You are on page 1of 35

Benchmark solutions for natural convection flows in

vertical channels submitted to different open boundary


conditions
G. Desrayaud, Eric Chénier, A. Joulin, Alain Bastide, Boris Brangeon, J.P.
Caltagirone, Y. Cherif, R. Eymard, C. Garnier, S. Giroux-Julien, et al.

To cite this version:


G. Desrayaud, Eric Chénier, A. Joulin, Alain Bastide, Boris Brangeon, et al.. Benchmark solutions
for natural convection flows in vertical channels submitted to different open boundary conditions.
International Journal of Thermal Sciences, 2013, 72, pp.18-33. �10.1016/j.ijthermalsci.2013.05.003�.
�hal-00841302�

HAL Id: hal-00841302


https://hal.science/hal-00841302
Submitted on 4 Jul 2013

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
Benchmark solutions for natural convection flows in vertical
channels submitted to different open boundary conditions✩
G. Desrayaud1 , E. Chéniera,2,∗, A. Joulinb,2, A. Bastidec , B. Brangeonc , J.P. Caltagironed , Y.
Cherifb , R. Eymarde , C. Garnierf,g , S. Giroux-Julienh, Y. Harnanei , P. Joubertj , N. Laaroussik, S.
Lassueb , P. Le Quéréf , R. Lia , D. Sauryi , A. Sergentf,g , S. Xinh , A. Zoubirh
a
Université Paris-Est, Laboratoire Modélisation et Simulation Multi Echelle, UMR-CNRS 8208, 5 Boulevard
Descartes, 77454 Marne-la-Vallée Cedex 2, France
b
Université Lille Nord de France, Laboratoire Génie Civil et géo-Environnement, EA 4515, Technoparc Futura,
62400 Béthune, France
c
PIMENT, 117 Avenue du Général Ailleret 97430 Le Tampon, France.
d
I2M – TREFLE Site ENSCBP - 16 Avenue Pey-Berland 33607 Pessac Cedex, France
e
Université Paris-Est, Laboratoire d’Analyse et de Mathématiques Appliquées, UMR-CNRS 8050, 5 Boulevard
Descartes, 77454 Marne-la-Vallée Cedex 2, France
f
CNRS, LIMSI, BP 133, 91403 Orsay, France
g
UPMC, Univ Paris 06, F-75005, Paris, France
h
CETHIL UMR 5008 Bâtiment Sadi Carnot 9 Rue de la Physique INSA de LYON 69621 Villeurbanne Cedex, France
i
Institut PPRIME UPR CNRS 3346 - Université de Poitiers, ENSMA -Téléport 2 1 avenue Clément ADER, BP
40109, 86961 Chasseneuil Futuroscope Cedex, France
j
LaSIE, Avenue Michel Crépeau 17042 La Rochelle Cedex 1, France
k
Ecole Supérieure de Technologie de Salé, Université Mohammed V-Agdal, Laboratoire d’Energétique, Matériaux et
Environnement (LEME), Avenue Prince Héritier, BP :227 Salé Medina, Maroc

Abstract
Comparison exercises have been carried out by different research teams to study the sensitivity of
the natural convection occurring in a vertical asymmetrically heated channel to four sets of open
boundary conditions. The dimensionless parameters have been chosen so that a return flow exists
at the outlet. On the whole, results provided by the partners are in good agreement; benchmark
solutions are then defined for each of the boundary conditions. Whilst the local and average
Nusselt numbers based on the entrance temperature do not depend much on conditions applied in
the aperture sections, the net fluid flow rates crossing the channel and the characteristics of the
recirculation cells are highly influenced. But we proved that these modifications of flow patterns
do not alter significantly the fluid flow rates leaving the channel through the exit section.
Keywords: Natural convection, Vertical channel, Recirculation flow, Open boundary conditions,
Benchmark solutions

Preprint submitted to International Journal of Thermal Sciences May 3, 2013


Nomenclature
A aspect ratio of the channel, = H/l
BC. boundary conditions
dw width of the downward flow (Eq. 13)
dψ width of the recirculation (Eq. 14)
(~ex , ~ez ) coordinate axes
Fr Froude number, = α20 /(g l3)
g gravitational acceleration, m/s2
GB boundary conditions (Eqs. 5c and 6c), see Tab. 1
GB-0 boundary conditions (Eqs. 5c and 6d), see Tab. 1
H channel height, m
l channel width, m
LB boundary conditions (Eqs. 5b and 6c), see Tab. 1
LB-0 boundary conditions (Eqs. 5b and 6d), see Tab. 1
f 1 ; Nu1
Nu inverse reduced temperature at the left wall; Nusselt number on the heated surface
based on the reference temperature (Eq. 10)
f 2 ; Nu2
Nu inverse reduced temperature between the left wall and the bulk; Nusselt
number on the heated surface based on the bulk temperature (Eq. 11)
p difference between static and hydrostatic pressures, = Π + z/Fr
Pr Prandtl number, = ν0 /α0
qin (z = 0) mass flow rate entering through z = 0 (Eq. 7)
qin (z = A) mass flow rate entering through z = A (Eq. 8)
Ra Rayleigh number, = gβ0 Φl4 /(λ0 ν0 α0 )
t time
T temperature, K
T1, · · · , T8 participating teams, see Tab. 2
(u, w) velocity components, u = ~v · ~ex , w = ~v · ~ez
~v velocity vector
(x, z) spatial coordinates


Supported by the French Research Group GDR AMETh: Analyse – Maîtrise des Ecoulements et Echanges Ther-
miques

Corresponding author
Email addresses: Eri .Chenieruniv-paris-est.fr (E. Chénier), Annabelle.Joulinuniv-artois.fr
(A. Joulin)
1
To the memory of Gilles Desrayaud, deceased in January 2009
2
Coordinators of the numerical exercise

2
Greek symbols
α thermal diffusivity, m2 /s
β thermal expansion coefficient, 1/K
θ reduced temperature, = (T − T 0 ) × λ0 /(lΦ)
θb bulk temperature (Eq. 9)
λ thermal conductivity, W/(m · K)
µ dynamic viscosity, kg/(m · s)
ν kinematic viscosity, m2 /s
Π static pressure
ρ density, kg/m3
Φ heat flux, W/m2
ψ stream function (Eq. 4)

Subscripts
0 reference temperature
m median value (Eq. 16), benchmark solutions (Fig. 9)
σ standard deviation (Eq. 15)

Others
spatial average on the heated wall (Eq. 12)
hi ensemble average (Eq. 15)

1. Introduction
Heat transfer and fluid flows driven by natural convection in open channels have been exten-
sively studied over the last past decades, for vertical or inclined configurations. This great interest
raised by this subject stems from its wide range of practical applications such as solar chimney,
solar energy collectors, Trombe walls, or the cooling of electronic components and many others
[1–4].

Since the precursory experimental works performed by Elenbaas in 1942 [5], who determined
the different flow regimes versus a modified Rayleigh number or Elenbaas number (Rayleigh
number calculated on the channel-chimney width divided by the aspect ratio), natural and mixed
convections have been widely studied in open channels, both numerically and experimentally.
Because the aim of this paper is to focus our attention on the influence of the boundary condi-
tions applied in the apertures of the channel, and then to define benchmark solutions, the detailed
description of the numerous contributions dealing with natural or mixed convection in vertical
channels is transferred to few relevant and recent papers chosen to provide complete reviews of
this topic (see [1, 6–13] and references therein). Among the first numerical simulations, we can
mention the contribution of Bodoia and Osterle [14] who investigated fluid flows and heat transfer
occurring through isothermal vertical plates. Their results were in good agreement with the Elen-
baas experimental data [5]. Since then, and despite the plenty of numerical studies, the choice of
3
the boundary conditions for open cavities is still a delicate issue.
For thermal natural convection, the distribution of the total heat fluxes conveyed by the fluid
flow through the open boundaries depends on heat transfered at walls but also on physical con-
ditions prevailing in the surroundings, on both sides of the apertures. It results a close coupling
between the dynamic and thermodynamic variables inside and outside the channel. Thus, the
thermal and kinematic inlet/outlet boundary conditions cannot be a priori prescribed without ac-
counting for the surrounding conditions [6, 15, 16]. Finally, the thermally driven channel behaves
as a thermal engine by converting the thermal gradient into a fluid flow. This corresponds to the so-
called thermosyphon effect. Although thermal natural convection was only considered here-above,
this applies for general mass transfer, whatever the origin of the density variation may be.
To overcome the issue of choosing the boundary conditions in inlet/outlet sections, some au-
thors proposed to extend the computational domain, upstream and/or downstream the channel.
Thus, the boundary conditions are pushed away from the true inlet/outlet sections of the channel.
One of the ideas leading to the displacement of the open boundaries far from the channel limits
is to reduce, as much as possible, the effect of an unawareness of the “real” boundary conditions.
St. Venant’s principle, often invoked in solid mechanics, may also be used in fluid mechanics:
if the artificial boundaries are placed sufficiently far away from the channel apertures, the veloc-
ity and temperature distributions at the entrance/exit of the channel are no longer affected by the
applied boundary conditions [17]. The issue now raised is how far the artificial open extensions
must be placed in order that the physical quantities in the inlet/outlet sections become insensitive
to the conditions set at boundaries. Although this approach seems attractive, the increase in size
of the computational domain proves to be expensive, both in memory and in computational time.
For these reasons, the domain extensions are often either relatively reduced or large but coarsely
discretized. The shapes of the walls at the entrance region, with sharp angles or smooth rounded
surfaces, affect also significantly the fluid flows and heat transfer. Using entrance walls with right
angles, Naylor et al. [18] predicted a fluid separation at the channel inlet which is approximatively
correlated with the dimensional flow rate. Their inlet boundary conditions were based on Jeffrey-
Hamel flow which consists in a similarity solution of isothermal flow caused by the presence of a
source or sink at the point of intersection of two walls. Kettleborough [17], with a parabolic ap-
proximation, and Nakamura et al. [19], with a full elliptic model, used boundary conditions which
physically correspond to fully developed flow entering a channel with a large sudden section re-
duction. They also found a separation of the boundary layer, but at some distance from the leading
edge at the entry to the channel. This issue was reconsidered in a recent paper by Boetcher and
Sparrow [20] who studied buoyancy-induced flow in an horizontal open-ended cavity. They ex-
amined the impact of the size of the extended domain, the boundary conditions on its surfaces, and
the mesh density required to achieve high accuracy. Just as the upstream extensions may modify
the flow and heat transfer in the channel-chimney, the increase in the computational downstream
domain may lead to modifications in the draft height and thermal transfer, for instance by adding
adiabatic extensions in order to enhance the sucking up of the chimney [21].
Another way to sidestep the question of choosing open boundary conditions in the entrance/exit
sections of the channel is to encapsulate it into a virtual enclosure [22, 23]. As a result, the
boundaries are solid walls on which it is easy to prescribe dynamical conditions. However, thermal
boundary conditions are not so easy to define since they may disturb the thermal field around the
4
channel and produce spurious natural recirculation cells.
The different strategies described here-above rest on an increase in size of the computational
domain, and therefore, in computational memory storage and CPU time. An alternative is to re-
strict simulations to the channel height only, and to model the thermal and dynamical boundary
conditions at the exact inlet/outlet sections. In comparison with numerical results by Bodoia and
Osterle [14], Aihara [15] proposed to link the pressure to the mass flow rate. This formulation
permits to substitute the unknown inlet velocity by a relation based on pressures. This writing is
related to a simplified model of the upstream fluid flow sucked by the pressure drop produced in
the channel aperture. In the upstream region, the fluid motion is then assumed incompressible,
isothermal, inviscid and stationary. Dalbert et al. [16] were among the first to deal with this pres-
sure drop which corresponds physically to the energy per volume unit required to bring the fluid
from rest far upstream to the entrance section. Using Jeffrey-Hamel flow extension at the entrance,
Naylor et al. [18] confirmed the relationship between the kinetic energy and the pressure drop in
the inlet section of the channel.

