Vous êtes sur la page 1sur 9

Performance and inuence of numerical sub-models on the CFD simulation

of free and forced convection in double-glazed ventilated facades


M. Coussirat
a,
*, A. Guardo
a
, E. Jou
a
, E. Egusquiza
a
, E. Cuerva
b
, P. Alavedra
b
a
Centre de Diagno`stic Industrial i Fluidodina`mica (CDIF), Universitat Polite`cnica de Catalunya, Av. Diagonal 647, ETSEIB. Pab D-1, 08028 Barcelona, Spain
b
Catedra UPC Grupo JG, Universitat Polite`cnica de Catalunya, Av. Diagonal 647, 08028 Barcelona, Spain
1. Introduction
Over the last decades trends in facade design, particularly for
ofce buildings, have changed radically and large double-glazing
envelopes are now used more frequently. Double-glazed facades
(DGF), resulting from an architectural movement in which
aesthetic considerations play a relevant role [1], are currently
being tted in both new and existing buildings.
Even in moderate climatic zones, fully glazed buildings need
shading devices in order to reduce cooling loads. The current trend
is migrating fromexternal shading to internal shading which is less
expensive and offers mechanical and aesthetic advantages [2]. In
the 1970s the climate-window concept was developed, and the
forced ventilated cavity concept was introduced. A more recent
development is the double-skin facade concept, in which the cavity
is ventilated by means of a free or forced convection ow of
ambient air.
Since thermal comfort must be assured for the buildings
occupants, once the conguration of the facade is selected, energy
concerns must be carefully analyzed for the facade to performwell.
Predicting the performance of a double-skin facade is not a trivial
exercise. The temperature and airows are the result of many
simultaneous thermal, optical, and free/forced convective turbu-
lent uid ow processes which interact and are highly dynamic.
These processes depend on geometric, thermo-physical, optical
and aerodynamic properties of the various components of the
double-skin facade structure and of the building itself. The
temperature inside the ofces, the ambient temperature, wind
Energy and Buildings 40 (2008) 17811789
A R T I C L E I N F O
Article history:
Received 19 July 2007
Received in revised form 14 March 2008
Accepted 18 March 2008
Keywords:
Double-glazed facades
Building thermal performance
Computational uid dynamics
Turbulence models
Numerical model validation
A B S T R A C T
Double-glazed facades (DGF) are an attractive option in contemporary architecture and are increasingly
usedincommercial buildings. Theyoffer some advantages comparedwithsinglefacade systems but require
careful design. Thesolar-collector-likeconstructionleads tohightemperatures inthefacadecavities andthe
possibility of the building overheating. This is undesirable effect, especially in Mediterranean climates. A
possible solution for reducing thermal overheating is to use the air channel between the two layers of glass
to evacuate the solar radiation absorbed by the facade. A suitable simulation procedure for modeling these
facades would be very useful for designing buildings of this type.
The use of computational uid dynamics (CFD) has been broadly extended in order to gain insight into
this problem, but selecting suitable sub-models for the convection, radiation and turbulence effects
remains a big challenge. In this work, several modeling tests were carried out on a well-documented
experimental test case taken from the open literature in order to obtain a suitable model of the
aforementioned thermo-uid-dynamics effects. Fluid and solid phase temperatures for a DGF
conguration were obtained for three different radiation models and ve different turbulence models,
compared with experimental results available in the literature, and validated according to numerical
verication and validation methodologies. From the results obtained it can be concluded that the P-1
radiation model seems to better predict the temperature of the solid phases present in the double facade.
The RNG k e turbulence model seems to perform better than the other turbulence models tested for
predicting heat transfer when there are zones of low velocities within the facade conguration. Only this
combination of sub-models achieved numerical validation at pre-dened levels for the tested case.
2008 Elsevier B.V. All rights reserved.
* Corresponding author. Tel.: +34 93 4015943; fax: +34 93 4015812.
E-mail address: coussirat@mf.upc.edu (M. Coussirat).
Abbreviations: CFD, computational uid dynamics; CVA, control volume
approach; DGF, double-glazed facade(s); DO, discrete ordinates radiation model;
DTRM, discrete transfer radiation model; SA, SpalartAllmaras turbulence model;
SKE, standard k e turbulence model; SKW, standard k v turbulence model;
SSTKW, shear stress transport k v turbulence model; RKE, realizable k e tur-
bulence model; RNGKE, renormalization groups k e turbulence model.
Cont ent s l i st s avai l abl e at Sci enceDi r ect
Energy and Buildings
j our nal homepage: www. el sevi er . com/ l ocat e/ enbui l d
0378-7788/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.enbuild.2008.03.009
speed, wind direction, transmitted and absorbed solar radiation
and angles of incidence each of which are highly transient
govern the main driving forces. This typically results in highly
erratic airows inside the structure [2].
The advantages and disadvantages of DGF have been exten-
sively debated (see e.g. [3,4]). A critical review of double-glazed
facades froma building physics point of viewis given by Gertis [5].
He argues that the suitability of double-glazed facades for
continental climate areas is questionable. Whereas glass facades
have the advantage of saving a large amount of energy in winter,
reducing both illumination and heating costs by taking advantage
of solar radiation and day light [6], they may become highly
inefcient in summer. In Mediterranean climates, where large
solar gains are a constant condition throughout the year, such large
semi-transparent areas can produce signicant over-heating in
buildings, even in winter [7,8].
The difference in the performance of natural and forced uid
owin DGF is also not clear. Hien et al. [9] point out that a DGF with
natural convection (the stack effect) is effective enough to extract
solar heat gain inside the facade cavity and maintain a lower
internal surface temperature. Accelerating air change rates using a
mechanical fan assists the extraction process, but the amount of
energy saved is negligible due to the installation and maintenance
cost. The characteristics of the external surface play a very
important role in the external surface temperature that is reached.
The heat extraction mechanism in the cavity between the external
and internal surfaces denes the rate of solar heat transfer to the
internal spaces. In this mechanism, turbulence plays a funda-
mental role. The interaction between the ow and the geometry
denes the turbulence level in the channel, and this level is a
fundamental parameter for heat transfer rates. A study of how the
turbulence level can be enhanced could be interesting for
optimizing design.
For the above reasons we believe that a reliable solution must
be found in order to select the most efcient facade conguration.
If heat removal is not estimated properly, the cooling costs could
exceed the foreseen management values. DGF may represent an
attractive and efcient solution to this problem [10]. The air
channel located between the two layers of the facade can be used
to ventilate the facade by introducing outdoor or indoor air.
Multidimensional and time-dependent temperature and uid
ow elds can be computed by means of analytical computations,
numerical computations or experimental benchmarks, all of which
have contributed to a better understanding of DGF systems [9].
Due to the distinctive characteristics of the problem, simple
analytical computations involving mean values (the control
volume approach, CVA) for temperature or uid velocity do not
give results in agreement with the real case. In addition, CVA needs
experimental information that is not always available. Experi-
mental information is used extensively in DGF design, but
normally this information is vague and general. Specic experi-
mental benchmarks for a certain case may be possible, but seldom
practical from the cost point of view. Dimensional analysis could
play an important role in optimizing experimental benchmark
costs. There have been attempts to provide design correlations
with a wide validity range (see e.g. [11]), but the majority of works
deal with natural convection DGF. There are fewcomparisons with
forced convection systems. Computational uid dynamics (CFD)
modeling has some advantages compared with CVA despite this
approach requiring a large amount of computational resources.
The current computer development allows more detailed numer-
ical modeling of cases with complex geometries and highly
interactive physical phenomena. At present, a commercial CFD
code allows coupled thermo-uid-dynamics problems to be solved
for complex geometries, and plays an important role in design
Nomenclature
a absorption coefcient (m
1
)
C
p
specic heat (J kg
1
K
1
)
E absolute error
F
b
external specic body forces (kg m
2
s
2
)
F
w1
radiation shape factor
g gravity forces (m s
2
)
h heat transfer coefcient, q
w
=T
w
T
0

