Vous êtes sur la page 1sur 8

Science Academy Transactions on Renewable Energy Systems Engineering and Technology (SATRESET)

Vol. 1, No. 4, December 2011, ISSN: 2046-6404


Science Academy Publisher, United Kingdom
www.sciacademypublisher.com


116
Numerical investigation of the Rutland 913 wind turbine airfoils
effect on the aerodynamic structure flow
Zied DRISS*, Walid TRIKI and Mohamed Salah ABID
Laboratory of Electro-Mechanic Systems (LASEM), National School of Engineers of Sfax (ENIS), University of Sfax (US), BP. 1173, Road
Soukra, km 3.5, Sfax, TUNISIA
*Email: zied.driss@enis.rnu.tn (Corresponding author)
Abstract In this paper, we are interested in studying the effect of the Rutland 913 wind turbine airfoils. The numerical results
obtained in the cases of the airfoils type SD2030 and BM4640 are particularly predicted and analyzed. The results, from
application of the computational fluid dynamics (CFD) code "Fluent", are presented in the transversal and longitudinal planes
of the considered control volume. The Navier-Stokes equations are solved by a finite volume discretization method. The
turbulence model used is the RNG k-. The objective is to study the effect of the airfoil type on the aerodynamic structure flow
around the horizontal-axis wind turbine.
Keywords Wind turbine, Rutland 913, Airfoil, Aerodynamic, Turbulent, CFD.


1. Introduction
Aerodynamic and structural optimization of wind turbine
airfoils has become a subject of considerable interest. It
involves the determination of the geometry of an
aerodynamic configuration, which satisfies certain objectives
[1-6]. To do this, a deep research has resulted in the
development of a series of Computational Fluid Dynamics
(CFD) tools based on the solution of the Reynolds averaged
Navier-Stokes (RANS) equations. This research has mostly
been related to the aerodynamics of wind turbines, notably
the development of basic solution algorithms, numerical
schemes for solution of the flow equations, grid generation
techniques and the modelling of boundary layer turbulence
[7-8]. These elements together form the basis of all CFD
codes, of which some already exist as standard commercial
software. Already, some of experiences have been exploited
to validate the CFD codes [9-12]. Then, we can consider that
CFD has matured to become a tool for predicting and
understanding the flow physics of modern wind turbine
rotors. Numerical calculations using commercial CFD codes
were made by several authors. [13-15].
In this paper, a numerical investigation within the CFD code
"Fluent" was carried to study the effect of the horizontal-axis
wind turbine airfoils.

2. Geometrical configuration
The geometrical configuration consists on the Rutland
913 wind turbine (Figure 1). This turbine is a six-blade
orizontal-axis wind turbine. It can be equipped by a scoop to
increase the output power of the wind turbine. Also, energy
capture can be improved at lower wind speeds [16-18].
Nevertheless, in this paper we are interested of the two types
of airfoils called SD2030 and BM4640 (Figure 2).


Figure 1. Rutland 913 wind turbine.


(a) SD2030 airfoil type (b) BM4640 airfoil type
Figure 2. Airfoil type.

3. Numerical model
The purpose of this section is to give an overview of
issues related to the turbulence models provided in
"FLUENT". The Reynolds-averaged Navier-Stokes (RANS)
equations govern the transport of the averaged flow
quantities, with the whole range of the scales of turbulence
being modeled. The RANS-based modeling approach greatly
reduces the required computational effort and resources and
is widely adopted for practical engineering applications. An
entire hierarchy of closure models is available in "FLUENT"
including Spalart-Allmaras, k- and its variants, k- and its
variants and the RSM. In our simulation, we are interested to
the RNG k- model [19-20].
Z. Driss et al. / SATRESET, Vol. 1, No. 4, pp. 116-123, December 2011