Since a long time, simulations performed by numericists rely on reference solutions to check
both the validity of discretization schemes and their numerical implementations. For laminar flows
and closed systems, such benchmark solutions are well documented. For example, one may men-
tion the lid driven cavity for isothermal fluid flows and the Rayleigh-Bénard problem or the differ-
entially heated cavity (window problem) for flows driven by buoyancy. But to our best knowledge,
and despite very numerous numerical references, such a work is still lacking for free convection
flowing through open boundaries. This issue was already addressed by the French community
dealing with thermal sciences for a geometrical model restricted to the channel height [24]. But
this first attempt to define reference solutions turned out to be unsuccessful: the gaps between
results provided by the research teams were too large, and no reference solution was defined. A
possible explanation of these discrepancies was recently proposed in [25]. For one particular set
of boundary conditions, Le Quéré showed that the use of Neumann type boundary conditions may
give one or several nontrivial combinations of velocity-pressure fields which satisfy the homoge-
neous Stokes operator, in addition to the unavoidable constant pressure mode. This recognition
leads him to propose an algorithm in which the solution is sought as a combination of particular
solution of the inhomogeneous Stokes or unsteady Stokes problem, plus a linear combination of
the modes of the Stokes kernel so as to satisfy the pressure drop between the inlet and outlet.
Considering the previous works [24, 25] and the obvious interest of the numericists community
to have reference solutions for free convection with open boundaries, we reformulated the test case
problem [24] on the basis of four boundary conditions frequently met in such studies. This new
approach leads us to study the sensitivity of fluid flows and heat transfer to the different prescribed
boundary conditions.

The purpose of this paper is then twofold. First, it aims at contributing to the numerical study of
natural convection in vertical channel by considering some sets of boundary conditions, but always
restricting the geometry to the physical domain, without any downstream or upstream domain
extensions. These boundary conditions are, for example, implemented in some commercial codes
or free packages solving the Navier-Stokes and energy equations. The second objective concerns
5
the definition of reference solutions that could be very useful in the community of numericists
for validating the numerical schemes in open systems, when flows are driven by natural or even
mixed convection. In order to propose a numerical exercise, both relevant and numerically delicat,
and to emphasize the effect of boundary conditions on fluid flow and heat transfer, we focused our
attention on a problem which gives rise to a recirculation flow at the exit of the channel [26, 27].
The rest of the paper is organized into two main sections followed by a conclusion. Section
2 provides, first the mathematical model with the four sets of boundary conditions, and then de-
scribes the different variables used for comparisons. Section 3, which deals with the results and
their discussions, is subdivided into three subsections. A qualitative description of the flow and
temperature fields is first given. Then, the results of the different research teams are synthesized
and commented; reference solutions are defined. Finally, the influence of the different boundary
conditions on heat and fluid flow is presented and analyzed. A conclusion summing up the key
points of this contribution ends the paper.

2. Description of the test cases


2.1. Heat and fluid flow equations
A vertical parallel plate channel of width l and height H is formed by two walls, one partially
heated at a constant and uniform heat flux φ on its half middle section and the remaining walls are
adiabatic (see fig. 1) [26]. The fluid flow is assumed laminar and two-dimensional. The thermal
radiations and the heat conduction inside the solid walls are disregarded. Accounting for the small
relative temperature difference occurring between the heated wall and the aperture, the Navier-
Stokes and energy equations are expressed with the Boussinesq approximation. The problem to
be numerically solved is restricted to the channel height. The governing flow and heat transfer
equations, written in dimensionless form, read:


∇ · ~v = 0 (1)

∂~v → − →

+ ∇ · (~v ⊗ ~v) = − ∇ p + Pr∇2~v + Ra Pr θ~
ez (2)
∂t
∂θ → −
+ ∇ · (~vθ) = ∇2 θ (3)
∂t
with ~v = ue~x + w~ez the dimensionless velocity vector, p = Π + z/Fr the dimensionless departure of
the static pressure from the hydrostatic pressure, and θ = λ0 (T −T 0 )/(φl) the reduced dimensionless
temperature. The reference temperature T 0 is set to the temperature of the surroundings. Thermal
conductivity, thermal and viscous diffusivity and thermal expansion coefficient are expressed at
T 0 by λ0 , α0 , ν0 and β0 respectively. The length, velocity and pressure are scaled by l, α0 /l and
ρ0 α20 /l2 . The dimensionless parameters governing the fluid flow and heat transfer are the Prandtl
number Pr = ν0 /α0 , the Rayleigh number Ra = gβ0 φl4 /(λ0 ν0 α0 ) and the aspect ratio of the channel
A = H/l. The dimensionless streamfunction is defined as usual by
∂ψ ∂ψ
= u, = −w (4)
∂z ∂x
6
2.2. Boundary conditions
Four sets of boundary conditions are referenced in this paper. In each case, the conditions
at the solid walls are the same; only the boundary conditions at the inlet and outlet sections are
modified.
Boundary conditions at solid plates.

• Left wall, x = 0,




 ∀z ∈]0; A[, ~v(0, z) = ~0





 ∀z ∈]0; A/4[∪]3A/4; A[, ∂θ

 (0, z) = 0


 ∂x


 ∂θ

 ∀z ∈]A/4; 3A/4[, (0, z) = −1
∂x
• Right wall, x = 1, ∀z ∈]0; A[, 

 ~
 ~v(1, z) = 0


 ∂θ

 (1, z) = 0
∂x
Channel inlet, z = 0, ∀x ∈]0; 1[, 


 u(x, 0) = 0


 ∂w

 (x, 0) = 0 (5a)


 ∂z
 θ(x, 0) = 0

Two pressure boundary conditions are studied. They result from Bernoulli’s theorems which
assume stationary, incompressible and inviscid fluid flows.

• Local Bernoulli relation (LB):


1 2
∀x ∈]0; 1[, p(x, 0) = − w(x, 0) (5b)
2
• Global Bernoulli relation (GB):
1 2
∀x ∈]0; 1[, p(x, 0) = − qin (z = 0) (5c)
2
R1
where qin (z = 0) = 0
w(x, 0) dx denotes the dimensionless mean velocity or flow rate.

Channel outlet, z = A. Although the designation “outlet” (or “exit”) is somewhat incorrect in our
problem, since fluid partially enters through the top boundary and creates a recirculation
flow, but for sake of simplicity, the use of this term is however kept. To establish the bound-
ary relations, it is then necessary to distinguish the conditions corresponding to entering
(w(x, A) < 0) and exiting (w(x, A) > 0) fluid flows. Furthermore, two pressure conditions
are studied when the fluid flows out of the channel.
7
• Exiting fluid:



 u(x, A) = 0




 ∂w


 (x, A) = 0
 ∂z
For all x ∈]0; 1[, such that w(x, A) ≥ 0, 
 ∂θ (6a)




 (x, A) = 0


 ∂z

 p(x, A) = 0

• Entering fluid:



 u(x, A) = 0


 ∂w
For all x ∈]0; 1[, such that w(x, A) < 0, 
 (x, A) = 0 (6b)


 ∂z
 θ(x, A) = 0

with one of the two following pressure boundary conditions:


– Local relation
1 2
For all x ∈]0; 1[, such that w(x, A) < 0, p(x, A) = − w(x, A) (6c)
2
– Uniform Pressure

For all x ∈]0; 1[, such that w(x, A) < 0, p(x, A) = 0 (6d)

The four sets of boundary conditions are synthesized with notations (LB, LB-0, GB and GB-0)
according to the rule given in Tab. 1. From a mnemonic point of view, the notations LB (Local
Bernoulli, Eq. (5b)) and GB (Global Bernoulli, Eq. (5c)) designate the boundary conditions at the
lower channel aperture. The addition of the suffix -0 indicates that the outlet pressure is uniformly
set to zero (see Eqs. (6a) and (6d)), otherwise local expressions are applied for the return flow
(Eqs. (6a) and (6c)). It must be emphasized that the pressure boundary condition is a continuous
function of the abscissa, even if a reversal flow occurs at the channel exit.
Before introducing the variables which will lead to quantitative comparisons, let us perform
some remarks about thermal radiation. First, and as announced at the beginning of Sec. 2.1,
the thermal radiation has been neglected, since the aim of the paper is to focus the analysis on
pressure boundary conditions and the definition of benchmark solutions in the scope of pure natural
convection. However, from a physical point of view, working with transparent fluids like air gas
makes the surface temperature sensitive to the surface radiation phenomena. In that case, it is worth
pointing out that, despite small relative temperature differences, surface radiations may strongly
modify the flow characteristics, even if the emissivities are very small [28]. Consequently, the
radiation exchanges must always be taken into account in the numerical model if comparisons
with experimental data are planned (see the works by Li et al. [29]).

8
2.3. Monitored variables
Comparisons have been carried out for mean and local quantities at different discrete vertical
coordinates located in the heated region and in the downstream adiabatic domain, namely for
z ∈ {3A/8; A/2; 5A/8; 3A/4; 7A/8; A}. These concern some

dynamic aspects:

• the mass flow rate entering the channel through the bottom section z = 0
Z 1
qin (z = 0) = w(x, 0) dx (7)
0

• the mass flow rate entering into the channel through the top section z = A
Z 1
|w(x, A)| − w(x, A)
qin (z = A) = dx (8)
0 2

thermal variables:

• the bulk temperature


Z 1
1
θb (z) = w(x, z)θ(x, z) dx (9)
qin (z = 0) 0

• the inverse of the temperature at the left wall

f 1 (z) = 1
Nu (10)
θ(0, z)
whose expression corresponds to the local Nusselt number Nu1 (z) on the heated wall
only, with a reference temperature equal to the surrounding value; Nu1 (z > 3A/4) = 0.

• the inverse of the difference between the temperature at the left wall and the bulk
temperature
f 2 (z) = 1
Nu (11)
θ(0, z) − θb (z)
whose expression represents the local Nusselt number Nu2 (z) on the heated wall only,
with the bulk temperature as reference temperature; Nu2 (z > 3A/4) = 0.
• the average Nusselt numbers on the heated wall
Z
2 3A/4
Nui = Nui (0, z) dz, i ∈ {1, 2} (12)
A A/4

characterization of flow patterns:

9
• the length of the downward flow in z-section. For z given, if there exists 0 < xw (z) < 1
such that w(xw (z), z) = 0, then

dw (z) = 1 − xw (z) (13)

• the recirculation length in z-section. For z given, if there exists 0 < xψ (z) < 1 such that
ψ(xψ (z), z) = ψ(1, z), then
dψ (z) = 1 − xψ (z) (14)
with the streamfunction defined in Eq. (4).

3. Results and discussions


The numerical exercise has been carried out for Ra = 5 × 105 , A = 10 and air as working fluid
with Pr = 0.71.

Results were obtained by eight partners which are parts of different laboratories: CETHIL,
I2M–TREFLE, LAMA, LaSIE, LEME, LGCgE, LIMSI, MSME, PIMENT and PPRIME. Five of
the numerical codes were developed by the research teams, two are free CFD software packages
(Aquilonr, FDS) and the last one is commercial (Fluentr). Table 2 indicates, for the different
partners, the name of the laboratories involved, and some characteristics of the numerical schemes.
More details about the numerical method are provided in the appendix section.

3.1. Qualitative description of the fluid flow and temperature field


The fluid flow and the temperature field, illustrated in Fig. 2 for boundary conditions LB
(Tab. 1), are qualitatively almost the same whatever the boundary conditions considered in this
paper. The fluid is heated along the mid-height of the central section of the left wall by a uniform
and constant heat flux Φ (Fig. 2(d)). Consequently, the fluid rises up due to the density variations
and creates a dynamic boundary layer along the left plate (Fig. 2(c)). Due to too large head loss
in the upstream adiabatic part, the feeding of this dynamic boundary layer cannot exclusively hap-
pen by an air supply coming from the lower aperture of the channel and then an incoming of fluid
through the top open section of the channel is created over a width dw (z = A) (Fig. 2(c)). Therefore,
a recirculation flow occurs with a measured size at the channel exit dψ (z = A) (Fig. 2(a)).
It must be underlined that the field of the horizontal component (Fig. 2(b)) displays positive
and negative values in the vicinity of the left and right corners of the entrance region. These
patterns are typically met with BC. (5b), i.e. when the inlet pressure is set to the local dynamic
pressure. If the pressure boundary condition in the aperture is kept uniform BC. (5c), these local
variations in the horizontal velocity do not exist. The small fluid pocket, defined by u(x, z) > 0
and located near the center of the channel exit section, has been identified by all the participants
working on LB problem. When other boundary conditions are applied, it disappears.

