at 12.5 °C and 800 W/m2 respectively, thermal cycles, suggesting that the variations between the
148 T. Wu, C. Lei / Solar Energy 133 (2016) 141–154

Fig. 3. Comparisons of the 2D CFD model and the 1D THBM. (a) Time series of predicted water and air temperatures; (b) Time series of predicted
maximum water and air temperatures; and (c) Variations of the differences between the CFD model and the THBM in the predicted daily maximum water
and air temperatures.
T. Wu, C. Lei / Solar Energy 133 (2016) 141–154 149

CFD model and the THBM are at least in part due to the The variations between the CFD model and the THBM
start-up effects of the diurnal models. Unfortunately, may also be attributed to the fact that the CFD model is
extending the CFD simulations to more diurnal cycles is 2D, whereas the THBM is only 1D. The effect of the hor-
not feasible due to the constraints of available computing izontal boundaries is neglected in the 1D THBM, but is
resources. In order to minimise the start-up effects and accounted for in the 2D CFD model. Since full-scale exper-
focus on the thermal performance of the water wall system iment is expensive to run and difficult to control under real-
at the quasi-steady state, only the data obtained in the last istic climate conditions, the predicted results from the
thermal cycle (i.e. in Day 7) are analysed and presented in experimentally validated THBM provide an alternative
what follows. way to validate the present CFD model. It is confirmed

Fig. 4. Snapshots of isotherms in the (a) water column and (b) uncontrolled room air obtained with the water wall thickness dw = 0.15 m in Day 7 under
the winter climate condition.
150 T. Wu, C. Lei / Solar Energy 133 (2016) 141–154

through the above comparisons that the present CFD 0.5 m in the water column or uncontrolled room, respec-
model is capable of predicting the water and room air tem- tively. DH is the difference between the higher and lower
peratures in the opaque water wall system to a reasonable points where the temperatures are extracted (i.e.
level of accuracy, and thus will be adopted to explore the DH = 2 m in the present case). The comparison of
thermal performance of the opaque water wall system over Fig. 5a and b confirms that the temperature stratification
a range of parameters. in the water column is indeed stronger than that in the
Whilst the THBM can quickly predict the major fea- uncontrolled room air. The former varies over the range
tures of a water wall system, it does not resolve the details of 0.05–0.1 for the different thicknesses of the water col-
of the convective flows pertinent to the water wall system umn, whereas the latter only varies over the range of
due to the 1D nature of the model. In contrast, the CFD 0.01–0.05. It is also noteworthy that the temperature strat-
approach can resolve full details of the convective flows ifications of both water and uncontrolled room air decrease
such as the thermal stratifications in the water column with the increase of the water column thickness.
and air space. Snapshots of isotherms in the water column Fig. 6 illustrates the time series of the average water and
and uncontrolled room air obtained with the thickness of air temperatures obtained with different water column
the water column dw = 0.15 m in the seventh thermal cycle thicknesses under the winter climate condition. It is clear
under the winter climate condition are presented in Fig. 4. in Fig. 6 that the temperature fluctuations of both water
It is seen from the isotherms that the interior thermal strat- and air reduce with the increase of the water column thick-
ifications are evident in both the water column and uncon- ness, and the temperature fluctuation of water is much lar-
trolled room air, although the strength of the temperature ger than that of air in the present opaque water
stratification in the water column is evidently stronger than wall system. This result along with the results shown in
that in the uncontrolled room air. Figs. 3–5 demonstrates that the water column acts as a
buffer layer to mitigate temperature fluctuations in the
3.2. Effect of the thickness of the water column uncontrolled room air.
Quantitative data indicating the temperature fluctua-
For comparison purpose, the thickness of the water col- tions in water and air obtained with different thicknesses
umn dw is varied from 0.075 m to 0.3 m in this study. All are listed in Table 2. In this table, the temperature fluctua-
the boundary conditions remain the same as those tion index (TFI) is calculated as (Tiwari and Singh, 1996):
described in Section 3.1. In order to quantify the tempera- T max  T min
ture stratifications in the water column and in the uncon- TFI ¼ ð19Þ
T max þ T min
trolled room air, time series of the vertical temperature
gradient (VTG) obtained with different water column where Tmax and Tmin are the daily maximum and minimum
thicknesses are plotted in Fig. 5. Here, the VTG is defined averaged water or air temperatures, respectively. The
as: results in Table 2 show that the temperature fluctuation
ðT 2:5  T 0:5 Þ=DH of water is almost three times than that of uncontrolled
VTG ¼ ð18Þ room air for all the water column thickness considered.
DT =H
For the purpose of exploring energy saving potential of
where T2.5 and T0.5 are the water or air temperatures various energy saving strategies, the heat flux to or from a
extracted along the centre line at the heights of 2.5 m and temperature-regulated room has been examined in the