117
3.1. Reynolds averaged Navier-Stokes (RANS) equations
In Reynolds averaging, the solution variables in the
instantaneous Navier-Stokes equations are decomposed into
the mean (ensemble-averaged or time-averaged) and
fluctuating components. For the velocity components (i=1, 2,
3):
'
i i i
u u u = (1)

where
i
u and
'
i
u are the mean and fluctuating velocity
components (m.s
-1
).
Likewise, for pressure and other scalar quantities denoted by
i
:
'
i i i
= (2)

Substituting expressions of this form for the flow variables
into the instantaneous continuity and momentum equations
and taking a time (or ensemble) average (and dropping the
overbar on the mean velocity u ) yields the ensemble-
averaged momentum equations, they can be written as
follows:
( )
i
i

u 0
t x

=

(3)
( ) ( )
i i j
j i
p
u u u
t x x

=


( )
j ' ' i l
ij i j
j j i l j
u
u u 2
u u
x x x 3 x x
l 1

l


l ( )
l
(4)

These equations are called Reynolds-averaged Navier-Stokes
(RANS) equations. They have the same general form as the
instantaneous Navier-Stokes equations, with the velocities
and other solution variables now representing ensemble-
averaged values. Additional terms appear that represent the
effects of turbulence. These Reynolds stresses
' '
i j
u u must
be modeled in order to close equation (4).
3.2. Boussinesq hypothesis
The Reynolds-averaged approach to turbulence modeling
requires that the Reynolds stresses in equation (4) be
appropriately modeled [21]. A common method employs the
Boussinesq hypothesis to relate the Reynolds stresses to the
mean velocity gradients:
j ' ' i i
i j t t ij
j i i
u
u u 2
u u k
x x 3 x
1
1

=


( )
( )
(5)

The Boussinesq hypothesis is used in the Spalart-Allmaras
model, the k- models and the k- models. The advantage of
this approach is the relatively flow computational cost
associated with the computation of the turbulent viscosity
t
.
In the case of the Spalart-Allmaras model, only one
additional transport equation, representing turbulent
viscosity, is solved. In the case of the k- and k- models,
two additional transport equations are solved, and
t
is
computed as a function of k and . In this case, k is the
turbulence kinetic energy, is the turbulence dissipation rate
and is the specific dissipation rate.
3.3. RNG k- model
The RNG k- model was derived using a rigorous
statistical technique, called renormalization group theory [19-
20]. It is similar in form to the standard k- model, but
includes the following refinements:
- The RNG model has an additional term in its equation that
significantly improves the accuracy for rapidly strained
flows.
- The effect of swirl on turbulence is included in the RNG
model, enhancing accuracy for swirling flows.
- The RNG theory provides an analytical formula for
turbulent Prandtl numbers, while the standard k- model uses
user-specified, constant values.
- While the standard k- model is a high-Reynolds-number
model, the RNG theory provides an analytically-derived
differential formula for effective viscosity that accounts for
low-Reynolds-number effects. Effective use of this feature
does, however, depend on an appropriate treatment of the
near-wall region.
These features make the RNG k- model more accurate and
reliable for a wider class of flows than the standard k-
model. The RNG-based k- turbulence model is derived from
the instantaneous Navier-Stokes equations, using a
mathematical technique called "renormalization group"
(RNG) methods. The analytical derivation results in a model
with constants different from those in the standard k- model
and additional terms and functions in the transport equations
for k and . A more comprehensive description of RNG
theory and its application to turbulence can be found [22].

The RNG k- model has a similar form to the standard k-
model:
( ) ( )
i k eff
i j j
k
k k u
t x x x
1


=



( )

k b M k
G G Y S (6)
( ) ( )
i eff
i j j

u
t x x x
1


=



( )

( )
2
1 k 3 b 2

C G C G C R S
k k
(7)

The scale elimination procedure in RNG theory results in a
differential equation for turbulent viscosity:
2
3

k
d 1.72 d
1 C
1


=



( )
(8)
Where:
eff

= (9)

C 100
(10)

This equation is integrated to obtain an accurate description
of how the effective turbulent transport varies with the
effective Reynolds number (or eddy scale), allowing the
model to better handle low-Reynolds-number and near-wall
flows. In the high-Reynolds-number limit, equation (8) gives
2
t
k
C

= (11)
Z. Driss et al. / SATRESET, Vol. 1, No. 4, pp. 116-123, December 2011

118
with

C =0.0845, derived using RNG theory. It is interesting


to note that this value of

C is very close to the empirically-


determined value of 0.09 used in the standard k- model. In
"FLUENT", the effective viscosity is computed using the
high-Reynolds-number form in equation (11). However, there
is an option available that allows using the differential
relation given in equation (8) when its essential to include
low-Reynolds-number effects.