10
3.2. Comparison of numerical contributions
Figures 3, 4, 5 and 6 provide Nu f 1 (z) (Eq. 10) and Nu f 2 (z) (Eq. 11), viz the local Nusselt
numbers Nu1 (z) and Nu2 (z) when A/4 ≤ z ≤ 3A/4, the local sizes of both the return flow dw (z)
(Eq. 13) and the recirculation flow dψ (z) (Eq. 14), the local bulk temperature θb (z) (Eq. 9), and four
overall values defined by the mean Nusselt numbers on the heated plate Nu1 and Nu2 (Eq. 12), and
the flow rate penetrating the channel through the bottom qin (z = 0) (Eq. 7) and the top qin (z = A)
(Eq. 8) sections, for boundary conditions GB-0, LB-0, GB and LB. The results of each research
team have been collected and graphically represented using histograms with the corresponding
numerical values accurate within 4 significant digits, and topped by three statistical quantities, the
average (hxi), the standard deviation (xσ ) and median value (xm ) defined as follows:

1X
n p
hxi = xi , xσ = hx2 i − hxi2 (15)
n i=1
(
x(n+1)/2 , if n is odd, with x1 ≤ · · · ≤ xn
xm = (16)
(xn/2 + xn/2+1 )/2, if n is even, with x1 ≤ · · · ≤ xn
with x any of the recorded variables mentioned here-above. When results are in excellent agree-
ment, the mean and median values are very close, and the standard deviation is small in comparison
with the average value. In that case, the mean or median values provide very good approximations
of the reference solution. In contrast, if some results depart significantly from the other contribu-
tions, the mean value can differ substantially from the median value. This bias which is introduced
is related to the relative small number of partners involved in the comparison exercice. In that
case, the median value is proved to be better representative of the results than the average value.
Therefore, the reference results are defined by the measure of the median xm .
On the whole, the results provided by the different teams are in quite good agreement. A
deeper insight is now proposed for the 4 sets of boundary conditions.

We first consider boundary conditions GB-0 and the associated results presented in Fig. 3. We
notice that the standard deviations are small (xσ /hxi < 0.6%) for the average Nusselt numbers
x ≡ Nu1 and Nu2 (Fig. 3(f)). This good agreement between the different teams is confirmed when
we look at the local values Nu f 1 (z) (Fig. 3(a)) and Nu
f 2 (z ≤ 3A/4) (Fig. 3(b)), since the average
discrepancy does not exceed 1.7%. Beyond z = 3A/4, the agreement between the participants
becomes all the more worse since one goes closer to the top section (Nu f 2 i ≈ 7% for z = A.).
g2σ /hNu
f
Let us recall that Nu1 (z) is inversely proportional to θ(0, z) (Eq. 10), and therefore its dispersion
results necessary from disagreements in the temperature distribution along the left plate. The
second definition Nuf 2 (z) is inversely proportional to the difference between the wall temperature
and the bulk temperature θb (z) (Eq. 11). Thus, the more the temperature gap decreases, the
more this value will be sensitive to the computational uncertainties. The variations of the bulk
temperature θb (z) (Fig. 3(e)) are in good accordance, except for two families of results that were
provided by teams T1 and T6. This discrepancy shown on the bulk temperature is found on the
dynamical variables as well. Indeed, their two flow rates qin (z = 0) and qin (z = A) (Fig. 3(f)),
their sizes of the recirculation flow dψ (z) (Fig. 3(d)) and their lengths corresponding to the fluid
11
flow penetrating the channel through the top section dw (z) (Fig. 3(c)) differ slightly from the data
provided by the other participants: whereas both the width of the recirculation dψ (z) looks smaller
and the return flow seems to penetrate less deeply into the channel for team T6, the perfect opposite
behavior is observed for results given by T1. Attempts of explanation for these departures are now
given.
The numerical method developed by T6 results from a recent analysis of the discrete Stokes
problem based on a staggered grid formulation with a projection method to uncouple the veloc-
ity and pressure [25, 30, 31]. The usual homogeneous Neumann boundary conditions applied to
the pressure increment may lead to an increase of the kernel size of the Stokes operator making
the solution of the full nonlinear equations indefinite. This arises in particular when Neumann
type boundary conditions are imposed on the velocity component normal to the inlet and outlet
boundaries. The natural convection problem under consideration falls into this category. An ap-
propriate algorithm has been proposed that is derived from the principle of superposition. The
final numerical solution is a linear combination of a particular solution provided by a conventional
finite volume scheme and of the (velocity, pressure) modes belonging to the Stokes kernel. These
novel works clearly raise the issue of the choice of the artificial boundaries conditions but also of
their numerical implementation. As it has been shown in [25, 30, 31] the numerical method used
by T6 leads to control two parameters in the linear combination of the Stokes’s kernel modes. To
determine them, two specific constraints of the flow have to be specified. The choice has been
made to impose the mean pressure in the outlet section equal to zero and the mean pressure in the
inlet section equal to −(qin (z = 0))2 /2. The discrepancies observed on GB-0 results are then due to
the fact that T6 imposed mean pressures on the inlet and outlet sections and not uniform pressures
as requested by Eqs. 5c and 6d. Thus, T6 solved a neighboring physical problem which seems
less restrictive for the flow field than GB-0 does. Consequently, the comparison with the results
by T6 shows the great dependence of the problem on the pressure boundary conditions.
The explanations for the departures of T1 results are unclear, but some specific points can be
pointed out. First, the FDS software used by T1 rests on a low Mach number approximation, suit-
able for low speed thermally-driven flows with large temperature variations. However, this model
should provide results in accordance with those obtained with Boussinesq approximation since
the maximal relative temperature never exceeds ≈ 5/290 = 1.7%. Another remark can be drawn
about the accuracy of computations. Indeed, the mass flow rate shows a slight increase of about
0.1% from the inlet to the outlet sections. Although this variation may be considered as negligible,
the natural convection flows in open channel are proved to be very sensitive to numerical errors,
and specially to mass conservation. Thus, this lack of numerical accuracy may explain the small
disagreements noticed on these results.

For boundary conditions LB-0 (Fig. 4), the data given by teams T2, T3 and T8 are overall in
good agreement: xσ /hxi is less or equal than 1% for x ≡ Nuf 1 (z), Nu
f 2 (z ≤ 5A/8), dw (z ≥ 3A/4),
dψ (z ≥ 3A/4), θb (z) and also for the mean Nusselt numbers and mass flow rates (Fig. 4(a)-(f)).
Lastly, the ratio between the standard deviation and the average value is about 2% for dw (5A/8)
and dψ (5A/8) and reaches up to 22% for Nu f 2 (A)! This latter very large gap is due to both the
bulk temperature accuracy (≈ 1%) and the small temperature difference between the wall and
bulk. To support this assertion, let us evaluate roughly Nuf 2 (z = A) from the bulk and wall
12
temperatures at z = A. The wall temperature at the channel exit is derived from the defini-
f 1 (Eq. 10), the mean value hNu
tion of Nu f 1 (A)i = 11.69 and the relative standard deviation
g1σ (A)/hNu
Nu f 1 (A)i = 0.1%: θ(x = 0, z = A) = 1/11.69 ± 0.1%. One gets θ(x = 0, z = A) − θb (A) =
(0.08554 ± 0.1%) − (0.08298 ± 0.9%) = 2.56 × 10−3 ± 8.3 × 10−4 , that is an error of 32%. Finally,
f 2 (A) ≈ 390 ± 32%, namely an average value very close to that reported in Fig. 4(b)
one has Nu
for z = A, but with an error about 45% larger than the corresponding standard deviation value
g2σ (A)/hNu
(Nu f 2 (A)i = 22%).

The results provided by teams T2, T3 and T8 for GB boundary conditions (Fig. 5) as well as
those given by T2, T3, T7 and T8 for LB boundary conditions (Fig. 6) are in excellent agreement,
both for the thermal and the dynamical quantities. The ratios between the standard deviations and
the mean values are less, or even much less than 1%, except dwσ (3A/4)/hdw (3A/4)i ≈ 1.6% and
qinσ (z = A)/hqin (z = a)i ≈ 1.9% for GB (Fig. 5(c) and (f)) and dψσ (5A/8)/hdψ (5A/8)i ≈ 1.2% for
LB (Fig. 6(d)).

3.3. Effect of boundary conditions on heat transfer and fluid flow


Figure 7 displays the vertical component of the velocity in different horizontal sections of the
channel, for the pressure outlet condition Eq. (6c) and two pressure inlet conditions Eq. (5b) and
Eq. (5c).
Major differences are shown in the entrance section. Indeed, when the inlet pressure is based
on the local kinetic energy per volume unit (Fig. 7(a)), the velocity consists essentially of a flat
profile far from the solid walls and it turns out to be of parabolic shape when the pressure is
uniform in the entrance section (Fig. 7(b)). However, in both situations, the velocity profiles
become almost parabolic at the entrance of the heated region at z = A/4 (filled squares). The
occurrence of an already established isothermal flow in the aperture section of the channel (see
Fig. 7(b)) is counterintuitive and somewhat questionable. Although the issue of choosing boundary
conditions capable to mimic the correct physical conditions is out of the scope of the present
contribution, we can however notice that preliminary experimental measurements [32] performed
on inlet velocity seem rather indicate uniform profiles in agreement with Fig. 7(a) obtained with
BC. (5b). Substituting outlet BC. (6c) by BC. (6d) does not alter the inlet velocity profiles.
The development of the dynamic boundary layer along the left wall as well as the back flow
entering through the top section at z = A are clearly visible in Fig. 7 (open triangles). We can
also notice that the return flow penetrates deeper into the channel for LB-boundary conditions: at
z = 3A/4 (filled circles), the magnitude of the downward fluid flow is slightly larger in Fig. 7(a)
than in Fig. 7(b). This qualitative observation is confirmed by the accurate measurements of dw (z)
and dψ (z) which are reported in Fig. 9(c) and Fig. 9(d) and commented afterwards.
Figure 8 presents the pressure and vertical velocity profiles in the vertical median section be-
tween the parallel plates versus the vertical coordinate, for LB, GB, LB-0 and GB-0 problems.
From Fig. 8(a), we see that the pressure is first decreasing in the lower adiabatic region because of
the head loss occurring in the channel for isothermal fluid flows. Then, beyond z = A/4, the buoy-
ancy force becomes active and the slope of the pressure gradient gets reversed. Additional features
are deduced from the spatial evolutions of these curves. We clearly see that the pressures GB and
GB-0 are linearly decreasing in the lower adiabatic region when BC. (5c) is applied in the aperture
13
sections, what confirms the flow is dynamically established upstream the heated region (see also
Fig. 7(b)). Above z = 7A/8, the pressure curves show a "S"-shape behavior, all the more marked
than the pressures in the inlet/outlet sections depend on the kinetic energy per volume unit. The
occurrence of fully developed flows in the lower adiabatic regions for GB and GB-0 cases are also
confirmed by the constant vertical velocity profiles (Fig. 8(b)). The increases in w(0.5, z), shown
in Fig. 8(b) for LB and LB-0 situations, are related to the dynamic boundary layer developments
which slow down the flows close to the wall and accelerate the fluid motions at the center of the
channels. But once the heated regions are reached, thermal boundary layers along the left surfaces
grow and modify the kinetic boundary layers by increasing the fluid velocities due to buoyancy
forces (see also Fig. 7). Since the flow rates must be conserved, it results some decreases in the
vertical velocities at the centers of the channel.

