You are on page 1of 15

Case Studies in Construction Materials 11 (2019) e00249

Contents lists available at ScienceDirect

Case Studies in Construction Materials


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

Three-dimensional numerical simulation of conduction,


natural convection, and radiation through alveolar
building walls
Mohamed Ouakarroucha , Karima El Azharya , Najma Laaroussia,* ,
Mohammed Garouma , Amir Feizb
a
University Mohammed V in Rabat, Materials, Energy and Acoustics Team (MEAT), avenue Prince Héritier, B.P: 227 Salé Médina, Morocco
b
University d’Evry Val-d’Essonne, 40 Rue du Pelvoux, 91020 Evry, France

A R T I C L E I N F O A B S T R A C T

Article history: Numerical simulation of the coupled heat transfers by conduction, convection, and
Received 7 February 2019 radiation through two kinds of the alveolar structures used in the construction are
Received in revised form 10 May 2019 numerically investigated. Owing to the low-temperature differences involved, the
Accepted 14 May 2019
tridimensional model is based on the Boussinesq approximation and constant
thermophysical fluid properties at mean temperature. The alveolar walls are assumed
Keywords: grey and diffuse. The symmetric and periodic boundary conditions for flows governed by
Alveolar
the incompressible Navier-Stokes equations are considered. Equations governing the
Natural convection
Surface radiation
natural convection in the alveolars, the radiative heat exchanges between their internal
Combined heat transfer surfaces and the heat conduction in the surrounding walls are solved using a finite volume
Thermal resistance method. The pressure-velocity coupling is solved by SIMPLE algorithm. The objective of this
study is to improve the effects of wall conduction and radiation heat exchange between
surfaces on the decrease of the thermal resistance in the alveolar walls. Two kinds of
concrete blocks with three and six-hole numbers are chosen and their thermal resistances
are numerically predicted. The comparison between the thermal resistances, specified in
terms of thermal regulations for buildings and numerical solutions, is presented.
© 2019 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY
license (http://creativecommons.org/licenses/by/4.0/).

* Corresponding author.
E-mail address: najma.laaroussi@um5.ac.ma (N. Laaroussi).

https://doi.org/10.1016/j.cscm.2019.e00249
2214-5095/© 2019 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/
4.0/).
2 M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249

Nomenclature

α Thermal diffusivity (m2s1)


A Enclosure aspect ratio, A=H/L
B Dimensionless radiosity
cp Specific heat (J. K1. kg1)
g Gravitational acceleration (m.s2)
hcv Convective heat transfer coefficient (W. m–2. K–1)
H Cavity height (m)
k Thermal conductivity (W. m1. K1)
L Length of cavity (m)
Li Length of internal cavity i (m)
P Pressure (Pa)
P* Dimensionless motion pressure
Pr Prandtl number, Pr=n0/α0
Pl Planck number, Pl = k0 DT/ (s T 4c L)
qr;i Net dimensionless radiative heat flux along
surface"i", qr,i/(s T 40 )
Qt* Dimensionless total heat flux, Qt*=Qr*+ Qc*
Qc* Dimensionless convective heat flux, - L@@Tx /DT
Qr* Dimensionless radiative heat flux, L.Qr/(k.DT)
RaL Thermal Rayleigh number, RaL = r0 g βTDT.L3/
(m0α0)
Raeff Thermal Rayleigh number, Raeff = r0 gβTDTi.Li3/
(m0α0)
T Temperature (K)
T0 Reference temperature T0=(Tc+Th)/2 (K)
(x*,y*,z*) Dimensionless cartesian coordinates.
V* Dimensionless velocity

Greek symbols
r Density (kg. m3)
β Stretching parameter.
βT Thermal coefficient of volumetric expansion (K1)
e Emissivity
DT Temperature difference, (Th-Tc) (K)
DTi Temperature difference, (Tmh,i-Tmc,i) (K)
s Stefan-Boltzmann constant, s = 5.672  108 (W m-2 K-4)
u Dimensionless temperature difference, (T-T0)/DT
u0 Dimensionless reference temperature ratio, Tc / DT

Subscripts
f Fluid
s Solid

1. Introduction

Research on the analysis and numerical resolution of coupling between the radiative transfer and the Navier-Stokes
equations in irregular geometries becomes the essential element for the employment of CFD simulations into industry-
related problems. This is due to the importance of this kind of geometry in many industrial applications such as thermal
insulation of buildings, solar energy collectors, fire control in buildings. In particular, the hollow structure has received
considerable attention because of its applicability in any engineering applications such as masonry walls in improving the
thermal and sound insulation performance of building walls and fire protection.
From the mid-Seventies, the increasing of computer performance has allowed to start treating three-dimensional
geometries. The published results by Mallinson and de Vahl Davis [1] established the first numerical studies of the heat
transfers by natural convection in parallelepiped cavities limited by four adiabatic and two vertical parallel walls maintained
M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249 3