(W m
2
K
1
)
k kinetic energy (m
2
s
2
)
k
f
uid thermal conductivity (W m
1
K
1
)
k
w
solid thermal conductivity (W m
1
K
1
)
L characteristic length (m)
p static pressure (Pa)
q heat ux (W m
2
)
R
k
convergence ratio
t time (s)
T static temperature (K)
u velocity (m s
1
)
U uncertainty
Greek letters
e turbulent kinetic energy dissipation (m
2
s
3
)
e* relative error
z emissivity
m dynamic viscosity (kg m
1
s
1
)
r uid density (kg m
3
)
s StefanBoltzmann constant (W m
2
K
4
)
t stress tensor (N m
2
)
Dimensionless groups
Bi Biot number [hL k
w
1
]
Br Brinkman number [mu
2
k
f
1
T
1
]
Ec Eckert number [u
2
C
p
1
T
1
], [Br Pr
1
]
Eu Euler number [dpr
1
u
2
]
Fr Froude number [u
2
L
1
g
1
]
Ma Mach number [uu
1
sound
]
Nu Nusselt number [hLk
f
1
]
Pr Prandtl number [C
p
mk
f
1
]
Re Reynolds number [d
p
mrm
1
]
Sr Strouhal number [Lt
1
u
1
]
Th Thring number [rC
p
uz
1
s
1
T
3
]
Subscripts/superscripts
D referred to experimental data
f referred to a uid
G referred to the grid
I referred to iterations
rad referred to radiation
reqd referred to a programmatic validation required
level
SN referred to numerical simulation
T referred to a turbulent quantity
V referred to validation
w referred to a wall
^ referred to a dimensionless quantity
M. Coussirat et al. / Energy and Buildings 40 (2008) 17811789 1782
tasks, but a previous code verication and validation with
experimental data is necessary [11]. Therefore, CFD could be a
good choice if the modeler knows suitable sub-models for
describing the phenomena involved.
2. Theoretical considerations
A dimensionless analysis under work conditions of the set of
equations used was carried out to determine the transport
mechanisms present in the study case. Dimensional analysis is a
powerful tool that provides insight into the physics background of
the problem (see e.g., Balocco [11,12], who shows a dimensionless
analysis performed with natural and mechanical ventilated double
facades, using an experimental database). In this work, dimen-
sional analysis is performed using the general transport equations.
The dimensionless equations corresponding to mass, momentum
and energy balances (in the uid and solid phases) are as follows:
Sr
@ r
@

t
_ _
u

r r Ma
2
r

r u (1)
Sr
@ u
@

t
_ _
u

r u Eu
1
r

r p
_ _

1
Re

r n

r u

r u
T

_ _

1
Re
T

r
m
t
r

r u

r u
T

_ _

1
Fr

F
b
(2)
Sr
@

T
@

t
_ _
u

r

T Ec
1
Re
v t :