Figure 9 gathers the whole statistical results for boundary conditions LB-0, LB, GB-0 and GB
together. Histograms and error bars stand for average and standard deviation values. Graphics are
then topped by the mean values hxi, plus the relative standard deviation error ±xσ /hxi reported
in parentheses and expressed in permillage (O/OO ). To emphasize the agreement or disagreement
between participants, these values are colored in blue if xσ /hxi < 1% and in red otherwise. As
announced in the first paragraph of Sec. 3.2, the reference solution is set to the median xm and
its value, accurate within four significant digits, is reported below the corresponding histogram.
Before considering the results in details, we can notice that xσ /hxi are rather small, what indicates
a good agreement between the results provided by the different partners.
Let us first analyze the thermal results. The comparison between the four studied cases LB-
0, LB, GB-0 and GB, for the average Nusselt number Nu1 (Fig. 9(f)) and Nu f 1 (z) (Fig. 9(a))
shows a maximal relative gap of 3.2%. However, when one considers the local Nusselt number
along the heated surface using the bulk temperature as reference, the relative difference increases
significantly with z (Fig. 9(b)): Nu2 (z)[LB-0]/Nu2 (z)[GB] − 1 is approximatively equal to 8% at
z = A/2, 14% at z = 5A/8, 23% at z = 3A/4. In the adiabatic region, this relative difference
f 2 (z)[LB-0]/Nu
Nu f 2 (z)[GB] − 1 grows dramatically from 114% at z = 7A/8 to more than 1000%
at z = A. Nonetheless, the average Nusselt value on the heated section Nu2 remains weakly
affected by the boundary condition changes since the relative difference does not exceed 8%.
The local bulk temperature is also very sensitive to the adopted model (Fig. 9(e)), but, contrary
to Nuf 2 (z), the relative difference is almost constant with z: θb (z)[LB-0]/θb (z)[GB] − 1 ≈ 43%.
This relative gap is easily explained by the following factors. Since the axial thermal diffusion
is essentially negligible with respect to the transverse diffusion, it is easy to prove that the bulk
temperature increases linearly in the heated section with a slope close to 1/qin (z = 0) and then it
keeps constant in the adiabatic region. Therefore, the departure in the local bulk temperature is
given by the relative difference obtained in the flow rate entering the channel at z = 0, namely
qin (z = 0)[GB]/qin (z = 0)[LB-0] − 1 ≈ 43% (Fig. 9(f)).
In the second step, let us focus on the dynamical aspects. A first insight shows that the flow pat-
terns generated by the natural convection as well as the flow rates are very sensitive to conditions
imposed at the inlet/outlet of the channel sections (Figs. 9(c),(d) and (f)). The maximal relative
gaps recorded at z = A lie from about 18% for dψ (z = A) and dw (z = A) up to 140% for qin (z = A).
A thoroughly study shows that applying uniform boundary conditions in the inlet section, namely
14
BC. (5c), creates larger flow rates qin (z = 0) than configurations where boundary conditions are
based on the local kinetic energy per volume unit (BC. 5b), i.e. qin (z = 0)[GB] > qin (z = 0)[LB]
and qin (z = 0)[GB-0] > qin (z = 0)[LB-0]. Furthermore, if we compare in Fig. 8(a) the mean pres-
sure drop in the entrance adiabatic region between GB and LB, and between GB-0 and LB-0, we
note that its value is larger if the inlet pressure is defined locally, what corresponds to the existence
of a dynamic entry zone where the fluid flow evolves from a flat profile to a Poiseuille-like profile
(Fig. 7(a)). We then deduce that the reduction in flow rate qin (z = 0) obtained with BC. (5b)
versus BC. (5c) is due to head losses induced by the fluid flow development in the lower adiabatic
region. Let us now consider the flow rates induced by the return flow for fixed inlet boundary
conditions. Fig. 9(f) indicates that qin (z = A) is reduced when the outlet pressure boundary con-
ditions depend on the fluid flow direction (BC. 6c), i.e. qin (z = A)[GB] < qin (z = A)[GB-0] and
qin (z = A)[LB] < qin (z = A)[LB-0]. The decrease in this flow rate is also correlated with the short-
ening of the return flow dw (Fig. 9(c)) and of the recirculation cell dψ (Fig. 9(d)). These behaviors
are directly explained by the intensity of the pressure imposed to the return flow, since it is smaller
with BC. (6c) than BC. (6d) and consequently less efficient to create an inflow circulation through
the upper aperture. If we now compare the total flow rate leaving the channel, namely
qout = qin (z = 0) + qin (z = A) (17)
we notice that qout [LB-0] < qout [LB] < qout [GB-0] < qout [GB], what is in accordance with the
behavior of qin (z = 0) (Fig. 9(f)). While the maximal relative difference of the flow rate entering
into the channel at z = 0 and defined by
qin (z = 0)[GB]
−1 (18)
qin (z = 0)[LB-0]
is about 43%, the value of expression (18) falls down to 10% if qin (z = 0) is substituted by qout
(Eq. 17) instead. Therefore, the flow rate leaving the channel through the top section is almost the
same whatever the boundary conditions are. This result is explained by the kind of flow crossing
the channel apertures. Indeed, this flow rate aims to feed the dynamic boundary layer created by
the wall heating and driven by the buoyancy force. And to do this, the fluid path fits to minimize
the total head loss as follows:
• the inflow rate qin (z = 0) (resp. qin (z = A)) is maximal (resp. minimal) through the lower
(resp. upper) aperture when both
– the velocity is fully developed all along the upstream adiabatic region (Eq. 5c), i.e. the
upstream head loss is minimal,
– the pressure at the outlet acts against the appearance of a return flow (Eq. 6c) by
producing a suction (negative pressure).
• the inflow rate qin (z = 0) (resp. qin (z = A)) is minimal (resp. maximal) through the lower
(resp. upper) aperture when both
– an entry length exists in the upstream adiabatic region (Eq. 5b), i.e. the upstream head
loss is maximal,
15
– the pressure at the outlet promotes the growth of a return flow (Eq. 6d) by pushing the
fluid into the channel (positive pressure).
This flow rate analysis confirms that the modification in the open boundary conditions acts
significantly on the dynamic variables and less on the heat transfer on the heated section of the
left plate. On the other hand, the Nusselt number Nu2 (z) clearly depends on the boundary condi-
tions in the downstream adiabatic region. This sensitivity is due to its dependency into the bulk
temperature θb (z) (or qin (z = 0)), which, in turn, is highly linked both to the flow structure and its
intensity.

4. Conclusion
A comparison exercise concerning the fluid flow and heat transfer in a vertical channel asym-
metrically heated at a constant heat flux has been initiated with the participation of eight teams
coming from different laboratories. The dimensionless parameters were chosen so that a return
flow occurs through the top section of the channel and four sets of conditions were proposed to
model the open boundaries.

Both local and overall quantities were extracted from the numerical results: two expressions of
the local Nusselt number along the heated plate, the local lengths of both the return flow and the
associated recirculation, the local bulk temperature and four overall values defined by the average
Nusselt numbers on the heated plate, the flow rate entering the channel through the bottom section
and through the top section of the channel. All data provided by the participants were illustrated
by histograms. The average, standard deviation and median statistical quantities were computed
and reported on the figures. Comparisons show a relative good agreement between the different
contributions.
For each of the four boundary conditions and the whole recorded variables, reference solutions
have been defined. These numerical values have been gathered and copied out on Fig. 9(a)-(f),
below their corresponding histograms. For benchmarking purpose, we recommend to compare at
least the flow rates qin (z = 0) and qin (z = A), and the different lengths of the recirculation dw (z)
and dψ (z) because these quantities are much sensitive to the boundary conditions than the thermal
variables.

The effects of boundary conditions on fluid flow and heat transfer have been discussed. Whereas
the local and mean Nusselt numbers based on the inlet temperature do not depend on the boundary
conditions, the Nusselt number constructed on the bulk temperature reveals to be much sensitive
to the conditions applied at the apertures. This large sensitivity is related to changes in flow pat-
terns and flow rates. Just as a uniform zero pressure at the upper section of the channel promotes
the return flow and its intensity, a constant pressure boundary condition to the lower aperture im-
prove the inward flow rate. But whatever the air supply through the open sections, the fluid flow
rates produced by the boundary layers along the heated surface are almost constant, what is in
good agreement with the weak influence of the boundary conditions on heat transfer. Finally, we
showed that the fluid flow is organized to minimize the total head loss.

16
Even if the issue of finding the correct boundaries is out of the scope of this paper, the fully
developed flow which appears in the lower adiabatic region when an uniform pressure is imposed
in the lower open section is somewhat questionable. Therefore, local pressure (BC. 5b) at the
lower aperture will be preferred to (BC. 5c) since it creates a non zero dynamic entry length.

Appendix
This exercise was originally organized in a french workshop for the Thermal French Society
(SFT) in May 2004 by G. Desrayaud [24] and recently readdressed by the French Research Group
AmeTH [33].
The main characteristics of the numerical schemes developed and/or used by the different
teams are detailed below.

T1 The CFD code used by PPRIME is FDS [34]. The model solves numerically a form of
the Navier-Stokes equations appropriate for low-speed, thermally-driven flow. The partial
derivatives of the conservation equations of mass, momentum and energy are approximated
as finite differences, and the solution is updated in time on a three-dimensional, rectilinear
grid. The core algorithm is an explicit predictor-corrector scheme that is second order accu-
rate in space and time. It approximates the governing equations on one or more rectilinear
grids. The mesh used in this laminar 2D case is 400 × 2560 and the boundary conditions
are exactly those describes in the bench excepted at the inlet/outlet where FDS used the
"OPEN" boundary condition [34]. This condition assumes that the pressure perturbation e p
is zero at an outgoing boundary and e p = −ρ∞ |u|2 /2 at an incoming boundary. Notice that
these boundary conditions are close to GB-0 boundary conditions described in this paper.
The second difference with GB-0 is, as mentioned above, that the low Mach approximation
is used instead of the Boussinesq approximation to solve the governing equations.

T2 Equations (1), (2) and (3) were solved either in their steady form for LB and LB-0 problems,
or in their transient form with a Crank-Nicolson scheme for cases GB and GB-0. Spatial
derivatives are expressed with a finite volume scheme with variables located at the center of
the control volumes. Centered approximations are used for convective, advective and diffu-
sive fluxes. A coupling between the mass equation (Eq. 1) and the pressure is enforced by
applying a penalization technique on a coarser grid. The momentum nonlinear contribution
and the pressure gradient are discretized so that the discrete kinetic energy balance mim-
ics the behavior of its continuous counterpart; a similar condition is required for thermal
equation. The discrete Navier-Stokes and energy equations, expressed in their stationnary
form or at the new time step, are solved simultaneously in a large nonlinear system by the
Newton-Raphson method. An under-relaxed procedure is used when time derivatives are
dropped. Several meshes have been considered and the grid 253 × 1520 was finally retained.
Additional details upon the numerical scheme and validation procedures can be found in
[35–37].

T3 The Navier-Stokes equations under the Boussinesq assumptions are discretized by finite dif-
ference schemes. After time discretization, we obtain a system of Helmholtz equations. The
17
velocity-pressure coupling is solved using a projection method which leads to a Poisson
equation for the pressure correction. The variables (u, w, p, T ) are located on a staggered
mesh and a spatial discretization scheme of second order is used. The Helmholtz and Pois-
son equations are solved respectively using TDMA and partial diagonalization. Moreover,
the Poiseuille mode corresponding to the kernel of the Stokes operator is used to correct
velocity and pressure to satisfy the boundary condition GB-0 at the inlet of the channel
[25, 38]. To implement (GB, LB, LB-0), at the projection step, we impose Dirichlet bound-
ary conditions on pressure at the inlet and outlet of the channel. Note that GB-0 is verified
at each time step and GB, LB, LB-0 are verified only when the flow reaches the steady state
[38].

T4 The usual dimensionless Boussinesq 2D Navier-Stokes equations were used. The time deriva-
tives in the momentum and in the energy equations are performed by a second-order back-
ward differentiation. The convection terms are approximate using a second-order Adams-
Bashford extrapolation method. The diffusion terms are implicitly treated. The resulting
Helmholtz systems are solved by a direct solver. Pressure-velocity coupling is obtained by
an incremental rotational projection method. A collocated finite volume method has been
used. Details on numerical method can be found in [39]. The GB-0 case has been computed
with a 48 × 600 grid size. The numerical code has been developed thanks to the environment
OpenFOAM [40].