at different temperatures. The three-dimensional effects were small except in the neighborhood of the end vertical walls and
in the junction zones between the vertical walls.
Heat transfer through hollow structures is taking place simultaneously with combined processes (convection, radiation,
and conduction) within the hollow blocks. Castro et al. [2] studied a 2D numerical simulation of conjugate heat transfer in a
structure with twelve rectangular holes. The horizontal top and bottom surfaces are considered adiabatic and the side walls
are heated with different temperatures. The effect of the hole surface radiation is performed. The combined heat transfer by
conduction, natural convection and surface radiation through building walls consisting of alveolar structures was studied
numerically by Abdelbaki and Zrikem [3].
Later, a numerical study that considers both transfers conductive, convective and radiation in hollow structures heated
from below or above was conducted by Ait-Taleb et al. [4]. The presented results were for structures which formed by a range
of three identical rectangular cavities. The heat transfer through the hollow structures was affected by the cavity’s aspect
ratio and the temperature difference between the horizontal side walls.
The resistance to fire of masonry walls constituted by hollow blocks was investigated from both experimental and
theoretical points of view by Nahhas et al. [5]. The theoretical model was developed to describe the experimental results
considering convection, conduction and radiation phenomena inside the walls. The results showed that the theoretical
model estimates correctly temperatures.
Coupled heat transfer through double hollow brick walls including an air layer was performed numerically by Boukendil
et al. [6]. The influence of mortar thickness related to a low emissivity of the structure surfaces was evaluated to reduce the
energy consumption of buildings. The thermal optimization and performance analysis on different types of light concrete
hollow bricks, using the finite element method, were established by del Coz Díazet al. [7,8].
Recently, a series of numerical studies, that consider both transfers conductive, convective and radiation in three-
dimensional geometry, are performed for cubical cavities [9–11].
Three-dimensional convection and radiation in a cubic cavity filled with transparent and participating media were
numerically analyzed in [12] with the use of the discrete ordinate method for solving the radiation transfer equation. It was
shown that the heat flux density grows in the case of a radiation-transmitting medium. For a participating medium with a
fixed value of optical medium thickness, it was found that an increase in the Rayleigh number leads to an increase in the heat
flux density.
The effect of surface radiation and aspect ratio on the 3D regimes in a differentially heated parallelepiped was analyzed
numerically in [13] when Prandtl and Rayleigh's numbers were fixed at Pr = 13.6 and Ra = 106. They showed that the flow
structure was considerably modified with the presence of relative intense spiral motion in the volume. The thermal
optimizing of holes arrangement of the hollow clay brick was carried out in [14]. They sized them for by 240  115  90 mm3
by calculating their equivalent thermal conductivity using 3D numerical simulation. Afterward, the authors conducted 3D
numerical simulation with finite volume method to find the optimum configuration of the number of holes and their
arrangement for 290  140  90 mm3 hollow clay brick [15].
Lorente et al. [16,17] paid attention to walls constructed with vertical hollow bricks. They proposed a 2D simplified
analytical model to calculate heat flux and thermal resistance of single vertical hollow brick in [16] and they also determined
the thermal resistance for different shapes of vertical hollow bricks using 2D conjugate simulation of heat transfer in [17].
Lorente and Bejan [18] formulated the problem of optimizing the internal structure of a vertical composite wall brick, by
considering the thermal insulation with a fixed mechanical strength.
Zhang et al. [19] combined the experiment and calculation in order to analyze the influence of the thermal performance of
a hollow block wall by modifying boundary condition, thermophysical properties and configuration of the hollow block
itself.
Antar and Baig [20] investigated numerically the conjugate heat transfer across a hollow block with a suitable count of
cavities in order to obtain an increase in thermal resistance of the wall. The results showed that increasing the number of
cavities while keeping the block width fixed, decreases the heat loss.
Khatamifar et al. [21] conducted a numerical study of the conjugate natural convection and heat transfer in a differentially
heated cavity. The effect of partition thickness and the Rayleigh number on thermal behavior was examined. They found that
the average Nusselt number increases with the Rayleigh number but decreases with partition thickness. Yang and Wu [22]
were carried out the effects of natural convection, thermal radiation and wall thermal conduction on heat flux distribution at
a heated plate located at the bottom of a three-dimensional rectangular enclosure. The distributions of convective, radiative
and Nusselt number at the surface of the heat source were obtained. The effect of convective heat transfer was reduced by
thermal radiation, but the total heat transfer was increased.
This work is aimed at investigating numerically the effects of wall conduction and radiation heat exchange among
surfaces on laminar natural convection heat transfer in a three-dimensional, differentially heated cavity modeling a concrete
hollow block. For this purpose, we consider two kinds of concrete alveolar building walls described in Th-U [23] : the
concrete block 10 of dimension 10  20  50 cm3 and the concrete block 20 of dimension 20  20  50 cm3 whose the
thermal resistances, taking into account the designated and specifications from Th-U [23], have to be equal to Rth = 0.12 m2.K.
W1 and Rth = 0.23 m2.K.W1 respectively.
4 M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249