Fig. 5. Time series of vertical temperature gradient in (a) the water column and (b) the uncontrolled room air obtained with different water column
thicknesses.
T. Wu, C. Lei / Solar Energy 133 (2016) 141–154 151

Fig. 6. Time series of the averaged (a) water and (b) uncontrolled room air temperatures obtained with different water column thicknesses.

thickness results in a larger thermal energy storage capac-


Table 2
Comparisons of the temperature fluctuation index (TFI) of water and air
ity, and thus the heat transfer into and out of the controlled
obtained with different water column thicknesses. room becomes smaller. When the water column thickness is
Water wall thickness dw (m) Water Air
0.3 m, the heat flux is positive throughout the thermal
cycle, suggesting that no supplemental energy is needed
0.075 0.313 0.106
0.15 0.177 0.060
for space heating in the controlled room.
0.3 0.094 0.032 Fig. 7b presents the supplemental energy consumptions
for heating the controlled room calculated with the differ-
ent water column thicknesses. Here, the supplemental
literature (e.g. Nayak et al., 1983; Kaushik and Kaul, energy consumption is calculated as:
1989). Here the time series of the heat flux entering into Z t2
the controlled room is presented in Fig. 7a for the different E¼ q00 dt for q00 < 0 ð20Þ
water column thicknesses. In this figure, a positive heat flux t1

means that heat is transferred from the uncontrolled room where q00 is the instantaneous heat flux from the uncon-
to the controlled room, whereas a negative heat flux repre- trolled room to the controlled room, and t1 and t2 are
sents heat transfer in the opposite direction. It is found in two time instants, over which period the supplemental
Fig. 7a that both the heat fluxes entering into the con- energy consumption is integrated. As described above,
trolled room and transferring out of the controlled room the negative heat fluxes represent that the air temperature
decrease with the increase of the water column thickness. in the uncontrolled room is lower than that in the con-
This is because that an increase of the water column trolled room, and thus heat is supplied to the controlled

Fig. 7. (a) Time series of the heat flux entering into the controlled room; (b) comparisons of supplemental energy consumptions for space heating with the
different water column thicknesses.
152 T. Wu, C. Lei / Solar Energy 133 (2016) 141–154

room in order to maintain the constant temperature Tc. the water column dw and the concrete slab dc fixed to
Fig. 7b indicates the supplemental energy consumptions 0.15 m. One climate condition is the typical winter condi-
decrease significantly with the increase of the water column tion in Sydney, Australia described in Section 3.1, and
thickness, and when the water column thickness is 0.3 m, the other climate condition is the typical summer condition
no supplemental energy for heating the space is required. in Sydney with a diurnal mean temperature of T0 = 27.5 °C
Therefore, increasing the water column thickness of the and the maximum solar radiation of Imax = 1000 W/m2.
present opaque water wall system can save a significant For the summer condition, the difference between the daily
amount of supplemental energy for space heating in winter. maximum and minimum ambient temperatures is also set
to DT = 15 °C, whereas the wind speed is assumed to be
3.3. Comparison of water wall and concrete wall models 3.681 m/s according to the statistical climate data
under different climate conditions for Sydney in January. Surface CS is maintained at
Tc = 25 °C, which is a typical temperature setting for an
In this section, the concrete wall model shown in Fig. 1b air-conditioned room in summer.
is compared with the opaque water wall model under two Fig. 8 presents the time series of the uncontrolled room
different climate conditions with both the thicknesses of air temperature in the water wall and concrete wall models

Fig. 8. Comparisons of the uncontrolled room air temperatures and the time lags in the water wall and concrete wall models under (a) winter and (b)
summer climate conditions.
T. Wu, C. Lei / Solar Energy 133 (2016) 141–154 153