T5 The structure and solver of the computational code Aquilonrare issued from previous works
[41], originally implemented with a Navier-Stokes finite volumes solver on the staggered
MAC mesh and using the Uzawa augmented Lagrangian [42] method to deal with the
divergence-free constraint. The discretization method with Discrete Operator Calculus is
an extension of the MAC (Marker And Cell) method with staggered grids to unstructured
meshes. The method is similar to Discrete Exterior Calculus based on differential geometry.
The primal and dual meshes enable to express gradient, divergence, curl operators as well
as Green, Gauss and Stokes theorems to obtain the continuum properties ∇ · (∇ ∧ ~v) = 0
and ∇ ∧ (∇ f ) = ~0 up to machine precision. The scheme is based on a node-center approach
avoiding interpolations, where the scalar or vector components unknowns are distributed
on nodes, faces and edges of the mesh stencils. The discretization is shown to locally and
globally conserve up to machine precision, mass, kinetic energy and vorticity in the absence
of viscosity. The spatial accuracy is found to be second-order on a structured or unstruc-
tured mesh including highly irregular meshes [43]. All the linear algebraic systems for the
three methods, e.g. the prediction steps, are solved with the Krylov BiCGS tab2 algorithm
preconditionned by the incomplete LU factorization of order zero ILU(0) [41–43]

T6 The code solves the unsteady two-dimensional Boussinesq equations in their elliptic form
and has been developing at LIMSI. The temporal scheme adopted in order to discretize the
Boussinesq equations is based on the second-order backward-Euler scheme for time ad-
vance and on the Adams-Bashforth extrapolation for convective terms. Thus, the diffusion
terms are treated implicitly and the convective ones explicitly. The continuous problem is
reduced to a discrete problem thanks to a finite volume method. The marker-and-cell (MAC)
18
type staggered grid arrangement is used. All variables are spatially discretized with a sec-
ond order centered scheme. The pressure-velocity coupled problem is solved by using a
prediction-projection algorithm. The Poisson equation, derived from the continuity equa-
tion, is associated to homogeneous Neumann boundary condition at the whole boundaries
(walls and artificial boundaries) [44]. Helmholtz systems are solved for the velocity fields
using a GMRES algorithm and the Poisson equation is solved with a multi-grid algorithm.
The governing equations have been integrated in time from fluid at rest up to reach the steady
state. The size of the kernel for the unsteady Stokes operator, obtained by a singular value
decomposition using the Lapack library, is equal to two. The two modes associated are the
trivial (zero velocity, constant pressure) mode and a Poiseuille mode. The superposition
principle [25, 30, 31] has been used to obtain the numerical solution as a linear combination
of a particular solution find with the finite volume scheme and both modes of the Stokes
operator kernel. The result is obtained for a 256 × 2048 grid.

T7 The spatial discretization of the channel is achieved within a Finite Volume Method. The
numerical code used is the commercial FLUENT code 12. Grid sensitivity has been studied
an a non-uniform grid (50 × 300) was retained, with a refinement near the vertical walls as
well as at the inlet and outlet boundaries. Numerical results were obtained from a pressure-
based segregated solver, under Boussinesq approximation. Pressure-velocity coupling was
solved using the SIMPLE algorithm. The Navier-Stokes equations were solved using, for
the pressure discretization, the PRESTO (PRESsure STaggering Option) scheme, and for
the momentum and energy equations, the QUICK scheme. The diffusion terms are central-
differenced and are second order accurate.
Note that the dynamic boundary conditions are slightly different from LB boundary condi-
tions. Let us note ~n the unit normal vector pointing outward. Then, for fluid flows

• entering the channel, ~v · ~n < 0, u = 0 and (∂w)/(∂z) = −(∂u)/(∂x);


• leaving the channel, ~v · ~n > 0, (∂u)/(∂z) = 0 and (∂w)/(∂z) = −(∂u)/(∂x).

The other boundary conditions are the same as LB. It is worth to point out that (∂u)/(∂x) is
relatively small in the inlet/outlet sections, and then (∂w)/(∂z) ≈ 0. Considering the no slip
boundary conditions at walls and (∂u)/(∂x) ≈ 0, we recover u ≈ 0, namely the boundary
conditions for LB.

T8 Both spatial and temporal discretization and the resolution procedure of discretized equations
are identical to T6. The marker-and-cell (MAC) type staggered grid arrangement is used.
The artificial boundaries of the computational domain which correspond to the inlet and the
outlet of the channel are considered to be on the scalar pressure nodes. The pressure equation
is associated to homogeneous Neumann boundary condition at the wall and Dirichlet bound-
aries conditions at the inlet and at the outlet so as to satisfy the local or the global Bernoulli
condition for pressure. The four set of boundaries condition proposed in the benchmark
have been treated. The whole results are obtained for a 128 × 1300 grid.

19
References
[1] E. Bacharoudis, M. Vrachopoulos, M. Koukou, D. Margaris, A. Filios, S. Mavrommatis, Study of the natural
convection phenomena inside a wall solar chimney with one wall adiabatic and one wall under a heat flux,
Applied Thermal Eng. 27 (2007) 2266–2275.
[2] F. Incropera, Convection heat transfer in electronic equipment cooling, J. Heat Transfer 110 (1988) 1097–1111.
[3] R. Ben Yedder, E. Bilgen, Natural convection and conduction in trombe wall systems, Int. J. Heat Mass Transfer
34 (1991) 1237–1248.
[4] B. Balunov, A. Babykin, R. Rybin, B. Krylov, V. Tanchuck, S. Grigoriev, Heat transfer at mixed convection in
vertical and inclined flat channels of the vacuum chamber of the iter international thermonuclear reactor, High
Temperature 42 (2004) 126–133.
[5] W. Elenbaas, Heat dissipation of parallel plates by free convection, Physica 9 (1) (1942) 1–23.
[6] H. Sun, R. Li, E. Chénier, G. Lauriat, On the modeling of aiding mixed convection in vertical channels, Heat
Mass Transfer (2012) 1–10.
[7] H. Sun, R. Li, E. Chénier, G. Lauriat, J. Padet, Optimal plate spacing for mixed convection from an array of
vertical isothermal plates, International Journal of Thermal Sciences 55 (2012) 16 – 30.
[8] G. Desrayaud, A. Fichera, Laminar natural convection in a vertical isothermal channel with symmetric surface-
mounted rectangular ribs, Int. J. Heat and Mass Transfer 23 (2002) 519–529.
[9] J. Hernandez, B. Zamora, Effects of variable properties and non-uniform heating on natural convection flows in
vertical channels, Int. J. Heat and Mass Transfer 48 (2005) 793–807.
[10] Q. Lu, S. Qiu, G. Su, W. Tian, Z. Ye, Experimental research on heat transfer of natural convection in vertical
rectangular channels with large aspect ratio, Experimental Thermal and Fluid Science 34 (1) (2010) 73 – 80.
[11] K. Ong, C. Chow, Performance of a solar chimney, Solar Energy 74 (1) (2003) 1 – 17.
[12] D. Ospir, C. Popa, C. Chereches, G. Polidori, S. Fohanno, Flow visualization of natural convection in a vertical
channel with asymmetric heating, International Communications in Heat and Mass Transfer 39 (4) (2012) 486–
493.
[13] A. Andreozzi, B. Buonomo, O. Manca, Thermal management of a symmetrically of a heated channel-chimney
system, International Journal of Thermal Sciences 48 (9) (2009) 742–748.
[14] J. Bodoia, J. Osterle, The development of free convection between heated vertical plates, ASME J. Heat Transfer
84 (1962) 40–44.
[15] T. Aihara, Effects of inlet boundary conditions on numerical solutions of free convection between vertical par-
allel plates, Report of the Institute of High Speed Mechanics, Tohuku University, Japan 28 (1973) 1–27.
[16] A. Dalbert, F. Penot, J. Peube, Convection naturelle laminaire dans un canal vertical chauffé à flux constant, Int.
J. Heat Mass Transfer 24 (1981) 1463–1473.
[17] C. Kettleborough, Transient laminar free convection between heated vertical plates including entrance effets,
Int. J. Heat Mass transfer 15 (1972) 883–896.
[18] D. Naylor, J. Floryan, J. Tarasuk, A numerical study of developing free convection between isothermal vertical
plates, J. Heat Transfer 113 (1991) 620–626.
[19] H. Nakamura, Y. Asako, T. Naitou, Heat transfer by free convection between two parallel flat plates, Num Heat
Transfer 5 (1982) 95–106.
[20] S. Boetcher, E. Sparrow, Buoyancy-induced flow in an open-ended cavity: Assessment of a similarity solution
and of numerical simulation models, International Journal of Heat and Mass Transfer 52 (15–16) (2009) 3850 –
3856.
[21] A. Andreozzi, B. Buonomo, O. Manca, Thermal and fluid dynamic behaviors in symmetrical heated channel-
chimney systems, International Journal of Numerical Methods for Heat and Fluid Flow 20 (7) (2010) 811–833.
[22] S. Ramanathan, R. Kumar, Correlations for natural convection between heated vertical plates, J. Heat Transfer
113 (1991) 97–107.
[23] J. Liu, W. Tao, Numerical analysis of natural convection around a vertical channel in a rectangular enclosure,
Heat and Mass Transfer 31 (1996) 313–321.
[24] G. Desrayaud, R. Bennacer, J.-P. Caltagirone, E. Chénier, A. Joulin, N. Laaroussi, K. Mojtabi, Etude numérique
comparative des écoulements thermoconvectifs dans un canal vertical chauffé asymmétriquement, in: VIIIième
Colloque Inter universitaire Franco-Québécois sur la Thermique des Systèmes, ART-06-14, 2007.
20
[25] P. Le Quéré, On the computation of some external or partially enclosed natural convection flows (keynote paper),
in: In 19th Int. Symp. Transport Phenomena, ISTP19, 2008, 8 pages, Reykjavik, Iceland.
[26] B. Webb, D. Hill, High Rayleigh number laminar natural convection in an asymmetrically heated vertical chan-
nel, J. Heat Transfer 111 (1989) 649–656.
[27] E. Sparrow, G. Chrysler, L. Azevedo, Observed flow reversals and measured-predicted nusselt numbers for
natural convection in a one-sided heated vertical channel, ASME J. Heat Transfer 106 (1984) 325–334.
[28] R. Li, H. Sun, E. Chénier, G. Lauriat, Effet du rayonnement surfacique sur les recirculations de convection
naturelle, in: Proc. of the annual Congrès Français de Thermique : Energie Solaire et Thermique, Vol. 19,
Perpignan, 2011, pp. 181–186.
[29] R. Li, M. Bousetta, E. Chénier, G. Lauriat, Effect of surface radiation on natural convective flows and onset of
flow reversal in asymmetrically heated vertical channels, International Journal of Thermal Sciences 65 (2013)
9–27.
[30] P. Le Quéré, On the computation of some external or partially enclosed natural convection flows, in: 7th Inter-
national Conference on Computational Heat and Mass Transfer, 2011, paper 401, Istanbul, Turkey.
[31] C. Garnier, A. Sergent, P. Le Quéré, Modélisation numérique d’un canal vertical asymétriquement chauffé avec
une condition aux limites à l’entrée de type vorticité, in: Proc. of the annual Congrès Français de Thermique,
SFT Bordeaux, 2012.
[32] Y. Cherif, A. Joulin, L. Zalewski, S. Lassue, Outils pour l’étude expérimentale de la convection naturelle entre
plaques planes verticales différentiellement chauffées, in: Proc. of the annual Congrès Français de Thermique,
Vol. 18, 2010.
[33] E. Arquis, Rapport intermédiaire GDR AMETh: Simulation numérique de l’écoulement d’air dans un canal
vertical partiellement chauffé. Exercice de comparaison (Mai 2010).
[34] Http://www.fire.nist.gov/fds.
[35] E. Chénier, R. Eymard, O. Touazi, Numerical results using a colocated finite volume scheme on unstructured
grids for incompressible fluid flows, Numerical Heat Transfer Part B 49 (3) (2006) 259–276.
[36] O. Touazi, E. Chénier, R. Eymard, Simulation of natural convection with the collocated clustered finite volume
scheme, Computers & Fluids 37 (2008) 1138–1147.
[37] E. Chénier, R. Eymard, R. Herbin, A collocated finite volume scheme to solve free convection for general
non-conforming grids, Journal of Computational Physics 228 (2009) 2296–2311.
[38] Z. Amine, S. Xin, S. Giroux-Julien, C. Ménézo, étude numérique des transferts thermo-convectifs dans un canal
d’air vertical à flux imposé, in: Proc. of the annual Congrès Français de Thermique, SFT Perpignan, 2011, pp.
163–168.
[39] R. Peyret, T. D. Taylor, Computational methods for fluid flow, Springer Series in Computational Physics, 1983.
[40] OpenFOAM 2.1, http ://www.openfoam.com, 2012.
[41] S. Vincent, J.-P. Caltagirone, P. Lubin, T. N. Randrianarivelo, An adaptative augmented lagrangian method for
three-dimensional multimaterial flows, Computers & Fluids 33 (2004) 1273–1289.
[42] M. Fortin, R. Glowinski, Méthodes de Lagrangien augmenté. Application à la résolution numérique de prob-
lèmes aux limites. Paris, Dunod, Dunod, 1982.
[43] P. Angot, J. P. Caltagirone, P. Fabrie, A fast vector penalty-projection method for incompressible nonhomoge-
neous or multiphase Navier-Stokes problems, Applied Mathematics Letters 25 (11) (2012) 1681–1688.
[44] E. Gadoin, P. Le Quére, O. Daube, A general methodology for investigating flow instabilities in complex ge-
ometries : application to natural convection in enclosure, International Journal for Numerical Methods in Fluids
37 (2001) 175–208.