2. Mathematical modeling

2.1. Governing equation

In this study, the dimensions of the cavity and the temperature difference are such that the flows generated are assumed
to be laminar, incompressible and three dimensional. The viscous dissipation term in the energy equations is neglected. The
surfaces are supposed to be grey and diffuse and separated by the radiative non-participating medium. The net radiative heat
flux distributions are calculated by using the radiosity model.
The thermo-physical properties are assumed to be independent of temperature and are taken to be constant (Table 1),
except for the buoyancy term in the momentum equation for which the density varies with temperature according to the
Boussinesq's approximation:

rf ¼ r0 ð1  bT ðT  T 0 ÞÞ ð1Þ

The flow structure and the temperature distribution are governed by three dimensionless numbers, the Rayleigh number
RaL, the Prandtl number Pr, and the Planck number Pl. The dimensionless governing equations are obtained as follows:

: ~
V ¼ 0 ð2Þ

V :~
~ V  ¼ rP þ Prr2 ~
V   RaL Pruf ~
g
ð3Þ
jg j

~
V :uf ¼ r2 uf ð4Þ

The dimensionless equation of heat conduction in the solid walls is:

us ¼ 0 ð5Þ
The above governing equations are thus given in a dimensionless form by introducing the units of length, L, velocity, V0
=α0/L, pressure, P0=r0 V 20 and temperature difference (T-Tc), DT. The superscript "*" is used for dimensionless quantities.
For the boundary conditions, no-slip velocity conditions are applied at the boundaries. When the radiant interchange
between surfaces is accounted for, the thermal boundary conditions at the adiabatic walls must include the contribution of
the net radiative heat flux. The net radiative heat flux distributions are axially symmetric provided that the flow is also axially
symmetric. Otherwise, the radiosity distributions are being dependent on the wall temperature profiles; dissymmetry in the
flow field implies different radiative heat flux distributions at the vertical walls.

2.2. Radiation formulation

The Surface-to-Surface (S2S) radiation model can be used to account for the radiation exchange in an enclosure of gray-
diffuse surfaces. The energy exchange between two surfaces depends in part on their size, separation distance and
orientation. These parameters are accounted for by a geometric function called a "view factor".
For non-participating media, the coupling between the radiative transfer and Navier–Stokes equations is done via the
radiative heat exchange between surfaces.
Using s T 40 as the radiative flux density scale and the radiosity scale, the dimensionless radiative flux density along the
surface walls is written in the following dimensionless form:
XN Z
qr;i ð~
ri Þ ¼ Bi ð~
ri Þ  j¼1
Bj ð~
rj ÞKð~
ri ; ~
rj ÞdSj ð6Þ
Sj

Table 1
Thermophysical properties of at T0 = 310 K and atmospheric pressure;

Dry air

r0 (kg. m3) Cp0 (J. kg1. K1) K0(W. m1. K1) m0 (kg m1.s1) β0 (K1)
5
1.177 1005.7 0.025 1.74 10 0.0032

Concrete

rS (kg. m3) CpS (J. kg1. K1) KS (W. m1. K1) e


2200 1000 1.4 0,9
M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249 5

Bi ð~
ri Þ is the dimensionless radiosity along surface Si:
XN Z    
r i Þ ¼ ei h4i ð~
Bi ð~ r i Þ þ ð1  ei Þ j¼1
Bj ~ r i ;~
rj K ~ rj dSj ð7Þ
Sj

where hi is the dimensionless temperature parameter, which is defined as hi ¼ TT0 .


ri and ~
~ rj denote the position vectors of the elementary surfaces dSi and dSj and K is the kernel function defined as:
  dF dS dS
K ~r i ;~
rj ¼ j i
ð8Þ
dSj

where dF dSi dSj is the elementary view factor between dSi and dSj [24]. The surfaces involved are the six walls of the 3-D
cavity i.e. N = 6.
The interaction of natural convection with surface radiation comes only from the thermal conditions on the adiabatic
walls. The adiabatic boundary condition includes a conduction term and a radiation term Q r ðni Þ calculated with the aid of
Eq. (6). The dimensionless temperature at any adiabatic walls is such that:

ni ru þ Q  r ð~
~ niÞ ¼ 0 ð9Þ

where ~
ni is the unit normal vector to the wall Si .