r u
1
Re
1
Pr

k
f
r

T
_ _

1
Re
1
Pr
T

k
f
T
r

T
_ _

aL
Th

z a s

T
4
(3)
r
w

C
p
w
@

T
@

t
_ _

k
w

D

T
qL
0w
k
0w
T
0w
(4)
Eqs. (1)(3) refer to the uid phase and Eq. (4) refers to the solid
phase. In all equations each term is affected by known
nondimensional parameters. The nondimensionalization process
was performed by dividing by the convective term in Eqs. (2) and
(3) and the diffusive term in (4). For this reason the Strouhal
number does not appear in Eq. (4). The right hand term in this
equation is dimensionless heat generation.
A dimensional analysis is not valid unless it takes full account of
the boundary and initial conditions, since these conditions affect
the solution of the problem directly. To do this, it is necessary to
substitute the nondimensional variables (
^
), dened by a suitable
reference value, in each of the various types of boundary
conditions. For this case, the free-stream conditions become:
u = 1, p 0;

T 0, giving no new dimensional parameters. For
non-slip solid surfaces, the conditions become: u 0;

T 1 or
Nu

k@

T=@ nj

wall
, where n is the normal direction from the
surface. Thus, if the wall heat ux is specied, a new parameter
appears: the Nusselt number (Nu). In this case it is one of the
driving parameters that affect the solution.
In Eq. (4) the only parameter is dimensionless heat generation.
The prescribed boundary conditions in this case are a radiating
boundary and a convective heat ux for both the external and
internal walls [13]. For the radiating boundary a sort of Biot
number could be dened as Bi
rad
h
rad
L
0
=k
w
, where the heat
transfer coefcient is h
rad
sF
w1
T
3
1
. For the convective boundary
condition a standard Biot number denition is obtained,
Bi h
1
L=k
w
. In this case, the initial conditions do not introduce
new nondimensional parameters.
The order of magnitude of the dimensionless groups was
estimated by taking the physical property values for air from
experimental data and empirical correlations available in the
literature [14]. This was performed for both the outer (quasi-
stagnated ow) and inner (high-speed ow) air gaps and the solid
phase present in the analyzed geometry. The Reynolds number was
calculated using two times the depth of the air gap as the
characteristic length. A Reynolds analogy has been used to
estimate values of Pr
T
from Re
T
[15]. The results obtained are
shown in Table 1. Dimensionless analysis allows the problemto be
identied as heat transfer by conduction/radiation for solid
surfaces and convection in laminar ow (for the outer gap) or
turbulent ow (for the inner gap) for the uid phase. In Eq. (2)
pressure forces and gravitational body forces in the uid phase are
considerably larger than convective acceleration and viscous and
inertial forces. Interestingly, at this point, when low-speed ow is
present, some authors write Eq. (2) slightly differently by using the
thermal expansion coefcient in order to account for the density
changes due to temperature changes. The quotient between the
Grashof and Reynolds numbers (Gr/Re
2
) is identied by several
algebraic operations involving the grouping of the pressure
gradients and an operation involving gravity-related terms. This
quotient is a kind of Froude number (Fr), used to determine the
importance of buoyancy (for further details please refer to [15]).
In Eq. (3), convection is an important transport mechanism.
Turbulent mixing clearly becomes an important transport
mechanism in the inner gap, while in the outer gap the laminar
and turbulent transport contributions are in the same order of
magnitude. For both cases, the contribution of viscous dissipation
is negligible. The radiation term vanishes independently of the
value of the Thring number (Th) because the absorption coefcient
(a) for a transparent gas (suchas air in this case) is zero. The Nusselt
number (Nu) is an important parameter in the boundary condition
and therefore the convective heat transfer coefcients must be
estimated carefully.
In the solid phase the radiative Biot number (Bi
rad
) is
signicantly greater than the standard Biot number (Bi), which
means that the importance of radiation phenomena in the heat
transfer mechanism present in these surfaces is greater than the
contribution of the conduction and convection phenomena. This
requires a careful computation of the ow that goes into the
computational domain from the outer wall. Modeling this
boundary condition by means of a suitable radiation model
becomes of paramount importance.
Inside the ventilation cavity, accurate modeling of the ow
mixing in this case implies using a turbulence model capable of
accounting for low Reynolds number effects, as it is impossible to
apply both a laminar and a turbulence model to the same
Table 1
Dimensionless groups orders of magnitude
Outer gap Inner gap Solid phase
Bi 10
1
10
2