Turbulence, in general, is affected by rotation or swirl in the
mean flow. The RNG model in "FLUENT" provides an
option to account the effects of swirl or rotation by modifying
the turbulent viscosity appropriately. The modification takes
the following functional form:
t t0 s
k
f ,,

1

=


( )
(12)
where
t0
is the value of turbulent viscosity calculated
without the swirl modification using either equation (8) or
equation (11). is a characteristic swirl number evaluated
within "FLUENT".
s
is a swirl constant that assumes
different values depending on whether the flow is swirl-
dominated or only mildly swirling. This swirl modification
always takes effect for axisymmetric, swirling flows and
three-dimensional flows when the RNG model is selected.
For mildly swirling flows,
s
is set to 0.05 and cannot be
modified. However, for strongly swirling flows a higher
value of
s
can be used.

The inverse effective Prandtl numbers,
k
and

, are
computed using the following formula derived analytically
by the RNG theory:
0.6321 0.3679
mol
0 0 eff
1.3929 2.3929
1.3929 2.3929

=

(13)
where:
0
=1.0 .
In the high-Reynolds-number limit (
mol
eff

):
k
1.393 = (14)

The main difference between the RNG and standard k-
models lies in the additional term in the equation given by:
3

2
0
3

C 1

R
k 1
1



( )
=

(15)
Where:
S k

= (16)

The effects of this term in the RNG equation can be seen
more clearly by rearranging equation (7). Using equation
(15), the third and fourth terms on the right-hand side of
equation (7) can be merged, and the resulting equation can
be rewritten as:
( ) ( )
i eff
i j j

u
t x x x
1


=



( )

( )
2
1 k 3 b 2

C G C G C
k k

(17)
Where
2
C

is given by:
3

0
2 2
3

C 1

C C
1


( )
=

(18)

The RNG model is more responsive to the effects of rapid
strain and streamline curvature than the standard k- model,
which explains the superior performance of the RNG model
for certain classes of flows. The model constants C
1
and
2
C
in equation (7) have values derived analytically by the RNG
theory (Table 1).

Table 1. Constants of the RNG k- model
C
1
C
2
C


0

1.42 1.68 0.0854 4.38 0.012

4. Numerical results
Figure 3 shows the control volume considered to study
the flow around the Rutland 913 wind turbine equipped by a
scoop. The results are presented using a cylindrical
coordinate system (r, , z) with the origin in the center of the
wind turbine. The orientation of the z-axis coincides with
symmetrical axis of the scoop and the control volume
considered. Particularly, we are interested to the velocity
vectors, the velocity magnitude, the static pressure and the
dynamic pressure. These results are presented in different
longitudinal and transversal planes. The first longitudinal
plane is defined by =0 and it is localized between two
blades. The second longitudinal plane is defined by =30
and it passes through the first blade. The third transversal
plane is defined by z=0 and it intercepts all the wind turbine
blades.

Figure 3. Control volume considered.
4.1. Velocity vectors
Figures 4, 5 and 6 compare the velocity vectors colored
by velocity magnitude for the SD2030 and BM4640 airfoil
type. These results are presented in the longitudinal and
transversal planes defined respectively by =0, =30 and
z=0. According to these results, it is clear that the flow is
uniform and presents a horizontal direction in the inlet of the
control volume. While approaching to the wind turbine
equipped by the scoop, the velocity field is affected by the
geometric configuration. In fact, a flow deceleration appears
Z. Driss et al. / SATRESET, Vol. 1, No. 4, pp. 116-123, December 2011