Table 3 have a worse overall performance in summer than that in


Comparisons of the time lags in the water wall and concrete wall models winter.
under the different climate conditions.
The time lags of the daily peak temperature calculated
Climate conditions Water wall (h) Concrete wall (h) for the water wall and concrete wall models are also shown
Winter 5.45 4.48 in Fig. 8, which compares the time histories of the averaged
Summer 5.08 4.45 air temperature in the uncontrolled room against that of
the prescribed Sol-air temperature in Day 7. Clearly the
under the winter and summer climate conditions respec- time lag is larger in the water wall model than that in the
tively. It is worth noting that the start-up effect of the concrete wall model under both climate conditions, which
CFD model has also been tested for the concrete wall sys- means that the uncontrolled room air temperature
tem, and the results show that a quasi-steady state is responds to the switch of the thermal forcing more quickly
achieved in the concrete wall system after only 2 full ther- in the concrete wall system than that in the water all sys-
mal cycles. The data presented in Fig. 8 for the concrete tem. A quantitative comparison of the time lags in the
wall model is extracted from Day 7, which is consistent water wall and concrete wall models under the different cli-
with that for the water wall model. It is clear in this figure mate conditions is given in Table 3. It is found in this table
that the uncontrolled room air temperature in the water that the concrete wall model results in a similar time lag to
wall model has a smaller fluctuation in both winter and that of the water wall model under both climate conditions.
summer than that in the concrete wall model. Under the It is also seen in Table 3 that the time lags resulting from
winter condition (refer to Fig. 8a), the uncontrolled room the concrete wall system in winter and summer are almost
air temperature is higher than the controlled room temper- the same, whereas the time lag resulting from the water
ature for most of the time in the water wall model, whereas wall model in summer is clearly less than that in winter.
in the concrete wall model, the daily maximum temperature Fig. 9 depicts the calculated heat fluxes through Surface
in the uncontrolled room is significantly higher and the CS in the water wall and concrete wall models under winter
daily minimum temperature in the uncontrolled room is and summer climate conditions. Similar to that observed
significantly lower than the controlled room temperature. for the uncontrolled room air temperature, the fluctuation
Therefore, the water wall model is better than the concrete of the heat fluxes is evidently smaller in the water wall
wall model in maintaining a stable temperature in the model than that in the concrete wall model. It is seen that
uncontrolled room under the winter condition. This is also both the water wall and concrete wall models produce
true under the summer condition (refer to Fig. 8b). How- higher heat fluxes in summer than that in winter. Further,
ever, from the thermal comfort and energy consumption the heat fluxes are positive throughout almost the whole
points of view, the concrete wall model is better than the thermal cycles under the summer condition, which means
water wall model during the late night and early morning excessive heat must be removed by air-conditioning in
in summer. The comparison of Fig. 8a and b also shows order to maintain a comfort environment in the controlled
that the uncontrolled room air temperatures in both the room. In winter, however, the heat loss from the controlled
water wall and concrete wall models are significantly higher room must be compensated by supplemental heating in
than the controlled room temperature in summer, which order to maintain thermal comfort. It can be seen in
indicates that both the water wall and concrete wall models Fig. 9 that negative heat flux occurs for approximately half

Fig. 9. Comparisons of the heat fluxes through Surface CS in the water wall and concrete wall models under the different climate conditions.
154 T. Wu, C. Lei / Solar Energy 133 (2016) 141–154

Table 4 Acknowledgement
Comparisons of supplemental energy consumptions in the water wall and
concrete wall models under the different climate conditions.
The financial support of the Australian Research Coun-
Climate conditions Water wall (kJ/m2) Concrete wall (kJ/m2) cil (Discovery Project Grant DP130100900) is gratefully
Winter 63.11 897.94 acknowledged.
Summer 3613.32 3652.93