2.3. Numerical solution

Calculations were carried out by using the finite-volume code AnsysFluent 14.0 [25]. The governing equations were
solved sequentially with a decoupled implicit scheme.
The convective terms and the diffusive terms are discretized using a second-order upwind scheme for both momentum
and energy equations.
The velocity pressure coupling was solved with the SIMPLE algorithm and the pressure was calculated with a body-force
weighted scheme. Results of the simulations were collected and processed employing in-house software.
Non-uniform grid spacings were used both in x, y and z directions. The calculation domain was mapped with rectangular
structured grids. These grids were non-uniformly distributed with higher concentrations of grid points close to the walls
where the highest velocities and temperature gradients are expected to develop.
The exact solution was assumed to be reached when the maximum residuals applied to all dependent variables were less
than 108. The radiation fluxes were updated once every 10 iterations and the radiosity field was assumed converged when
its maximum normalized was less than 10-6.

3. Code validation

The code was verified considering problems with non-participating media in a three-dimensional cavity in order to
validate the numerical procedure of the radiation exchanges between the surfaces.The three-dimensional differentially
heated cavity with convection and radiation transfers is studied numerically. The radiosity method is used to solve the
radiative heat transfer equation. The obtained solutions are compared with those reported by Colomer et al. [12], for the
same level of accuracy.The main results of this verification process are described below.

3.1. Numerical convection coupled with radiation heat transfer in a cubical box

The numerical solutions for a three-dimensional natural convection heat transfer in an air-filled cubical cavity are
presented. The 3-D cavity has two differentially heated isothermal vertical walls and four adiabatic walls with an emissivity
of 1 (Fig. 1). The temperature difference between the hot and the cold vertical surfaces (DT = Th-Tc) is equal to 20 K. The
computations are conducted for Rayleigh numbers of 103, 104, 105 and 106. The Prandtl number is fixed at 0.71.
The validation of the simulation of the three-dimensional cavity has been completed by the comparison of present results
with that obtained by Colomer et al. [12].
The thermal boundary conditions are:
u = 0 the cold vertical surface at x* = 1 (10)

u = 1 the hot vertical surface at x* = 0 (11)

@u
 þ Q r ¼ 0 the adiabatic walls at y ¼ 0 and 1 ð12Þ
@y
6 M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249

Fig. 1. Three-dimensional differentially heated cavity scheme.

@u
 þ Q r ¼ 0 the adiabatic walls at z ¼ 0 and 1 ð13Þ
@z
The results are presented for the dimensionless total heat flux Qt*at the hot wall, which can be calculated as:

Q t ¼ Q c þ Q r ð14Þ

Q r and Q c
are the radiation and convection contribution to the total heat flux Q t .

@u 
Q t ¼   þ Q r  ð15Þ
@x x ¼0

All the computations are carried out for the three-dimensional cavity of L = 0.025 m side length.

3.2. Effects of surface radiation

A mesh study was conducted to analyze the influence of the mesh spacing on the final result by using non-uniform grids
varying from 213 to 813 control volumes. The reference problem (RaL = 106, Pr = 0.71, Pl = 0.043, u0 = 15) was used.
A grid independence test was conducted for the largest Rayleigh number RaL = 106 and was carried out by evaluating the
influence of grid size on mean differences of temperature, heat flux and velocity fields between one simulation and its
previous coarser solution (Table 2).
The result presented in table2 showed that the difference lower than 2% was obtained by comparing the third and fourth
discretization. For the range of parameters considered in the present work, a (65  65  65) grid represents a good
compromise between accuracy and computational costs and is as fine enough to perform the numerical studies. So, the
results were computed with a 65  65  65 control volumes.
The effects of radiation were considered for four Rayleigh numbers RaL = 103 to 106.The Planck number was set to Pl = 0.043
and the dimensionless reference temperature ratio was u0 = 15 (Table 3).
It was shown that radiation significantly increases heat transfer. The contribution of convection heat transfer becomes
more important as the Rayleigh number increases, and the contribution of radiative heat transfer remains almost constant.
The agreement between the results reported in Table 3 and those obtained by Colomer et al. [12] was quite good.
The result presented in Fig. 2 illustrates the average heat flux at the hot wall as a function of the position along the y-axis
when radiation was taken into account. The present results were in excellent agreement with those reported in [12].

Table 2
Mean difference for the temperature inside the cavity, heat flux and velocity fields between two consecutive meshes, and total heat flux at the hot wall for
differentially heated cavity (e = 1, RaL = 106, Pr = 0.71, u0 = 15).

Mesh du dQ t dV x dV y dV z Q t

21  21  21 – – – – – 12.65
33  33  33 0.3% 6% 6.4% 7.1% 3.9% 11.95
65  65  65 0.2% 2.5% 3.1% 4.2% 1.6% 11.63
81  81  81 0.1% 0.3% 0.6% 1.1% 0.7% 11.60
Stretching βx=βy = 1.08 and βz = 1
M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249 7

Table 3
dimensionless average heat flux at heated wall (e = 1, Pr = 0.71, Pl = 0.043, u0 = 15).