21
List of Tables
1 Notations for constructing the boundary conditions. . . . . . . . . . . . . . . . . 24
2 Some characteristics of the numerical codes and labels of teams. . . . . . . . . . 24

22
List of Figures
1 Geometry and boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . 25
2 Fluid flow and temperature field for LB boundary conditions (see Tab. 1). The ver-
tical thick line located on the left wall emphasizes the heated section. (a) Stream-
lines and size of the recirculation at the top section, dψ (z = A) (Eq. 14); (b)
Horizontal component of the velocity, u(x, z); (c) Vertical component of the veloc-
ity, w(x, z), and size of the downward flow at the top section, dw (z = A) (Eq. 13);
(d) Temperature field. Data provided by team T2. . . . . . . . . . . . . . . . . . 26
3 Histograms of results obtained for GB-0 (Tab. 1). Data provided by the different
teams (Tab. 2) and statistical quantities: average, standard deviation and median
values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4 Histograms of results obtained for LB-0 (Tab. 1). Data provided by the different
teams (Tab. 2) and statistical quantities: average, standard deviation and median
values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5 Histograms of results obtained for GB (Tab. 1). Data provided by the different
teams (Tab. 2) and statistical quantities: average, standard deviation and median
values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
6 Histograms of results obtained for LB (Tab. 1). Data provided by the different
teams (Tab. 2) and statistical quantities: average, standard deviation and median
values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
7 Vertical component of the velocity at different horizontal sections of the channel
for (a) LB and (b) GB boundary conditions (Tab. 1). The vertical thick line located
on the left wall emphasizes the heated section. Data provided by team T2. . . . . 33
8 (a) Pressure and (b) vertical velocity profiles as a function of z at mid-width of the
channel for boundary conditions LB, GB, LB-0 and GB-0 (Tab. 1). Data provided
by team T8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
9 Histograms and error bars stand for average values (hxi) and standard deviations
(xσ ), for LB-0, LB, GB-0 and GB (Tab. 1). Boxes are topped by the average values
and the relative standard deviation (xσ /hxi) reported in parentheses and expressed
in permillage. Values below the graphics are the reference solutions which are set
to the median values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

23
Channel Inlet
Relation 5b Relation 5c
Channel Relation 6c LB GB
Outlet Relation 6d LB-0 GB-0

Table 1: Notations for constructing the boundary conditions.

Team Laboratory Code Grid Stationary scheme Boundary conditions


T1 PPRIME FDS [34] 400 × 2560 no GB-0
T2 MSME, LAMA Laboratory code [35–37] 253 × 1520 yes/no GB, GB-0, LB, LB-0
T3 CETHIL Laboratory code [25, 38] 200 × 1200 no GB, GB-0, LB, LB-0
T4 PIMENT, LaSIE Laboratory code [39] (with OpenFoam [40]) 48 × 600 no GB-0
T5 I2M–TREFLE Aquilonr[41–43] 100 × 1000 no GB-0
T6 LIMSI Laboratory code [25, 30, 31] 256 × 2048 no GB-0
T7 LGCgE, LEME Fluentr 50 × 300 yes LB
T8 LIMSI Laboratory code 128 × 1300 no GB, GB-0, LB, LB-0

Table 2: Some characteristics of the numerical codes and labels of teams.

24
∂w
u= =0 
∂z 

 −0.5w2 (Eq. 6c)
∂θ 
if w > 0: = 0, P = 0 if w < 0: θ = 0, p = 
 or
∂z 
 0 (Eq. 6d)

∂θ
=0 A/4
∂x

∂θ 1 ∂θ
= −1 =0
∂x ∂x
A/2
w, ~ez

∂θ
=0 A/4 u, ~ex
∂x




 −0.5w2 (Eq. 5b)
∂w 
u= = 0, θ = 0, p = 
 or
∂z 
 −0.5(q (z = 0))2 (Eq. 5c)
in
Figure 1: Geometry and boundary conditions

25
dψ(z=A) dw(z=A)

(a) (b) (c) (d)


Figure 2: Fluid flow and temperature field for LB boundary conditions (see Tab. 1). The vertical
thick line located on the left wall emphasizes the heated section. (a) Streamlines and size of the
recirculation at the top section, dψ (z = A) (Eq. 14); (b) Horizontal component of the velocity,
u(x, z); (c) Vertical component of the velocity, w(x, z), and size of the downward flow at the top
section, dw (z = A) (Eq. 13); (d) Temperature field. Data provided by team T2.

26
(Tab. 2) and statistical quantities: average, standard deviation and median values.
Figure 3: Histograms of results obtained for GB-0 (Tab. 1). Data provided by the different teams

20

40

60

80

10

20
0

0
8.313 7.342
8.256 xm=8.255 7.284 xm=7.286

z=3A/8

z=3A/8
8.254 7.286
8.255 <x>=8.257 7.269 <x>=7.291
8.232 xσ=2.6E-02 7.266 xσ=2.4E-02
8.228 7.303

T2
T1

T2
T1
8.262 7.290
7.881 6.284
7.807 xm=7.807 6.236 xm=6.238
z=A/2

z=A/2
7.802 6.238
7.812 <x>=7.805 6.227 <x>=6.242
7.786 xσ=4.0E-02 6.221 xσ=1.9E-02
(b) Nu

(a) Nu
7.735 6.250
7.810 6.240

f 1 (z), Nu1 (z ≤ 3A/4)


7.958 5.740
f 2 (z), Nu2 (z ≤ 3A/4)

7.853 5.691

T4
T3

T4
T3
xm=7.853 xm=5.693
z=5A/8

z=5A/8
7.845 5.693
7.865 <x>=7.849 5.684 <x>=5.698
7.834 xσ=6.0E-02 5.678 xσ=1.9E-02
7.735 5.703
27

7.855 5.694
9.049 5.712
8.917 xm=8.917 5.665 xm=5.665
z=3A/4

z=3A/4
8.890 5.664
8.947 <x>=8.877 5.501 <x>=5.638
8.606 xσ=1.5E-01 5.534 xσ=7.9E-02

T6
T5

T6
T5
8.700 5.676
9.029 5.713
27.87 10.04
27.22 xm=27.22 9.886 xm=9.886
z=7A/8

z=7A/8
27.09 9.891
27.37 <x>=27.06 9.879 <x>=9.912

median: xm

median: xm
27.30 xσ=7.3E-01 9.877 xσ=5.4E-02
25.37 9.926
27.18 9.884
54.18 12.22
T8

T8
51.76 xm=51.76 11.95 xm=11.95
51.31 11.95
z=A

z=A
52.19 <x>=52.14 11.94 <x>=12.04
58.55 xσ=3.6E+00 12.26 xσ=1.3E-01
45.38 12.00
51.58 11.94
values.
by the different teams (Tab. 2) and statistical quantities: average, standard deviation and median
Figure 3i (continued Fig. 3): Histograms of results obtained for GB-0 (Tab. 1). Data provided

0.2

0.4

0.6

0.8

0.2

0.4

0.6

0.8
0

1
z=3A/8

z=3A/8
T2
T1

T2
T1
z=A/2

z=A/2
0.2919 0.2150
0.1334 0.08906

T4
T3

T4
T3
xm=0.1334 xm=0.09023
z=5A/8

z=5A/8
0.1330 0.09200
(d) dψ (z)

(c) dw (z)
0.1380 <x>=0.1432 0.09023 <x>=0.09890
0.1344 xσ=6.9E-02 0.09024 xσ=5.2E-02
28

0.0392 0.02596
0.1326 0.08984
0.5857 0.4000
0.5392 xm=0.5392 0.3713 xm=0.3713
z=3A/4

z=3A/4
0.5380 0.3700
0.5456 <x>=0.5369 0.3709 <x>=0.3683
0.5405 xσ=3.1E-02 0.3723 xσ=2.1E-02

T6
T5

T6
T5
0.4705 0.3219
0.5389 0.3714
0.6743 0.4725
0.6652 xm=0.6652 0.4641 xm=0.4640
z=7A/8

z=7A/8
0.6610 0.4630
0.6670 <x>=0.6618 0.4637 <x>=0.4618

median: xm

median: xm
0.6657 xσ=1.2E-02 0.4646 xσ=9.2E-03
0.6349 0.4404
0.6648 0.4640
0.7049 0.5200
T8

T8
0.7068 xm=0.7049 0.5020 xm=0.5018
0.7000 0.5030
z=A

z=A
0.7080 <x>=0.7022 0.4987 <x>=0.5007
0.7072 xσ=7.8E-03 0.5001 xσ=1.1E-02
0.6841 0.4794
0.7047 0.5018
values.
by the different teams (Tab. 2) and statistical quantities: average, standard deviation and median
Figure 3ii (continued Fig. 3): Histograms of results obtained for GB-0 (Tab. 1). Data provided

100

120

0.02

0.04

0.06

0.08

0.12
20

40

60

80

0.1
0

0
7.000
0.01590
6.924 0.01617
xm=6.947 xm=0.01614

z=3A/8
6.930 0.01610
0.01630 <x>=0.01602
Nu1

6.947 <x>=6.955 0.01615

T2
T1
xσ=2.8E-04

T2
T1
6.980 0.01539
xσ=2.5E-02 0.01614
6.960
(f) Nu1 , Nu2 , qin (z = 0) and qin (z = A)

0.03236
6.941 0.03226 xm=0.03226

z=A/2
0.03213
8.584 0.03240 <x>=0.03205
0.03230 xσ=5.5E-04
8.483 0.03072
xm=8.484 0.03221
8.482
Nu2


8.519 <x>=8.498 0.04693

T4
T3

T4
T3
0.04837 xm=0.04831

z=5A/8
8.484 0.04818
xσ=4.3E-02

(e) θb (z)
8.430 0.04860 <x>=0.04784
0.04846 xσ=8.9E-04
29

8.506 0.04605
0.04831
77.63 0.06454
0.06431 xm=0.06431

z=3A/4
77.60 0.06407
xm=77.63
qin(z=0)

77.80 0.06470 <x>=0.06395


0.06451

T6
T5
77.43 <x>=78.17 xσ=1.1E-03

T6
T5
0.06124
77.66 0.06425
xσ=1.4E+00
81.50 0.06374
77.59 0.06442 xm=0.06439

z=7A/8
0.06420
0.06460 <x>=0.06390

median: xm

median: xm
16.62 0.06462 xσ=1.1E-03
0.06133
17.52 0.06439
xm=17.52
qin(z=A)