Bi
rad
10
1
Ec 10
9
10
7

Eu 10
3
10
1

Fr 10
3
10
2

Ma 10
5
10
4

Nu 10
0
10
0
Pr 10
1
10
1

Pt
T
10
1
10
5

Re 10
2
10
3

Re
T
10
5
10
1

Th 10
1
10
2

Sr 10
1
10
2

M. Coussirat et al. / Energy and Buildings 40 (2008) 17811789 1783


computational model. One of the intended objectives of this work
was to check whether it is possible to obtain an accurate model
using an eddy viscosity model that is able to take into account
these lowReynolds number effects together with buoyancy effects
without overestimating the kinetic energy (e.g., for stagnation
point anomalies) or the dissipation rate (e.g. for predicting
turbulent behavior at low ow velocities).
Accurate modeling of heat transfer in the solid phase implies
using a radiation model capable of accounting for the energy
absorption/emission/transmission in the semi-transparent sur-
faces of the DGF as well as predicting the solid phase temperatures
properly.
3. Modeled case
A well-documented experimental database from Manz et al.
[16] was selected to validate and calibrate the CFD code for DGF
problems. In this study, experimental data (surface and air
temperatures, and solar irradiation) were derived for a single-
story glass DGF with free and forced convection heat transfer (see
Fig. 1). The uncertainties for the experimental data are considered
suitable for this work (3% for the solar irradiation measurement
and 0.5 K for the temperature measurements) [16].
3.1. Geometrical model
A vertical section of the DGF studied numerically is shown in
Fig. 1. This comprises (fromoutside to inside) an insulating glazing
unit whose outer pane has solar protection properties and a low-
emissivity coating, an outer air gap, a metallized shading screen, an
inner air gap and a single pane with a low-emissivity coating on the
outer side. Air can ow between the gaps (each with a width of
0.02 m) through two openings at the top and bottom of the facade.
Table 2 gives the thickness and solar and infrared properties of all
layers. The dimensions of the geometrical model were taken
following the indications of the experimental setup developed by
Manz et al. [16].
3.2. Applied numerical models
The importance of turbulence and radiation in DGF problems
has been highlighted above, and sub-models that allow these
inuences to be detected are necessary in CFD simulation.
Some guidelines concerning the applied sub-models will now
be presented.
3.2.1. Radiation models
Heating or cooling of surfaces due to radiation and/or heat
sources or sinks due to radiation within the uid phase can be
computed by means of a radiation model. For particular problems,
one radiation model may be more appropriate than others
depending on conditions such as the optical thickness of the
problem, the presence of scattering effects, the emissivity and the
nature of the surfaces (opaque, semi-transparent, specular, etc.),
the uid involved and the presence of localized heat sources. For
this case it has been demonstrated that it is important to nd a
suitable radiation model that is able to reproduce properly the
physical phenomena present in our specic analyzed case. The
radiation models used in this work can be seen in Table 3.
3.2.2. Turbulence models
It is of paramount importance to determine the most adequate
turbulence model for computing mean and local values of
velocities and temperature elds when one is modeling the
DGF. The system of equations is dened by averaged continuity,
momentum and energy equations. Fluctuations of these mean
quantities are calculatedby means of a turbulence model toobtain
representative velocity and thermal elds. As most of the practical
interest is to determine the mean quantities, this technique is
used widely in modeling turbulent ows. An extensive list of
references with a description of these equations can be found in
the publishedliterature (see e.g. [23]). The researchonthis subject
has led to the development of several classes of turbulence
models. The classicationscheme showninTable 4 only takes into
account the models usedinthis work, whichare basedonthe eddy
viscosity concept.
Fig. 1. Vertical section of the DGF showing the solution domain (dimensions in
meters).
Table 2
Layer properties of the facade
Layer number Name Thickness
(mm)
Emissivity outside
(dimensionless)
Emissivity inside
(dimensionless)
Solar transmittance
(dimensionless)
Solar absorptance
(dimensionless)
1 Solar protective, low-emissivity glass 10 0.89 0.05 0.39 0.33
2 Air 18 -
3 Glass 14 0.89 0.89 0.78 0.03
4 Air 70 -
5 Metallized shading screen 1 0.51 0.81 0.09 0.17
6 Air 113
7 Low-emissivity glass 6 0.22 0.86 0.004 0.68
Optical data taken from Manz et al. [16].
M. Coussirat et al. / Energy and Buildings 40 (2008) 17811789 1784
The standard formulation of these models is only valid for
turbulent ows that are far fromwalls, where terms corresponding
to shear stress are small and the turbulent viscosity becomes
isotropic. Walls are the main source of vorticity and turbulence
because there are large gradients of variables and a more vigorous
momentum transport in their vicinity. To take into account the
non-isotropic nature of turbulence in near-wall regions, a two-
layer modeling approach (for the k e family models) and an
enhanced wall function approach (for the SpalartAllmaras or the
k v family models) have been implemented together with the
tested turbulence model.
3.3. Mesh design and boundary conditions
In addition to the aforementioned, appropriate modeling of the
turbulence phenomena involved in this case implies using a highly
rened homogeneous mesh, but this idea involves high computa-
tional resources. Another of the intended purposes of this work
was to check whether it is possible to obtain an accurate model
when the idea of using a homogeneous mesh is relaxed in order to
minimize simulation times. According to this, the mesh should be
designed to properly dene a minimum cell size to compute the