RaL 103 104 105 106

Q c Q r Q c Q r Q c Q r Q c Q r
Present 1.494 3.057 2.056 3.250 4.047 3.317 8.060 3.446
[13] 1.434 3.162 2.062 3.233 3.983 3.385 8.102 3.568

4. Results and discussion

4.1. Alveolar building: coupling of convection, radiation and conduction

The estimation of heat transfer through the structures can be done by the modeling of thermal behavior, essentially due to
the coupling of the convective and radiative thermal transfers within the three-dimensional cellular structures.
This part of the study is concerned with the simulation of heat transfer through two different structures of concrete
blocks among the most used in the building. The objective of these simulations is to find the values of thermal resistances by
specifying the relative importance of each involved heat transfer.
To this end, two configurations are considered consisting of concrete hollow blocks corresponding to NF P 14-13 [23] of
dimensions 10  20  50 cm3 and 20  20  50 cm3, having rows of holes and formed by several rectangular cavities
separated by conductive walls as shown in Fig. 3. The thermal resistances of these two blocks used in the building and given
in the specified Th-U [23] are 0.12 m2.K.W1 et 0.23 m2.K.W-1, respectively.

4.2. Parameters of the calculation

The heat transfer of natural convection, wall thermal conduction and surface thermal radiation was investigated
numerically in the block of alveolar (Fig. 3).
The thermal resistance of a wall, consisting of a simple block of alveolar structures separating two environments at
different ambient temperatures, was studied numerically taking into account the periodicity of the structure.
The periodic boundary conditions were used to simulate a portion of the geometry to have a representative element
consisting of a parallelepiped cavity closed by six solid wall thickness.
However, the symmetry boundary conditions were applied on the vertical transverse walls and the periodic boundary
conditions on the horizontal walls. The two vertical walls are in contact with cold and hot ambient temperatures Te and Ti
(Ti>Te) with a uniform surface heat transfer coefficients hce and hci, respectively.
Based on the dimensions of the air cavity (volume less than 1000 cm3) and the maximum temperature differences (less
than 30  C), it is assumed that the flow is laminar, because the maximum Rayleigh numbers based on (DT = Te-Ti) are less
than Rayleigh number corresponding to the end of laminar regime (108).
So, the Boussinesq approximation is applicable. The radiative heat exchanges between surfaces were calculated using the
radiosity method, assuming that the medium is semi-transparent, non-diffusing and has a very low absorption coefficient.

Fig. 2. Average total heat flux distribution in the y* direction versus z*/ (z* = 0.5; x* = 1).
8 M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249

Fig. 3. (a) concrete block20, (b) concrete block10.

The study is carried out for two configurations of the concrete blocks, as shown in Figs. 4 and 5. The vertical wall thickness
of concrete blocks separating the cells is equal to 1.7 cm, with a thermal conductivity of 1.4 Wm1 K1. The top and bottom
wall surfaces are periodic, while the vertical walls which are not in contact with the hot and cold environments are
considered as planes of symmetry.
The hexahedral mesh was chosen. The ratio of the length of two successive divisions was 1 in the x-direction, 1.08 in the y-
direction and 1.02 in the z-direction.The thermophysical properties of dry air used for the calculations are given in Table 1. It
is proposed to determine the most appropriate mesh to reduce the calculation time without compromising accuracy. It has
been found that the grid 38  25  63 has good accuracy and the most reasonable cost of calculation.

4.3. Concrete blocks of 10

Heat transfer through the concrete blocks occurs mainly by conduction in the solid walls, by natural convection in the air
filling the block holes and by radiation between the internal surfaces of the holes. A series of calculations were conducted for
a hollow concrete block with 10 cm width and the thermal conductivity of the solid walls is kb = 1.4 W.m1. K1). The two
horizontal periodic walls have the same thickness (0.85 cm) and the four vertical walls are 1.7 cm in thickness. The surface
exchange coefficients were set at the values stipulated in [23] (hce = 25 W.m-2. K1 and hci = 7.6925 W.m-2. K1) and the
interior temperature was maintained at Ti = 293 K.The outdoor temperature was increasing from Te = 268 K to 283 K.
The maximum Rayleigh number based on the total width of the block and the difference between the indoor and outdoor
temperatures is RaL = 2.7  105 and RaL = 6.9  105 according to Te = 268 K–283 K.
The convective heat transfer within the cell was better represented by the effective Rayleigh number, Raeff based on the
depth of the cell and the difference between the average temperatures of the hot and cold internal surfaces, Tmc,I and Tmf,i
deduced from an integration of the calculated temperature distributions.
Table 4 and Fig. 6 show that the average total heat flux is divided by 2.5 when the maximum temperature difference D
T = Ti-Te decreases from 25 K to 10 K. It is deduced that the thermal resistance can be considered constant in this range of
temperatures. The calculated value, Rth = 0.13 m2.K.W1 is very close to that reported in [23].
The contribution of heat transfer by conduction represents 55% of the total heat flux. On the other hand, the convective
heat flux is divided by 3.2 and the radiative heat flux by 2.2. This result shows that the relative proportion of heat transfers
decreases by convection (from 16% to 12.4%) and increases by radiation (from 28% to 31.4%) when DTmax decreases.