17.61 0.06339
T8

T8
17.70 <x>=17.11 0.06438 xm=0.06437
0.06417

z=A
17.72 0.06460 <x>=0.06382
xσ=8.9E-01
15.11 0.06452 xσ=1.1E-03
0.06130
17.52 0.06437
(Tab. 2) and statistical quantities: average, standard deviation and median values.
Figure 4: Histograms of results obtained for LB-0 (Tab. 1). Data provided by the different teams

10

20
0.2

0.4

0.6

0.8
0.02

0.04

0.06

0.08

0.12

0
0.1

1
0
7.191 xm=7.198

z=3A/8
0.0210

z=3A/8
xm=0.0209

z=3A/8
7.198 <x>=7.196
0.0206 <x>=0.0208 xσ=3.3E-03
7.198

T2
T2
xσ=1.9E-04

T2
0.0209

(a) f
6.166 xm=6.170

z=A/2
0.0419 xm=0.0418

z=A/2
z=A/2
6.172 <x>=6.169

Nu1 (z),
0.0410 <x>=0.0416 xσ=2.4E-03
xσ=3.8E-04 6.170
0.0418

5.624

T3
0.4144 xm=5.627

T3
(c) dw (z)

z=5A/8
xm=0.4141

T3
0.0627

(e) θb (z)

z=5A/8
xm=0.0627

z=5A/8
0.3980 <x>=0.4088 5.630 <x>=5.627
0.0615 <x>=0.0623

Nu1 (z
xσ=7.6E-03 xσ=2.4E-03
xσ=5.7E-04 0.4141 5.627
0.0627

0.5157 5.600 xm=5.602


0.0834 xm=0.5157

z=3A/4
z=3A/4
z=3A/4 xm=0.0834
0.0818 <x>=0.0829 0.5080 <x>=0.5132 5.602 <x>=5.616

≤ 3A/4)

T8
xσ=3.7E-03 xσ=2.2E-02

T8
xσ=7.5E-04 5.647

T8
0.0834 0.5159

0.0836 0.5424 9.692 xm=9.692


xm=0.5424

z=7A/8
xm=0.0836

z=7A/8
z=7A/8

0.0820 <x>=0.0830 0.5380 <x>=0.5409 9.707 <x>=9.697

median: xm
median: xm
median: xm
xσ=7.5E-04 xσ=2.1E-03 xσ=7.4E-03
0.0836 0.5424 9.690

0.0835 xm=0.0835 0.5552 xm=0.5530 11.68 xm=11.68

z=A
z=A
z=A

0.0819 <x>=0.0830 0.5530 <x>=0.5532 11.70 <x>=11.69


xσ=7.5E-04 xσ=1.5E-03 xσ=1.1E-02
0.0835 0.5516 11.68
30

100

120

100

200

300

400

500

600

700
0.2

0.4

0.6

0.8
20

40

60

80
0

0
(f) Nu1 , Nu2 , qin (z = 0) and qin (z = A)

8.469 xm=8.469

z=3A/8

z=3A/8
6.828 xm=6.840 8.449 <x>=8.464
Nu1


6.840 <x>=6.838 T2 xσ=1.1E-02

T2

T2
xσ=6.9E-03 8.475

(b) f
6.845
8.311 xm=8.311
z=A/2

z=A/2
Nu2 (z),
8.265 <x>=8.298
xσ=2.3E-02
8.316
9.005 xm=9.005
Nu2


8.959 <x>=8.998 0.5923 8.690


T3

T3

T3
(d) dψ (z)

xm=0.5912 xm=8.690
z=5A/8

z=5A/8
xσ=2.9E-02
9.029 0.5670 <x>=0.5835 8.612 <x>=8.666

Nu2 (z
xσ=1.2E-02 xσ=3.8E-02
0.5912 8.694

0.7257 xm=0.7257 10.52 xm=10.52


z=3A/4

z=3A/4
59.82 xm=59.82 0.7120 <x>=0.7211 10.34 <x>=10.51
qin(z=0)

≤ 3A/4)
xσ=6.5E-03 xσ=1.4E-01
T8

T8

T8
60.90 <x>=60.17 0.7257 10.67
xσ=5.2E-01
59.78
0.7646 xm=0.7644 50.98 xm=50.95
z=7A/8

z=7A/8
0.7530 <x>=0.7607 47.52 <x>=49.82
median: xm

median: xm

median: xm
xσ=5.4E-03 xσ=1.6E+00
0.7644 50.95
28.87
qin(z=A)

xm=28.87
28.22 <x>=28.65 0.7784 xm=0.7775 474.5 xm=471.0
z=A

z=A
xσ=3.1E-01 0.7680 <x>=0.7746 283.0 <x>=409.5
28.87
xσ=4.7E-03 xσ=9.0E+01
0.7775 471.0
(Tab. 2) and statistical quantities: average, standard deviation and median values.
Figure 5: Histograms of results obtained for GB (Tab. 1). Data provided by the different teams

10

20
0.2

0.4

0.6

0.8
0.02

0.04

0.06

0.08

0.12

0
0.1

1
0
7.325 xm=7.325

z=3A/8
0.01463

z=3A/8
xm=0.01463

z=3A/8
7.324 <x>=7.327
0.01470 <x>=0.01464 xσ=3.4E-03
7.332

T2
T2
xσ=4.1E-05

T2
0.01460

(a) f
6.265 xm=6.265

z=A/2
0.02919 xm=0.02919

z=A/2
z=A/2
6.264 <x>=6.266

Nu1 (z),
0.02933 <x>=0.02922 xσ=2.4E-03
xσ=8.2E-05 6.270
0.02914

5.716

T3
xm=5.716

T3
(c) dw (z)

z=5A/8
T3
0.04375

(e) θb (z)

z=5A/8
xm=0.04375

z=5A/8
5.716 <x>=5.717
0.04399 <x>=0.04381

Nu1 (z
xσ=1.7E-03
xσ=1.2E-04 5.720
0.04370

0.2591 5.694 xm=5.694


0.05820 xm=0.2591

z=3A/4
z=3A/4
z=3A/4 xm=0.05820
0.05851 <x>=0.05828 0.2680 <x>=0.2620 5.689 <x>=5.708

≤ 3A/4)

T8
xσ=4.2E-03 xσ=2.4E-02

T8
xσ=1.6E-04 5.742

T8
0.05814 0.2590

0.05828 0.4106 9.967 xm=9.965


xm=0.4106

z=7A/8
xm=0.05828

z=7A/8
z=7A/8

0.05862 <x>=0.05838 0.4140 <x>=0.4117 9.962 <x>=9.965

median: xm
median: xm
median: xm
xσ=1.7E-04 xσ=1.6E-03 xσ=2.2E-03
0.05825 0.4104 9.965

0.05823 xm=0.05823 0.4758 xm=0.4758 12.06 xm=12.05

z=A
z=A
z=A

0.05859 <x>=0.05835 0.4780 <x>=0.4765 12.05 <x>=12.05


xσ=1.7E-04 xσ=1.1E-03 xσ=4.2E-03
0.05822 0.4756 12.05
31

20

40

60

80
100

120

0.2

0.4

0.6

0.8

0
20

40

60

80
0

1
(f) Nu1 , Nu2 , qin (z = 0) and qin (z = A)

8.204 xm=8.207

z=3A/8
z=3A/8
6.968 xm=6.969 8.207 <x>=8.207
Nu1


6.969 <x>=6.974 xσ=2.6E-03


8.210

T2
T2

T2
xσ=7.9E-03

(b) f
6.985
7.667 xm=7.671

z=A/2
z=A/2
7.674 <x>=7.671

Nu2 (z),
xσ=2.8E-03
7.671
8.351 xm=8.362
Nu2


7.623

T3
8.362 <x>=8.363 xm=7.626
T3

T3
(d) dψ (z)

z=5A/8
z=5A/8

xσ=9.3E-03 7.635 <x>=7.628


8.374

Nu2 (z
xσ=5.1E-03
7.626

0.3811 8.519 xm=8.528


xm=0.3826

z=3A/4
z=3A/4

85.77 xm=85.77 0.3880 <x>=0.3839 8.528 <x>=8.556


qin(z=0)

≤ 3A/4)

T8
xσ=3.0E-03 xσ=4.6E-02
T8

T8
85.21 <x>=85.58 0.3826 8.622
xσ=2.7E-01
85.77
0.5955 23.78 xm=23.78
xm=0.5950

z=7A/8
z=7A/8

0.5950 <x>=0.5952 23.94 <x>=23.83

median: xm
median: xm

median: xm
xσ=2.6E-04 xσ=8.5E-02
0.5950 23.75
11.97
qin(z=A)

xm=11.97
0.6608 xm=0.6595 40.50 xm=40.50
12.45 <x>=12.13

z=A
z=A

xσ=2.2E-01 0.6550 <x>=0.6584 40.99 <x>=40.62


11.97
xσ=2.5E-03 xσ=2.6E-01
0.6595 40.39
(Tab. 2) and statistical quantities: average, standard deviation and median values.
Figure 6: Histograms of results obtained for LB (Tab. 1). Data provided by the different teams

10

20
0.2

0.4

0.6

0.8
0.02

0.04

0.06

0.08

0.12

0
0.1

1
0
7.265
xm=7.267

z=3A/8
0.01707 7.269

z=3A/8
xm=0.01706

z=3A/8
0.01712 <x>=7.268
<x>=0.01707 7.266
0.01705 xσ=2.6E-03
xσ=3.0E-05 7.271
0.01704

(a) f
6.217
0.03406 xm=6.219

z=A/2
xm=0.03404 6.220

z=A/2
z=A/2
<x>=6.219

Nu1 (z),

T7
T2
0.03405

T7
T2
T7
T2
<x>=0.03404 6.218
0.03403 xσ=1.7E-03
xσ=1.6E-05 6.221
0.03402
0.1540 5.674
0.05108 xm=5.674

(c) dw (z)

z=5A/8
xm=0.1534

(e) θb (z)

z=5A/8
xm=0.05104 5.674

z=5A/8
0.05084 0.1520 <x>=5.675
<x>=0.05100 <x>=0.1532 5.675
0.1530

Nu1 (z
0.05105 xσ=8.0E-04 xσ=1.4E-03
xσ=9.3E-05 0.1539 5.677
0.05102

T8
T3
T8
T3
T8
T3
0.4138 5.648
0.06790 xm=0.4139 xm=5.647

z=3A/4
xm=0.06788

z=3A/4
5.646
z=3A/4
0.06790 0.4140 <x>=5.659
<x>=0.06788 <x>=0.4139 5.645

≤ 3A/4)
0.06787 0.4140 xσ=2.1E-02
xσ=2.1E-05 xσ=6.6E-05

median: xm
0.4139 5.696

median: xm
0.06785

median: xm
0.06801 0.4864 9.837
xm=0.06801 xm=0.4862 xm=9.835

z=7A/8
z=7A/8
z=7A/8

0.06802 0.4860 9.835


<x>=0.06801 <x>=0.4862 <x>=9.835
0.06801 0.4860 9.836
xσ=8.7E-06 xσ=1.8E-04 xσ=8.7E-04
0.06800 0.4863 9.834

0.06792 0.5231 11.87


xm=0.06794 xm=0.5230 xm=11.87
0.06800 0.5220 11.87

z=A
z=A
z=A

<x>=0.06795 <x>=0.5244 <x>=11.87


0.06794 0.5230 11.87
xσ=3.1E-05 xσ=3.0E-03 xσ=3.7E-03
0.06794 0.5294 11.86
32

20

40

60

80
100

120

0.2

0.4

0.6

0.8

0
20

40

60

80
0

1
(f) Nu1 , Nu2 , qin (z = 0) and qin (z = A)

8.293
6.909 xm=8.296

z=3A/8
8.304

z=3A/8
xm=6.913 <x>=8.297
6.912 8.293
Nu1


<x>=6.915 xσ=4.6E-03
6.914 8.300

(b) f
xσ=6.5E-03
6.926 7.887
xm=7.889

z=A/2
7.900
z=A/2
<x>=7.891

T7
T2
Nu2 (z),
T7
T2

T7
T2
7.887
xσ=5.3E-03
8.574 7.892
xm=8.580
8.583 7.990
Nu2


<x>=8.583 0.2293 xm=7.991


(d) dψ (z)

z=5A/8
xm=0.2260
z=5A/8

8.577 0.2280 8.034


xσ=9.0E-03 <x>=0.2261 <x>=8.001
0.2230 7.989

Nu2 (z
8.597 xσ=2.7E-03 xσ=1.9E-02
0.2239 7.993

T8
T3
T8
T3

T8
T3
0.5960 9.166
xm=0.5953 xm=9.160

z=3A/4
73.50
z=3A/4

0.5950 9.154
xm=73.47 <x>=0.5941 <x>=9.189
qin(z=0)