turbulent mixing appropriately with the proposed geometry. Mesh
density depends on the near-wall modeling strategy adopted for
resolving the problem under turbulent ow conditions, and is
determined by the y
+
characteristic parameter [31].
The overall dimensions of the solution domain were
0.5 m 2.15 m (Fig. 1) with a differential thickness of 10 mm.
Although the bi-dimensional nature of the ow and transport
phenomena present in geometries of this type is widely known, a
three-dimensional model was built in order to perform suitable
turbulence modeling as turbulence is a 3D phenomena. It also took
advantage of specic features of the CFD solver used (Fluent
1
6.3,
[31]), such as solar load radiation models, which are only available
for three-dimensional geometries. Due to the regularity of the
volumes involved, a homogeneous hexahedral grid was used to
build the computational mesh in all cases.
Initial and boundary conditions were set as close as possible to
the experimental data available [16]. Horizontal boundaries and
the boundaries of the exhaust duct were adiabatic and the mass
ow was set to a constant value (0.1310 kg/m
2
s) at the exhaust in
the models, which was provided by a fan in the experimental
setup. Outdoor and indoor boundary conditions were chosen to be
open for heat ow and if not obstructed by solid parts also for
mass ow.
Thermal boundary conditions for the exterior and interior
layers were imposed as convective + radiative heat uxes and
convective heat ux, respectively. A value of 12 W/m
2
K for the
heat transfer coefcients used in the convective heat ux
conditions was set at the outdoor airlayer 1 boundary. This
value refers to the experimental values of the outdoor air
temperature measured and reported for the specic date (8th
September) and the physic location of the experimental setup [16].
The boundary condition for heat ux over the DGF due to radiation
was computed by means of a radiation model, which computes a
value for the radiative heat contribution and then imposes it on the
energy balance computation for the solid phases through a
fountain term or a surface heat ux (depending on the radiation
model selected). A value of 8 W/m
2
K for the convective heat
transfer coefcient was set at the indoor airlayer 7 boundary
[32], in reference to the indoor desired temperature (296 K).
3.4. Modeling work
NavierStokes equations together with energy balance were
solved using the Fluent
1
6.3 commercially available nite volume
Table 3
Radiation models used for this work
Model Discrete transfer radiation model (DTRM) Discrete ordinates (DO) P-1 Radiation model
Reference [17,18] [19,20] [21,22]
Features The main assumption of this model is
that the radiation leaving the surface
element in a certain range of solid angles
can be approximated by a single ray
This model transforms the radiative transfer
equation into a transport equation for radiation
intensity in the spatial coordinates, solving as
many transport equations as there are direction
vectors associated with a number of discrete
solid angles. The solution method is identical to
that used for the uid ow and energy equations
The P-1 radiation model is the simplest case of the
more general PN model, which is based on the
expansion of the radiation intensity into an
orthogonal series of spherical harmonics, in order
to solve the radiative transfer equation
Table 4
Turbulence models used for this work
Model One-equation model Two-equation models
SpalartAllmaras (SA) Standard k e (SKE) RNG k e (RNGKE) Realizable k e (RKE) Standard k v (SKW) SST k v (SSTKW)
Reference [24] [25] [26] [27] [28,29] [30]
Features Uses a differential
partial equation for
the transport of
turbulent viscosity.
In this way the
turbulent quantity
modeled is directly
the effective viscosity
Uses one differential
equation for the
turbulent velocity scale
and another for the
turbulent length scale.
The variables modeled
are the turbulent kinetic
energy, k, and the rate
of dissipation of turbulent
kinetic energy, e Then
the eddy viscosity is
computed and turbulence
effects are introduced in
the equations for
mean quantities
As the SKE model,
but using the
renormalization
group (RNG) methods.
This results in a model
with constants different
from those in the SKE
model, and additional
terms in the obtained
transport equations
As the SKE model,
but the term
realizable means
that the model satises
certain mathematical
constraints on the
normal stresses,
consistent with the
physics of turbulent
ows. It adopts a new
eddy-viscosity formula
and a new model
equation for dissipation
This model is based
on model transport
equations for the
turbulence kinetic
energy (k) and the
specic dissipation
rate (v), which can
also be thought of as
the ratio of e to k.
Production terms
have been added to
the model equations
The major way in
which the SST model
differs from the
standard model is a
modied turbulent
viscosity formulation
to account for the
transport effects of
the principal turbulent
shear stress
M. Coussirat et al. / Energy and Buildings 40 (2008) 17811789 1785
code software [31]. A pressure-based double-precision solver was
selected in order to solve the set of equations used. A structured
hexahedral grid was applied to the geometrical model. Second
order upwind discretization schemes were imposed on all the
transport equations solved (momentum, energy and turbulence).
PISO pressurevelocity coupling was chosen due to its suitability
for buoyancy-affected ows [31]. The uid was taken to be
incompressible, Newtonian and in a turbulent owregime. Air was
chosen as the simulation uid, for which the constants were
available in the software database. The incompressible ideal gas
law for density and power law for viscosity were applied to the
model to make these variables temperature-dependent. Simula-
tions were run in transient mode, making a 24-h analysis of the
behavior of the double facade. A time step of 1 h was taken as the