Fig. 4. Scheme of concrete blocks10.


M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249 9

Fig. 5. Scheme of concrete blocks 20.

Indeed, the effective Rayleigh number (Raeff) was decreasing from 2.25  105 to 9.15  104 when the temperature
difference DTmax decreases by 15 K. Also, the number of Planck was varying from 0.13 to 0.05, which means that the
proportion of the radiation increases when DTmax decreases. The results indicate that natural convection heat transfer in the
hollow block is reduced by heat conduction in the walls and radiation exchange among surfaces.
The z velocity distribution,Vz is shown in Fig. 7. It should be noted that the increase in a temperature difference DT tends
to refine the boundary layers. The temperature profiles in the concrete block walls have an identical appearance in Fig. 8,
regardless of the temperature differences DT within the cavity.
Fig. 9 shows the flow structures by representing the temperature contours corresponding to the reference temperature T0
= 288 K and DTmax = 10 K. The radiation exchanges produce a slight increase in the thickness of the diffusion layers and the
isotherm becomes increasingly distorted. A clear dissymmetry of the temperature distribution inside the cavity can be noted
in Fig. 9b.
Fig. 10 reveals the distortion of the isotherms in the cell ends and indicates a three-dimensional heat transfer effect. The
tightening of these isotherms in the vicinity of the internal walls reveals the importance of the temperature gradient that
appears in these regions to ensure the continuity of the heat flux.

Table 4
Influence of the external temperature variation on the average surface temperatures of the internal walls and of the external surfaces, on the heat flux of the
hot wall and on the thermal resistance of concrete block 10.

Temperatures Ra105

Te(K) Tmc,e(K) Tmc,i(K) Tmh,e(K) Tmh,i(K)


268 271.7 273.1 281.2 282.5 6.9
273 276.0 277.1 283.5 284.6 5.5
278 280.2 280.9 285.9 286.7 4.1
283 284.5 284.9 288.3 288.8 2.7

Heat flux Rth (m2.K.W1)

Te(k) Qt(W) Qcd(W) Qc(W) Qr(W)


268 2.54 1.40 0.41 0.73 0.132
273 2.03 1.12 0.31 0.60 0.132
278 1.52 0.84 0.21 0.46 0.132
283 1.01 0.56 0.13 0.32 0.133
10 M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249

Fig. 6. Effect of the variation of outdoor temperature on the heat flux by conduction, convection, and radiation of concrete block 10.

4.4. Concrete blocks of 20

The configuration studied is that of a concrete block of usual construction including two cells separated by a wall 1.7 cm
thickness as shown in Fig. 5. The thermophysical properties for the calculations are available in Table 1. The surface exchange
coefficients were hce = 25 W.m2. K-1 and hci = 7.6925 W.m2. K-1 and the interior and outdoor temperatures were maintained
at Ti = 293 K and Te = 273 K respectively.
The temperature and the velocity, Vz, distributions on the center line (0, y) are shown in Fig. 11 and 12. It can be seen that
the temperatures in the first cell on the hot side are higher than in the second cell situated on the cold side. It can also observe
the effect of thermal conduction in the vertical partition which constitutes a thermal bridge. The average surface
temperatures on the hot and cold side are Ts,c = 13.8  C and Ts,f=2.57  C respectively.
The low-temperature gradients in the central part and the high-temperature gradients in the vicinity of the partitions
indicate that the convective heat transfer is achieved throughout the laminar boundary layer regime. The partition can be
considered practically isothermal due to the high ratio of thermal conductivity (ks/ ka).
The Table 5 shows the distribution of the heat flux through the concrete block of 20. The value of the total transferred flux
through the external walls is Qt = 1.46 W. It has noticed that more than 50% is due to wall thermal conduction which is
therefore much more important than convection and radiation transfers, indicating the interest of a geometry allowing a

Fig. 7. z velocity distribution, Vz, on the perpendicular line to the center of the plane (xz).
z * = 0.5of concrete block 10
M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249 11

Fig. 8. Temperature distribution on the perpendicular line to the center of the plane (xz), z * = 0.5of concrete block 10.