119
on the entry, the exit and around the external envelope of the
scoop. However, inside the scoop the flow takes back his
movement in the wind turbine upstream. At the end of the
turbine wind blades, an acceleration of the flow is appeared.
This fact is also observed far from this domain and
particularly it is generated from the scoop collector.
Otherwise, its noted that the airfoil type have an effect on
the velocity vectors distribution. In fact, with the SD2030
airfoil, two developed recirculation zones appear around the
scoop and in the wind turbine downstream. Within a BM4640
airfoil, the same fact is observed. However, its noted that the
airfoil type has a direct effect on the recirculation zones
shape less developed with the BM4640 airfoil.


(a) SD2030 airfoil type

(b) BM4640 airfoil type
Figure 4. Velocity vectors colored by velocity magnitude in
the longitudinal plane defined by =0.


(a) SD2030 airfoil type

(b) BM4640 airfoil type
Figure 5. Velocity vectors colored by velocity magnitude in
the longitudinal plane defined by =30.

(a) SD2030 airfoil type

(b) BM4640 airfoil type
Figure 6. Velocity vectors colored by velocity magnitude in
the transversal plane defined by z=0.
4.2. Velocity magnitude
Figures 7, 8 and 9 compare the distribution of velocity
magnitude for the SD2030 and BM4640 airfoil type. These
results are presented in the longitudinal and transversal
planes defined respectively by =0, =30 and z=0.
According to these results, it is clear that the velocity
magnitude is uniform in the inlet of the control volume.
While approaching to the wind turbine equipped by the
scoop, the velocity magnitude value is affected by the
geometric configuration. In fact, the velocity magnitude
values decrease on the entry, the exit and around the external
envelope of the scoop. However, inside the scoop and
particularly in the downstream of the turbine blade, the
velocity magnitude values increase. Two wakes
characteristics of the highly velocity magnitude are appeared
in this area. Outside the scoop, two extended wakes are
created at the crossing of the scoop collector. On the meeting
of the scoop and the turbine blade, the velocity is null. The
wake characteristic of the weaker velocity magnitude appear
in the scoop excite. While comparing these results relative to
the SD2030 and BM4640 airfoil type, we note that the
maximal value of the velocity magnitude is gotten with the
SD2030 airfoil type.


(a) SD2030 airfoil type
Z. Driss et al. / SATRESET, Vol. 1, No. 4, pp. 116-123, December 2011

120

(b) BM4640 airfoil type
Figure 7. Distribution of velocity magnitude in the
longitudinal plane defined by =0.


(a) SD2030 airfoil type

(b) BM4640 airfoil type
Figure 8. Distribution of velocity magnitude in the
longitudinal plane defined by =30.


(a) SD2030 airfoil type

(b) BM4640 airfoil type
Figure 9. Distribution of velocity magnitude in the
transversal plane defined by z=0.
4.3. Static pressure
Figures 10, 11 and 12 compare the distribution of the
static pressure for the SD2030 and BM4640 airfoil type.
These results are presented in the longitudinal and transversal
planes defined respectively by =0, =30 and z=0.
According to these results, it is clear that the static pressure
presents a compression zone on the upstream of the scoop
and the wind turbine. The maximal values are obtained in the
scoop inside the collector neighborhood and just at the level
of the wind turbine. However, the depression zones are
localized around the external envelope of the scoop and in the
wind turbine downstream inside the scoop.


(a) SD2030 airfoil type

(b) BM4640 airfoil type
Figure 10. Distribution of static pressure in the longitudinal
plane defined by =0.


(a) SD2030 airfoil type

(b) BM4640 airfoil type
Figure 11. Distribution of static pressure in the longitudinal
plane defined by =30.

Z. Driss et al. / SATRESET, Vol. 1, No. 4, pp. 116-123, December 2011

121
While comparing these results relative to the SD2030 and
BM4640 airfoil type, we note that the compression and
depression zones shape are widely assigned by the airfoil
type. The weakest value of the static pressure is obtained
within a SD2030 airfoil type.