References
of the diurnal cycle in the concrete wall model in winter,
Albanese, V., Robinson, S., Brehob, G., Sharp, K., 2012. Simulated and
whereas the occurrence of negative heat flux is insignificant experimental performance of a heat pipe assisted solar wall. Sol.
in the water wall model. This comparison indicates that the Energy 86, 1552–1562.
water wall model has a better thermal performance than ASHRAE, 1975. Procedure for Determining Heating and Cooling Loads
the concrete wall model in winter. for Computerized Energy Calculations: Algorithms for Building Heat
A further quantitative comparison of the supplemental Transfer Sub Routines. ASHRAE, New York.
Balcomb, J.D., McFarland, R.D., 1978. A simple empirical method for
energy consumptions in the water wall and concrete wall estimating the performance of a passive solar heated building of the
models under the different climate conditions is given in thermal storage wall type. 2nd National Passive Solar Conference,
Table 4. It is clear that both the water wall and concrete Philadelphia, PA.
wall models consume significantly more supplemental Dutt, D.K., Rai, S.N., Tiwari, G.N., Yadav, Y.P., 1987. Transient analysis
energy in summer than that in winter. The difference in of a winter greenhouse. Energy Convers. Manage. 27 (2), 141–147.
Fluent Inc., 2011. ANSYS FLUENT User’s Guide. USA.
the supplemental energy consumptions between the water Gupta, A., Tiwari, G.N., 2002. Computer model and its validation for
wall and the concrete wall models is large in winter, but prediction of storage effect of water mass in a greenhouse: a transient
insignificant in summer. analysis. Energy Convers. Manage. 43 (18), 2625–2640.
Karabay, H., Arıcı, M., Sandık, M., 2013. A numerical investigation of
4. Conclusions fluid flow and heat transfer inside a room for floor heating and wall
heating systems. Energy Build. 67, 471–478.
Kaushik, S.C., Kaul, S., 1989. Thermal comfort in buildings through a
In this study, the thermal performance of an opaque mixed water-mass thermal storage wall. Build. Environ. 24, 199–207.
water wall system is numerically investigated for different McFarland, R.D., Balcornb, J.D., 1979. Effect of design parameter
thicknesses of the water column by means of transient changes on the performance of thermal storage wall passive systems.
CFD modelling for the climate of Sydney, Australia. To Third National Passive Solar Conference, San Jose, California.
Menter, F.R., 1994. Two-equation eddy-viscosity turbulence models for
model the turbulence flows and radiation transfer in the engineering applications. AIAA J. 32, 1598–1605.
water wall system, the shear-stress transport k–x turbu- Modest, M.F., 2013. Radiative Heat Transfer, third ed. Academic Press,
lence model and the Discrete Ordinates radiation model New York.
have been adopted. The present numerical results have Moustafa, M.A., Aripin, S., 2014. CFD evaluation of the pottery water wall
revealed that the thermal stratification and the temperature in a hot arid climate of Luxor, Egypt. J. Green Build. 9 (4), 175–189.
Nayak, J.K., Bansal, N.K., Sodha, M.S., 1983. Analysis of passive heating
fluctuation decrease with the increase of the water column concepts. Sol. Energy 30, 51–69.
thickness in the present water wall model under the typical O’Callaghan, P.W., Probert, S.D., 1977. Sol-Air temperature. Appl.
winter climate condition in Sydney, Australia. The energy Energy 3, 307–311.
performance in terms of supplemental energy consumption Robinson, B.S., Chmielewski, E., Knox-Kelecy, A., Brehob, G., Sharp,
is also enhanced by increasing the water column thickness. M.K., 2013. Heating season performance of a full-scale heat pipe
assisted solar wall. Sol. Energy 87, 76–83.
In addition, this study also shows that the thermal perfor- Robinson, B.S., Sharp, M.K., 2014. Heating season performance
mance of the water wall in winter is better than that in sum- improvements for a solar heat pipe system. Sol. Energy 110, 39–49.
mer since significantly less supplemental energy Ruivo, C.R., Ferreira, P.M., Vaz, D.C., 2013. Prediction of thermal load
consumption is needed in winter. A comparison between temperature difference values for the external envelope of rooms with
the present water wall system and a conventional concrete setback and setup thermostats. Appl. Therm. Eng. 51 (1), 980–987.
Sodha, M.S., Kaur, J., Sawhney, R.L., 1992. Effect of storage on thermal
wall system indicates that the temperature fluctuations are performance of a building. Int. J. Energy Res. 16 (8), 697–707.
smaller in the water wall system than that in the concrete Tiwari, G.N., Singh, A.K., 1996. Comparative studies of different heating
wall system for both winter and summer climate condi- techniques of a non-air conditioned building. Build. Environ. 31, 215–
tions, and a relatively larger time lag is found in the water 224.
wall model for both climate conditions considered. The Wilcox, D.C., 1998. Turbulence Modeling for CFD. DCW Industries Inc.,
La Canada, California.
water wall system requires significantly less supplemental Wu, T., Lei, C., 2015a. On numerical modelling of conjugate turbulent
energy for space heating than that of the concrete wall in natural convection and radiation in a differentially heated cavity. Int.
winter, whereas both the water wall and concrete wall sys- J. Heat Mass Transf. 91, 454–466.
tems have similar energy performance in summer. There- Wu, T., Lei, C., 2015b. Thermal modelling and experimental validation of
fore, it is concluded that an opaque water wall system is a semi-transparent water wall system for Sydney climate. Sol. Energy,
submitted for publication.
more suitable than a concrete wall system for regions Yumrutas, R., Kaska, O., Yıldırım, E., 2007. Estimation of total
with climate conditions similar to the winter condition in equivalent temperature difference values for multilayer walls and flat
Sydney, Australia. roofs by using periodic solution. Build. Environ. 42, 1878–1885.

You might also like