reduction of thermal bridge effects. The thermal resistance calculated for the concrete block of 20 is Rth = 0.236 m2.K.W1. As
can be seen, the difference between the present calculations and the results reported in Th-U [23] is within 3%.
Figs. 13 and 14 show the temperature and velocity contours in the plane (x* = 0.5) of the concrete block of20. It is observed
the development of very thin boundary layers at the vertical walls.
The thickness of the vertical boundary layers is clearly depending on the Rayleigh number whereas a noticeable
difference between the boundary layers thickness near the vertical and the horizontal walls.
Fig. 15 shows the effect of the radiation exchanges between the internal surface walls of the cells (e = 1). The heat transfers
by natural convection are decreased due to the influence of radiative transfers, but the structure of the flows in the cells
remains similar. On the other hand, the thermal resistance decreases considerably because the contribution of the surface
radiation to the heat transfer between the external and internal ambiances represents nearly 30% of the total heat flux. Also,
the results show that increasing the number of holes while keeping the block width constant decreases the heat loss
(increases the Rth-value) significantly.

Fig. 9. Isotherm corresponding to the reference temperatureofconcrete block 10, T0 = 288 K and DTmax = 10 K (a) With radiation exchanges (b) Without
radiation exchanges.
12 M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249

Fig. 10. Isotherms in the plan (z,y) corresponding toDTmax = 10k, (__)with radiation exchanges.
(- -) Without radiation exchanges, of concrete block 10

Fig. 11. z velocity distribution,Vz,on the perpendicular line to the center of the plane (xz), x* = 0.5 of concrete block 20.
M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249 13

Fig. 12. Temperature distribution on the perpendicular line to the center of the plane (xz), x* = 0.5of concrete block 20.

Table 5
Average surface temperatures of the internal walls and of the external surfaces, heat flux and thermal resistance through the concrete block of 20.

External walls Internal walls Internal walls


Hot side cavity Cold side cavity
Raeff = 1.86  105 Raeff = 2.02  105

Tmc,e Tmh,e Tmh,i Tmc,i Tmh,i Tmc,i


T(K) 275.72 286.95 286.39 281.68 280.53 275.42
Qt(W) 1.46 1.46 0.647 0.606 0.596 0.625
Qcd(W) – – 0.813 0.854 0.864 0.835
Qc(W) – – 0.206 0.184 0.195 0.205
Qr(W) – – 0.441 0.422 0.401 0.420
Rth(m2.K.W1) Present: 0.236 Reference [1] : 0.23 Erreur : 3%

Fig. 13. Isotherms on the plane (x * = 0.5) in concrete block of 20.


14 M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249

Fig. 14. Vz velocity contours on the plane (x * = 0.5) in concrete block of 20.

Fig. 15. Isotherms for 272 K, 275 K, 279 K, 282 K, 285 K, of concrete block 20, (a) without radiation transfer, (b) with radiation transfer, e = 1.

5. Conclusion

This study was undertaken in order to investigate the effects of the thermal conduction in the walls and the surface
radiation exchange, to highlight that the calculations and the measures are in agreement when the heat transfers on the
walls are correctly modeled without introducing specific thermal boundary conditions to an experiment.
M. Ouakarrouch et al. / Case Studies in Construction Materials 11 (2019) e00249 15

The influence on the thermal performance of concrete block walls produced by two configurations of concrete blocks was
analyzed using the indexes of thermal resistance and the heat flux. The three-dimensional continuity, momentum and
energy equations were solved for the calculations by considering the thermal radiation. Two typical concrete blocks were
adopted for the simulations under the same boundary conditions. It is demonstrated that the calculation results have a good
agreement with the data reported in Th-U [23].
The main conclusions drawn from this work are: the heat transfer through the alveolar structures increases considerably
when the radiative heat exchange is taken into account. The analysis of the heat transfer through the concrete block of 20
consisting of a thermally conducting vertical partition and two cells shows that the conductive heat exchange represents
more than 50% of the total heat flux. The thermal conduction in the vertical partition constitutes a thermal bridge, and the
thermal resistance of the concrete block of 20 is 50% greater than that of the concrete block of 10.