typical time step in building simulation programs [16].
Simulations were run on an HP Proliant DL385 workstation, and
simulation times ranged from 1 to 12 h depending on the case
studied. Numerical convergence of the model was checked based
on the normalized numerical residuals of all computed variables.
For a more complete convergence check, the static temperature at
different points of the geometry was also chosen as a monitor.
4. Results and discussion
4.1. Buoyancy-driven ow and ow patterns
In the uid phase, due to the presence of two differentiated ow
zones within the computational domain (a high-speed owzone in
the inner gap, and a quasi-stagnated ow zone in the outer gap),
free and forced convection effects in the ow patterns and heat
transfer were expected. Although the dimensionless analysis
revealed that gravitational body forces were relevant within the
momentum balance, a simulation series was run in order to
quantify the importance of the free convection effects for
estimating heat uxes. In this series, the air temperature at the
DGF exhaust was recorded and compared with experimental
results for a owmodel with and without considering gravitational
body forces for an elapsed time of 24 h [16]. The results obtained
can be seen in Fig. 2. The numerical results showed a better t to
experimental results when gravitational body forces (and hence
free convection effects) were considered within the ow model.
For subsequent simulations, gravitational body forces were
activated within the momentum equation.
The inuence of gravitational body forces on the velocity eld
within the DGF can be observed when comparing the velocity
contour plots near the exhaust at different times of the day. Fig. 3
shows the evolution of the velocity eld throughout the day. It can
be seen that in the presence of solar radiation (6, 12 and 18 h) free
convection effects are shown by the formation of a high-speed air
stream moving upwards around the shading screen, while in the
absence of solar radiation (24 h) this stream disappears.
4.2. Radiation model
In order to obtain accurate results for the temperature proles
in the solid phase, it was necessary to evaluate the performance
of different radiation models available in the CFD solver used
for simulations. A simulation set was run to test three
different radiation models (see Table 3). The RNGKE turbulence
model was selected for the radiation test cases because it
includes an analytically derived differential formula for effective
viscosity that accounts for low Reynolds number effects. In this
Fig. 2. Air temperature at the exhaust of the double-glazing for a ow model with
and without considering gravitational body forces.
Fig. 3. Velocity contour plots near the exhaust of the DGF at different daytimes (velocity in m/s).
M. Coussirat et al. / Energy and Buildings 40 (2008) 17811789 1786
series, the temperature in a solid phase (in this case, the
center point of layer 5, the metallized shading screen) through-
out the day was recorded for each radiation model tested and
compared with experimental results [16]. The results can be seen
in Fig. 4. The numerical results showed a better t to the
experimental results when the P-1 radiation model was used.
Due to this good t, the P-1 radiation model was used for
subsequent simulations.
4.3. Turbulence model
The performance of several turbulence models for predicting
the heat transfer phenomena involved in a DGF was tested. The
existence of two differentiated ow zones implies that it is
appropriate for an accurate model of the owmixing in this case to
use a turbulence model capable of accounting for low Reynolds
number effects, as it is impossible to apply both a laminar and a
turbulence model to the same computational model. For this
reason, a simulation series testing ve different turbulence models
(see Table 4) was run. In this series, the evolution of air
temperature at the exhaust of the DGF throughout the day for
each turbulence model was recorded and compared with
experimental results [16]. The results can be seen in Fig. 5. In a
rst approach, the numerical results showed a better t to the
experimental results when the SKE and RNGKE turbulence models
were used.
4.3.1. Mesh sensitivity of the numerical results
A mesh sensitivity test was carried out to prove mesh-
independence of the numerical results. Four simulation series
were run for each of the turbulence models tested, using a
sequentially rened homogeneous mesh for each set (where each
homogeneous mesh has twice the number of elements of the
previous mesh in the sequence, e.g., 16, 32, 64 and 128 thousand
control volumes). Air temperature at the DGF exhaust was again
taken as the control variable.
Fig. 4. Performance of different radiation models available in the CFD solver.
Fig. 5. Performance of turbulence models tested.
Fig. 6. Numerical results obtained for different mesh densities using RNGKE.
Fig. 7. Absolute error for the mesh-independency simulation series.
M. Coussirat et al. / Energy and Buildings 40 (2008) 17811789 1787
For all cases a clear dependence of the numerical results on the
mesh density was observed. An example of this can be seen in
Fig. 6, where numerical results for the four mesh densities tested
with the RNGKE turbulence model are compared with experi-
mental results [16]. It was also observed that the error was greater
in the hours of greater radiation intensity, which means that the
error in predicting uid temperatures is related to the miscalcula-
tion of the free convection effects in the DGF, which increase in the
hours of more solar intensity.
Following the reference pattern mentioned above, the absolute
error (see Fig. 7) and the relative errors (see e.g. Fig. 8, which shows
therelativeerrors obtainedat thetimeof highest radiationintensity)
of the numerical results obtained were calculated and analyzed. For
all turbulence models tested except for RKE a monotonic