73.46 9.150

≤ 3A/4)
0.5900 xσ=5.6E-02
<x>=73.47 xσ=2.4E-03

median: xm
0.5956 9.285
median: xm

median: xm
73.43
xσ=2.3E-02
73.47 0.6922 29.71
xm=0.6921 xm=29.71

z=7A/8
z=7A/8

0.6940 29.71
<x>=0.6913 <x>=29.71
0.6870 29.71
xσ=2.6E-03 xσ=1.3E-02
18.59 0.6920 29.68
xm=18.59
qin(z=A)

18.60 0.7208 61.33


<x>=18.60 xm=0.7204 xm=61.34
0.7200 61.35

z=A
z=A

18.61 <x>=0.7195 <x>=61.31


xσ=1.0E-02 0.7160 61.36
18.58 xσ=2.1E-03 xσ=7.2E-02
0.7212 61.18
w(x,A)=400 w(x,A) w(x,A)=400 w(x,A)

w(x,A)=300 w(x,A)=300

w(x,A)=200 w(x,A)=200
w(x,3A/4) w(x,3A/4)
w(x,A)=100 w(x,3A/4)=300 w(x,A)=100 w(x,3A/4)=300

w(x,A)= 0 w(x,3A/4)=200 w(x,A)= 0 w(x,3A/4)=200


w(x,A)= -50 w(x,A)= -50
w(x,3A/4)=100 w(x,3A/4)=100
w(x,A/2) w(x,A/2)
w(x,A/2)=200 w(x,3A/4)= 0 w(x,A/2)=200 w(x,3A/4)= 0

w(x,A/2)=100 w(x,A/2)=100

w(x,A/2)= 0 w(x,A/4)=200 w(x,A/2)= 0 w(x,A/4)=200


w(x,A/4) w(x,A/4)
w(x,A/4)=100 w(x,A/4)=100

w(x,0)=200 w(x,A/4)=0 w(x,0)=200 w(x,A/4)=0


w(x,0)
w(x,0)
w(x,0)=100 w(x,0)=100

w(x,0)= 0 w(x,0)= 0
0 0.5 1 0 0.5 1

(a) LB (b) GB
Figure 7: Vertical component of the velocity at different horizontal sections of the channel for
(a) LB and (b) GB boundary conditions (Tab. 1). The vertical thick line located on the left wall
emphasizes the heated section. Data provided by team T2.

A A

7A/8 LB 7A/8 GB

3A/4 GB 3A/4 GB-0

5A/8 5A/8

A/2 GB-0 A/2 LB-0


z
z

3A/8 3A/8 LB

A/4 LB-0 A/4

A/8 A/8

0 0
-6000 -5000 -4000 -3000 -2000 -1000 0 -40 -20 0 20 40 60 80 100 120 140
p(0.5,z) w(0.5,z)

(a) Pressure (b) Vertical component of the velocity


Figure 8: (a) Pressure and (b) vertical velocity profiles as a function of z at mid-width of the
channel for boundary conditions LB, GB, LB-0 and GB-0 (Tab. 1). Data provided by team T8.

33
graphics are the reference solutions which are set to the median values.
standard deviation (xσ /hxi) reported in parentheses and expressed in permillage. Values below the
for LB-0, LB, GB-0 and GB (Tab. 1). Boxes are topped by the average values and the relative
Figure 9: Histograms and error bars stand for average values (hxi) and standard deviations (xσ ),
Benchmark
values

10

15

20
Benchmark Benchmark

0.02

0.04

0.06

0.08

0.25

0.75

5
(xm)

0.1

0.5
values values

1
(xm) (xm)
7.198 7.196 (± 0.45o⁄oo)
0.02083 (± 9.1ooo⁄ )

z=3A/8
0.02094 7.267 7.268 (± 0.36o⁄oo)

z=3A/8

z=3A/8
0.01706 0.01707 (± 1.8o⁄oo) 7.286 ⁄ )
7.291 (± 3.3ooo
0.01614 0.01602 (± 17.4o⁄oo) 7.325 ⁄ )
7.327 (± 0.46 ooo
0.01463 ⁄ )
0.01464 (± 2.8 ooo

(a) f

LB−0
⁄ )
6.169 (± 0.38ooo

LB−0

LB−0
6.170
0.04182 ⁄ )
0.04157 (± 9.1ooo
6.219 (± 0.28o⁄oo)

z=A/2
6.219
0.03404 (± 0.48o⁄oo)

z=A/2

z=A/2

Nu1 (z),
0.03404
6.238 6.242 (± 3.0o⁄oo)
0.03226 0.03205 (± 17o⁄oo)
6.265 6.266 (± 0.39 ooo⁄ )
0.02919 ⁄ )
0.02922 (± 2.8 ooo
5.627 5.627 (± 0.42ooo⁄ )
0.06232 (± 9.1o⁄oo) 0.4088 (± 19o⁄oo)

LB
0.06270 0.4141

(c) dw (z)

z=5A/8
LB

LB
5.675 (± 0.25o⁄oo)

(e) θb (z)
5.674

z=5A/8

z=5A/8
0.05104 0.05100 (± 1.8o⁄oo) 0.1534 0.1532 (± 5.3 ⁄oo) o
5.693 5.698 (± 3.3o

oo)
0.04784 (± 19o⁄oo) 0.0989 (± 529o⁄oo)

Nu1 (z
0.04831 0.0902
5.716 5.717 (± 0.30 o⁄oo)
0.04375 ⁄ )
0.04381 (± 2.8 ooo

GB−0
GB−0

GB−0
5.602 5.616 (± 3.9o⁄oo)
0.08338 0.08287 (± 9.1o⁄oo) 0.5157 0.5132 (± 7.2o⁄oo)

z=3A/4
5.647 5.659 (± 3.8o⁄oo)

z=3A/4

z=3A/4
0.06788 0.06788 (± 0.31o⁄oo) 0.4139 0.4139 (± 0.16o⁄oo)
5.665 ⁄ )
5.638 (± 14ooo

≤ 3A/4)
0.06431 0.06395 (± 18o⁄oo) 0.3713 0.3683 (± 58o⁄oo)
⁄ ) ⁄ ) 5.694 ⁄ )
5.708 (± 4.2 ooo
0.05820 0.05828 (± 2.8 ooo 0.2591 0.2620 (± 16 ooo

0.08303 (± 9.0o⁄oo) ⁄ ) 9.692 9.697 (± 0.76o⁄oo)

GB
0.08356 0.5424 0.5409 (± 3.8ooo

GB

GB

z=7A/8
9.835 (± 0.088o⁄oo)
z=7A/8

z=7A/8
0.06801 0.06801 (± 0.13o⁄oo) 0.4862 0.4862 (± 0.37o⁄oo) 9.835
0.06439 0.06390 (± 17o⁄oo) 0.4640 0.4618 (± 20o⁄oo) 9.886 ⁄ )
9.912 (± 5.4ooo
0.05828 ⁄ )
0.05838 (± 2.9 ooo 0.4106 ⁄ )
0.4117 (± 4.0 ooo 9.965 ⁄ )
9.965 (± 0.22 ooo

0.08348 0.08298 (± 9.0o⁄oo) 0.5530 0.5532 (± 2.7o⁄oo) 11.68 11.69 (± 0.97o⁄oo)


0.06794 0.06795 (± 0.45o⁄oo) 0.5230 0.5244 (± 5.6o⁄oo) 11.87 11.87 (± 0.31o⁄oo)

z=A
z=A

z=A
0.06437 0.06382 (± 17o⁄oo) 0.5018 0.5007 (± 22o⁄oo) 11.95 ⁄ )
12.04 (± 11ooo
0.05823 ⁄ )
0.05835 (± 2.9 ooo 0.4758 ⁄ )
0.4765 (± 2.3 ooo 12.05 ⁄ )
12.05 (± 0.35 ooo
34

Benchmark Benchmark

100

100

200

300

400

500
Benchmark

0.25

0.75
values values
20

40

60

80

0.5
values
0

0
(xm) (xm)

1
(xm)
(f) Nu1 , Nu2 , qin (z = 0) and qin (z = A)

6.840 6.838 (± 1.0o⁄oo) 8.469 ⁄ )


8.464 (± 1.3ooo

z=3A/8
8.296 8.297 (± 0.56o⁄oo)

z=3A/8
6.913 6.915 (± 0.94 ⁄oo) o
8.257 (± 3.1 ⁄oo)
o
Nu1


8.255
6.947 6.955 (± 3.6o⁄oo) 8.207 8.207 (± 0.32o⁄oo)

(b) f
LB−0

LB−0
6.974 (± 1.1 o⁄oo)

LB−0
6.969 8.311 8.298 (± 2.8o⁄oo)
7.891 (± 0.67o⁄oo)

z=A/2
7.889

z=A/2

Nu2 (z),
7.807 7.805 (± 5.1o⁄oo)
9.005 8.998 (± 3.2o⁄oo)
7.671 7.671 (± 0.36o⁄oo)
8.580 8.583 (± 1.0o⁄oo)
8.666 (± 4.4o⁄oo)
Nu2


0.5835 (± 20o⁄oo) 8.690


LB

LB
0.5912
(d) dψ (z)

LB
8.498 (± 5.1o⁄oo)

z=5A/8
8.484 8.001 (± 2.4o⁄oo)
z=5A/8

0.2260 0.2261 (± 12o⁄oo) 7.991


8.363 (± 1.1 o⁄oo) 0.1432 (± 482o⁄oo) 7.853 7.849 (± 7.7o⁄oo)

Nu2 (z
8.362 0.1334
7.626 7.628 (± 0.66ooo⁄ )
GB−0

GB−0
GB−0
0.7257 0.7211 (± 9.0o⁄oo) 10.52 10.51 (± 13o⁄oo)
59.82 60.17 (± 8.6o⁄oo)

z=3A/4
9.189 (± 6.1o⁄oo)
z=3A/4

0.5953 0.5941 (± 4.1o⁄oo) 9.160


qin(z=0)

73.47 73.47 (± 0.32 ⁄oo)


o
⁄ ) 8.877 (± 17o⁄oo)

≤ 3A/4)
0.5392 0.5369 (± 58ooo 8.917
⁄ )
0.3839 (± 7.7 ooo 8.528 ⁄ )
8.556 (± 5.4ooo
77.63 78.17 (± 17o⁄oo) 0.3826

85.58 (± 3.1 o⁄oo) 0.7607 (± 7.1o⁄oo) 50.95 49.82 (± 33o⁄oo)


GB

GB
85.77 0.7644

GB

z=7A/8
29.71 (± 0.43o⁄oo)
z=7A/8

0.6921 0.6913 (± 3.8o⁄oo) 29.71


0.6652 0.6618 (± 18o⁄oo) 27.22 27.06 (± 27o⁄oo)
28.87 28.65 (± 11o⁄oo) 0.5950 ⁄ )
0.5952 (± 0.44 ooo 23.78 ⁄ )
23.83 (± 3.5ooo
qin(z=A)

18.59 18.60 (± 0.56o⁄oo) 0.7775 0.7746 (± 6.1o⁄oo) 471.0 409.5 (± 219o⁄oo)


17.11 (± 52o⁄oo) 0.7204 0.7195 (± 2.9o⁄oo) 61.34 61.31 (± 1.2o⁄oo)

z=A
z=A

17.52
0.7049 0.7022 (± 11o⁄oo) 51.76 52.14 (± 70o⁄oo)
11.97 12.13 (± 19 o⁄oo) ⁄ ) ⁄ )
0.6595 0.6584 (± 3.8 ooo 40.50 40.62 (± 6.5ooo

You might also like