(a) SD2030 airfoil type

(b) BM4640 airfoil type
Figure 12. Distribution of static pressure in the transversal
plane defined by z=0.
4.4. Dynamic pressure
Figures 13, 14 and 15 compare the distribution of the
dynamic pressure for the SD2030 and BM4640 airfoil type.
These results are presented in the longitudinal and transversal
planes defined respectively by =0, =30 and z=0.
According to these results, it is clear that the wakes
characteristics of the weakest values of the dynamic pressure
are created on the entry, the exit and around the external
envelope of the scoop. The first weak takes root in the scoop
collector by merging his surface from the two sides. In the
scoop inside, the dynamic pressure increase at the time of the
air flow advancement. While approaching of the wind
turbine, the dynamic pressure decreases again. The same fact
is observed in the wind turbine downstream where the wake
zone is developed more and more. In the scoop outside, this
wake crosses the one coming from the scoop collector.
Otherwise, the wakes characteristics of the strong values of
the dynamic pressure are created in the scoop inside and
more precisely in the downstream blades. In the scoop
outside, others wakes are developed and generated from the
scoop collector. While comparing these results relative to the
SD2030 and BM4640 airfoil type, we note that the maximal
value of the dynamic pressure is gotten within an SD2030
airfoil type.


(a) SD2030 airfoil type

(b) BM4640 airfoil type
Figure 13. Distribution of dynamic pressure in the
longitudinal plane defined by =0.


(a) SD2030 airfoil type

(b) BM4640 airfoil type
Figure 14. Distribution of dynamic pressure in the
longitudinal plane defined by =30.


(a) SD2030 airfoil type

Z. Driss et al. / SATRESET, Vol. 1, No. 4, pp. 116-123, December 2011

122

(b) BM4640 airfoil type
Figure 15. Distribution of dynamic pressure in the transversal
plane defined by z=0

5. Conclusion
In this paper, the aerodynamic characteristics of a Rutland
913 wind turbine have been studied. Particularly, we have
compared the numerical results obtained from the application
of the commercial CFD code "Fluent" in the cases of the
airfoils type BM4640 and SD2030. From the numerical
results, it has been noted that the airfoil type has a direct
effect on the recirculation zones shape which are less
developed with the BM4640 airfoil. The compression and
depression zones shape are widely assigned by the airfoil
type. The weakest value of the static pressure is obtained
with the SD2030 airfoil.
In the future, we propose to develop an experimental
investigation within a particle velocimetry laser (PIV)
system.

Nomenclature
2
C

model constant
C
1
constant of the turbulence model

C
2
constant of the turbulence model
C
3
constant of the turbulence model

C

constant of the turbulence model

C constant of the turbulence model


G
b
generation of turbulence kinetic energy, kg.m
-1
.s
-3

G
k
generation of turbulence kinetic energy, kg.m
-1
.s
-3

k turbulence kinetic energy, m
2
.s
-2

p static pressure, Pa
Re Reynolds number
R

source term of the dissipation rate of the turbulent
kinetic energy, kg.m
-1
.s
-3

S scalar measure of the deformation tensor
S
k
source term of the turbulent kinetic energy, kg.m
-1
.s
-3

S
ij
mean strain rate, s
-1

S

source term of the dissipation rate of the turbulent
kinetic energy, kg.m
-1
.s
-3

t time, s
u
i
velocity components, m.s
-1

i
u mean velocity components, m.s
-1

'
i
u fluctuating velocity components, m.s
-1

x cartesian coordinate, m
y cartesian coordinate, m
Y
M
contribution of the fluctuating dilatation in compressible
turbulence to the overall dissipation rate, kg.m
-1
.s
-3


Greek symbols
angular coordinate, rad

s
swirl constant

k
inverse effective Prandtl number for k


inverse effective Prandtl number for
turbulence dissipation rate, m
2
.s
-3

dynamic viscosity, Pa.s

t
turbulent viscosity, Pa.s

eff
effective viscosity, Pa.s

mol
molecular viscosity, Pa.s
strain rate
molecular kinematic viscosity, m
2
.s
-1

specific dissipation rate, m
2
.s
-3

density, kg.m
-3

i
scalar quantities
i
mean scalar quantities
'
i
fluctuating scalar quantities
swirl number