convergence (error decreases as the mesh density increases) was
observed. As can be seen in the gures, an oscillatory convergence
was obtained with RKE. It was also observed that for every
simulation series studied, SKE and RNGKE showed the smallest
error in all cases, independently of the mesh density used.
4.3.2. Verication and validation of the results obtained
The verication and validation methodology and procedures of
CFD simulations provide a pragmatic approach for estimating
simulation errors and uncertainties [33]. The present analysis
allowed us to treat the simulation errors as either stochastic or
deterministic and takes into account the uncertainties in both the
simulation and the data for assessing the level of validation.
Several CFD verication and validation procedures are avail-
able in the literature [3337]. Wilson et al. [34] state that
verication is accomplished by parameter convergence studies
using multiple solutions (at least 3) withsystematic meshrening
while keeping all other parameters constant. Following the
procedure referenced above, it was possible for the simulation
sets with monotonic convergence to estimate the validation
uncertainty (U
V
), dened as the combination of all uncertainties
that we know how to estimate (i.e. the grid uncertainty (U
G
), the
iteration uncertainty (U
I
), the simulation numerical uncertainty
(U
SN
) and the uncertainty in the experimental data (U
D
)).
According to these authors [29], validation is accomplished by
comparing simulations with benchmark CFD data, including
experimental uncertainty estimates (U
D
). If three variables U
V
, e*
and U
reqd
(programmatic validation required) are considered,
there are six combinations. For three cases, e* < U
V
and validation
is achieved at the U
V
level, but U
V
< U
reqd
for only one of these, so
validation is also achieved at U
reqd
.
For the turbulence model benchmark described in the mesh
sensitivity tests, by taking, for example, the air temperature at the
DGF exhaust at a given time (e.g. 17 h) as the control variable for
the validation studies, convergence ratios (R
k
) can be estimated
(see Table 5) and uncertainties can be obtained for those situations
in which a monotonic convergence is obtained (0 < R
k
< 1). The
uncertainties and relative errors obtained can be seen in Table 6.
According to the above criteria, setting U
reqd
= 1% as our required
validation level, although for many combinations of grid density
and turbulence models validation is achieved at the U
V
level, only
RNGKE achieves validation at the U
reqd
level.
5. Conclusions
CFDproves tobe a useful tool for modeling owandheat transfer
of DGF including conduction, convection and radiation heat transfer
phenomena. It was possible to reproduce an experimental case
numerically and to obtain velocity and temperature proles
Fig. 8. Relative errors obtained at the time of highest radiation intensity.
Table 5
Values obtained for the control variable and convergence ratios for the turbulence model benchmark test
Turbulence model Control variable (T
exhaust, 17 h
(K)) Convergence ratio, R
k
Grid 4, 16 kel Grid 3, 32 kel Grid 2, 64 kel Grid 1, 128 kel Study 1, Grid 4/3/2 Study 2, Grid 3/2/1
SA 307.99 310.15 312.07 313.21 0.89 0.60
SKE 310.84 312.52 314.13 315.58 0.96 0.89
RNGKE 309.93 312.18 314.32 315.64 0.96 0.62
RKE 309.66 312.05 310.05 311.37 0.83 0.65
SKW 310.27 311.89 313.53 314.55 1.01 0.62
SSTKW 309.43 311.74 313.52 314.99 0.76 0.83
Table 6
Relative errors and validation uncertainties for the turbulence models benchmark test
Turbulence model Relative error (against control variable) Validation uncer-
tainty, U
V
U
G
U
SN
U
D
U
I
Grid 4, 16 kel Grid 3, 32 kel Grid 2, 64 kel Grid 1, 128 kel Study 1 Study 2 Study 1 Study 2 Study 1 Study 2
SA 2.88 2.20 1.59 1.23 4.85 0.55 4.85 0.53 4.85 0.53 0.16 0.05
SKE 1.98 1.45 0.94 0.48 11.53 4.11 11.53 4.11 11.53 4.11 0.16 0.05
RNGKE 2.27 1.56 0.88 0.47 14.98 0.68 14.98 0.67 14.98 0.67 0.16 0.05
SKW 1.65 1.13 0.81 0.55 0.52 0.53 0.16 0.05
SSTKW 2.42 1.69 1.13 0.81 1.82 2.32 1.82 2.31 1.82 2.32 0.15 0.05
All values expressed as a percentage of the experimental value of the control variable.
M. Coussirat et al. / Energy and Buildings 40 (2008) 17811789 1788
within the DGF. The results were compared with experimental data
and validated according to numerical verication and validation
procedures.
The inuence of gravitational body forces on the velocity eld
within the DGF was observed when the velocity elds were
compared within the facade. It was seen that in the presence of
solar radiation, free convection effects are shown by the formation
of a high-speed air stream moving upwards around the hot
surfaces, while in the absence of solar radiation this free
convection stream disappears.
The performance of several turbulence models for predicting
the heat transfer phenomena involved in a DGF was tested. The
existence of two differentiated ow zones implies that accurate
modeling of the ow mixing in this case involves using a
turbulence model capable of accounting for low Reynolds number
effects, as it is impossible to apply both a laminar and a turbulence
model in the same computational model. It was observed that for
every simulation series studied, SKE and RNGKE showed the
smallest error in all cases, independently of the mesh density used.
For the turbulence model benchmark, convergence ratios were
estimated and validation uncertainties were obtained for those
turbulence model/grid density combinations in whicha monotonic
convergence was obtained. Following the verication and valida-
tion criteria stated by Wilson et al. [33], and setting U
reqd
= 1% as
our required validation level, it was observed that although for
many combinations of grid density and turbulence models
validation is achieved at the U
V
level, only RNGKE achieved
validation at the U
reqd
level.
Acknowledgments
The authors acknowledge the economical support received
from Grupo JG Consultora de Proyectos S.A. for developing this