References

[1] G.D. Mallinson, G. de Vahl Davis, Three-dimensional natural convection in a box: A Numerical Study, J. Fluid Mec. 83 (part1) (1977) 1–31.
[2] S. Castro-Cardoso, E. Chénier et, G. Lauriat, Transferts de Chaleur dans une structure alvéolaire, Comptes Rendus du Congrès SFT, Grenoble (2003) 507–
512.
[3] A. Abdelbaki, Z. Zrikem, Simulation numérique des transferts thermiques couplées à Travers les parois alvéolaires des bâtiments, Int. J. Therm. Sci. 38
(1999) 719–730.
[4] T. Ait-Taleb, A. Abdelbaki, Z. Zrikem, Numerical simulation of coupled heat transfers by conduction, natural convection, and radiation in hollow
structures heated from below or above, Int. J. Therm. Sci. 47 (2008) 378–387.
[5] F.Al. Nahhas, R.A.M.I. Saada, G. Bonnet, P. Delmotte, Resistance to the fire of walls constituted by hollow blocks: experiments and thermal modeling,
Appl. Therm. Eng. 1 (27) (2007) 258–267.
[6] M. Boukendil, A. Abdelbaki, Z. Zrikem, Numerical simulation of coupled heat transfer through double hollow brick walls: effects of mortar joint
thickness and emissivity, Appl. Therm. Eng. 125 (2017) 1228–1238.
[7] J.J. del Coz Díaz, P.J. García Nieto, J.L. Suárez Sierra, I. Peñuelas Sánchez, Nonlinear thermal optimization and design improvement of a new internal light
concrete multi-holed brick walls by FEM, Appl. Therm. Eng. 28 (8-9) (2008) 1090–1100.
[8] J.J. del Coz Díaz, P.J. García Nieto, J. Domínguez Hernández, F.P. Álvarez Rabanal, A. fem, Comparative analysis of the thermal efficiency among floors
made up of clay, concrete and lightweight concrete hollow blocks, Appl. Therm. Eng. 30 (17–18) (2010) 2822–2826.
[9] D.W. Pepper, K.G.T. Hollands, Summary of benchmark numerical studies for 3-D natural convection in an air-filled enclosure, Numer. Heat Transf. Part A
Appl. 42 (1-2) (2002) 1–11.
[10] E. Tric, G. Labrosse, M. Betrouni, A first incursion into the 3-D structure of natural convection of air in a differentially heated cubic cavity, from accurate
numerical solutions, Int. J. Heat Mass Transf. - Theory Appl. 43 (21) (2000) 4043–4056.
[11] I.D. Piazza, M. Ciofalo, MHD free convection in a liquid-metal filled cubic enclosure, I. Differential heating, Int. J. Heat Mass Transf. - Theory Appl. 45 (7)
(2002) 1477–1492.
[12] G. Colomer, M. Costa, R. Consul, A. Oliva, Three-dimensional numerical simulation of convection and radiation in a differentially heated cavity using the
discrete ordinates method, Int. J. Heat Mass Transf. - Theory Appl. 47 (2004) 257–269.
[13] L. Kolsi, A. Abidi, Ch. Maatki, M.N. Borjini, H.B.E.N. Aissia, Combined radiation-natural convection in three-dimensional verticals cavities, Therm. Sci. 15
(2) (2011) 327–339.
[14] L.P. Li, Z.G. Wu, Z.Y. Li, Y.L. He, W.Q. Tao, Numerical thermal optimization of the configuration of multi-holed clay bricks used for constructing building
walls by the finite volume method, Int. J. Heat Mass Transf. - Theory Appl. 51 (2008) 3669–3682.
[15] L.P. Li, Z.G. Wu, Y.L. He, G. Lauriat, W.Q. Tao, Optimization of the configuration of 29014090 hollow clay bricks with 3-D numerical simulation by
finite volume method, Energy Build. (4) (2008) 1790–1798.
[16] S. Lorente, M. Petit, R. Javelas, Simplified analytical model for thermal transfer in vertical hollow brick, Energy Build. 24 (1996) 95–103.
[17] S. Lorente, M. Petit, R. Javelas, The effects of temperature conditions on the thermal resistance of walls made with different shapes vertical hollow
bricks, Energy Build. 28 (1998) 237–240.
[18] S. Lorente, A. Bejan, Combined flow and strength geometric optimization: internal structure in a vertical insulating wall with air cavities and prescribed
strength, Int. J. Heat Mass Transf. 45 (16) (2002) 3313–3320.
[19] Y. Zhang, K. Du, J. He, L. Yanga, Y. Lia, S. Lia, Impact factors analysis on the thermal performance of hollow block wall, Energy Build. 75 (2014) 330–341.
[20] M.A. Antar, H. Baig, Conjugate conduction-natural convection heat transfer in a hollow building block, Appl. Therm. Eng. 29 (2009) 3716–3720.
[21] M. Khatamifar, W. Lin, S.W. Armfield, D. Holmes, M.P. Kirkpatrick, Conjugate natural convection heat transfer in a partitioned differentially-heated
square cavity, Int. Commun. Heat Mass Transf. 81 (2017) 92–103.
[22] G. Yang, J.Y. Wu, Effects of natural convection, wall thermal conduction, and thermal radiation on heat transfer uniformity at a heated plate located at
the bottom of a three-dimensional rectangular enclosure, Numer. Heat Transf. 69 (issue 6) (2016) 589–606.
[23] R.ègles Th-U, Fascicule 4/5, Calcul Des Caractéristiques Thermiques Des Parois Opaques, CSTB, 2012.
[24] M.F. Modest, Radiative Heat Transfer, second ed., Academic Press, 2003, pp. 162–197 Chapter 5.
[25] Ansys Fluent 14 User’s Guide (n.d), www.fluentusers.com.

You might also like