work.
References
[1] M. Nicoletti, Architectural expression and low energy design, Renewable Energy
15 (1998) 3241.
[2] J. Hensen, M. Bartak, D. Frantisek, Modeling and simulation of a double-skin
facade system, ASHRAE Transactions 108 (2) (2002) 12511259.
[3] E.V. Fachinstitute Geba ude-Klima (Ed.), Proceedings of Doppelfassaden und
technische geba udeausru stung, Bietigheim-Bissingen, Germany, 1997.
[4] W. Hien, W. Liping, A. Chandra, A. Pandey, W. Xiaoling, Effects of double glazed
facade on energy consumption, thermal comfort and condensation for a typical
ofce building in Singapore, Energy and Buildings 37 (2005) 563572.
[5] K. Gertis, Sind neuere fassadenentwicklungen bauphysikalisch sinnvoll? Teil 2:
glas-doppelfassaden (GDF), Bauphysik 2 (1999) 5466.
[6] A. Viljoen, J. Dubiel, M. Wilson, M. Fontoynont, Investigation for improving the
daylighting potential of double-skinned ofce buildings, Solar Energy 59 (1997)
179194.
[7] D. Faggembauu, M. Costa, M. Soria, A. Oliva, Numerical analysis of the thermal
behaviour of ventilated glazed facades in Mediterranean climates. Part I. Devel-
opment and validation of a numerical model, Solar Energy 75 (2003) 217228.
[8] D. Faggembauu, M. Costa, M. Soria, A. Oliva, Numerical analysis of the thermal
behaviour of ventilated glazed facades in Mediterranean climates. Part II. Appli-
cations and analysis of results, Solar Energy 75 (2003) 229239.
[9] W.N. Hien, W. Liping, A.N. Chandra, A.R. Pandey, W. Xiaolin, Effects of
double glazed facade on energy consumption, thermal comfort and condensa-
tion for a typical ofce building in Singapore, Energy and Buildings 37 (2005)
563572.
[10] D. Faggembauu, M. Costa, M. Soria, A. Oliva, Strategies to reduce thermal over-
heatings in Mediterranean climates using large glazed areas, in: FIER Forum
International sur les Energies Renovables, 2, Te touan, Marroc, (2002), pp. 307
312.
[11] C. Balocco, A non-dimensional analysis of a ventilated double facade energy
performance, Energy and Buildings 36 (2004) 3540.
[12] C. Balocco, M. Colombari, Thermal behavior of interactive mechanically ventilated
double glazed facade: non-dimensional analysis, Energy and Buildings 38 (2006)
17.
[13] F. White, Heat and Mass Transfer, Addison-Wesley Publishing Co., 1991.
[14] R. Reid, J. Prausnitz, B. Poling, The Properties of Gases and Liquids, McGraw-Hill,
1987.
[15] F. White, Viscous Fluid Flow, McGraw-Hill, 1991.
[16] H. Manz, A. Schaelin, H. Simmler, Airow patterns and thermal behavior of
mechanically ventilated glass double facades, Building and Environment 39
(2004) 10231033.
[17] N.G. Shah, A new method of computation of radiant heat transfer in combustion
chambers, Ph.D. Thesis, Imperial College of Science and Technology, 1979.
[18] M.G. Carvalho, T. Farias, P. Fontes, Predicting radiative heat transfer in absorbing,
emitting, and scattering media using the discrete transfer method, in: W.A.
Fiveland et al., (Eds.), Fundamentals of Radiation Heat Transfer, vol. 160, ASME
HDT, 1991, pp. 1726.
[19] G.D. Raithby, E.H. Chui, A nite-volume method for predicting a radiant heat
transfer in enclosures with participating media, Journal of Heat Transfer 112
(1990) 415423.
[20] E.H. Chui, G.D. Raithby, Computation of radiant heat transfer on a non-orthogonal
mesh using the nite-volume method, Numerical Heat Transfer B 23 (1993) 269
288.
[21] P. Cheng, Two-dimensional radiating gas owby a moment method, AIAA Journal
2 (1964) 16621664.
[22] R. Siegel, J.R. Howell, Thermal Radiation Heat Transfer, Hemisphere Publishing
Corporation, Washington, DC, 1992.
[23] P.A. Durbin, B.A. Petterson Reif, Statistical Theory and Modeling for Turbulent
Flows, Wiley, 2001.
[24] P. Spalart, S. Allmaras, A one-equation turbulence model for aerodynamic ows,
American Institute of Aeronautics and Astronautics, Technical report AIAA-92-
0439, 1992.
[25] B.E. Launder, D.B. Spalding, The numerical computation of turbulent ows,
Computer Methods in Applied Mechanics and Engineering 3 (1974) 269289.
[26] D. Choudhury, S.-E. Kim, W.S. Flannery, Calculation of turbulent separated ows
using a renormalization group based k e turbulence model, vol. 149, American
Society of Mechanical Engineers, Fluids Engineering Division (publication) FED,
1993, pp. 177187.
[27] T.-H. Shij, W.W. Liou, A. Shabbir, Z. Yang, J. Shu, New k e eddy viscosity model
for high Reynolds number turbulent ows, Computers and Fluids 24 (1995) 227
238.
[28] D.C. Wilcox, Reassessment of the scale-determining equation for advanced
turbulence models, AIAA Journal 26 (1998) 12991310.
[29] D.C. Wilcox, Multiscale model for turbulent ows, AIAA Journal 26 (1998) 1311
1320.
[30] F.R. Menter, Two-equation eddy-viscosity turbulence models for engineering
applications, AIAA Journal 32 (1994) 15981605.
[31] Fluent Inc., Fluent 6.3. Users guide, 2006.
[32] I. Pe rez-Grande, J. Meseguer, G. Alonso, Inuence of glass properties on the
performance of double-glazed facades, Applied Thermal Engineering 25 (2005)
31633175.
[33] F. Stern, R.V. Wilson, H.W. Coleman, E.G. Paterson, Comprehensive approach to
verication and validation of CFD simulations. Part 1. Methodology and proce-
dures, Journal of Fluids Engineering 123 (2001) 793802.
[34] R.V. Wilson, F. Stern, H.W. Coleman, E.G. Paterson, Comprehensive approach to
verication and validation of CFD simulations. Part 2. Application for RANS
simulation of a cargo/container ship, Journal of Fluid Engineering 123 (2001)
803810.
[35] H.W. Coleman, F. Stern, Uncertainties, CFD code validation, Journal of Fluids
Engineering 119 (1997) 795803.
[36] P.J. Roache, Verication of codes and calculation, AIAAJournal 36 (1998) 696702.
[37] J. Cadafalch, C.D. Pe rez-Segarra, R. Co nsul, A. Oliva, Verication of nite volume
computations on steady-state uid ow and heat transfer, Journal of Fluids
Engineering 124 (2002) 1121.
M. Coussirat et al. / Energy and Buildings 40 (2008) 17811789 1789

Vous aimerez peut-être aussi