References
[1] G. Bouzgarrou, Z. Driss, N. Bahloul, H. Kchaou, M.S. Abid, "Etude
de linfluence dun diffuseur sur la structure arodynamique dune
olienne axe horizontal type Rutland 913", 14
me
Journes
Internationales de Thermique (JITH2009), Djerba, Tunisie, pp. 1-6,
2009.
[2] J.C. Dai, Y.P. Hu, D.S. Liu, X. Long, "Aerodynamic loads calculation
and analysis for large scale wind turbine based on combining BEM
modified theory with dynamic stall model", Renewable Energy, 36,
1095-1104, 2011.
[3] R. Lanzafame, M. Messina, "Horizontal axis wind turbine working at
maximum power coefficient continuously", Renewable Energy, 35, pp.
301-306, 2010.
[4] K. Pope, I. Dincer, G.F. Naterer, "Energy and exergy efficiency
comparison of horizontal and vertical axis wind turbines", Renewable
Energy, 35, pp. 2102-2113, 2010.
[5] T. Sharpe, G. Proven, "Crossflex: Concept and early development of a
true building integrated wind turbine", Energy and Buildings, 42, pp.
2365-2375, 2010.
[6] C. Sicot, P. Devinant, S. Loyer, J. Hureau, "Rotational and turbulence
effects on a wind turbine blade: Investigation of the stall mechanisms",
Journal of Wind Engineering and Industrial Aerodynamics, 96, pp.
1320-1331, 2008.
[7] Z. Driss, S. Karray, H. Kchaou, M.S. Abid, "Computer simulations of
laminar flow generated by an anchor blade and a Maxblend impellers",
Science Academy Transactions on Renewable Energy Systems
Engineering and Technology, Vol. 1, N. 3, pp. 68-76, 2011.
[8] Z. Driss, G. Bouzgarrou, W. Chtourou, H. Kchaou, M.S. Abid,
"Computational studies of the pitched blade turbines design effect on
the stirred tank flow characteristics", European Journal of Mechanics
B/Fluids, 29, pp. 236-245, 2010.
[9] A. Damak, Z. Driss, H. Kchaou, M.S. Abid, "Conception et ralisation
dune soufflerie aspiration", 4
me
Congrs International Conception et
Modlisation des Systmes Mcaniques (CMSM11), Sousse, Tunisie,
pp. 1-7, 2011.
[10] K. Kishinami, H. Taniguchi, J. Suzuki, H. Ibano, T. Kazunou, M.
Turuhami, "Theoretical and experimental study on the aerodynamic
characteristics of a horizontal axis wind turbine". Energy, 30, pp.
2089-2100, 2005.
[11] Z. Driss, A. Damak, H. Kchaou, M. S. Abid, "Experimental
investigation on wind tunnel", Tunisian Japanese Symposium on
Science, Society and Technology (TJASSST'11), Hammamet, Tunisia,
pp. 1-4, 2011.
[12] A. Damak, Z. Driss, A. Kaffel, M.S. Abid, "Experimental study of
helical Savonius rotor", Tunisian Japanese Symposium on Science,
Society and Technology (TJASSST'11), Hammamet, Tunisia, pp. 1-4,
2011.
[13] Z. Driss, M. S. Abid, "Effect of the horizontal-axis wind turbine airfoil
on the aerodynamic structure flow", Tunisian Japanese Symposium on
Science, Society and Technology (TJASSST'11), Hammamet, Tunisia,
pp. 1-4, 2011.
[14] A. El Kasmi, C. Masson, "An extended k- model for turbulent flow
through horizontal-axis wind turbines". Journal of Wind Engineering
and Industrial Aerodynamics, 96, pp. 103-122, 2008.
[15] C.G. Gebhardt, S. Preidikman, J.C. Massa, "Numerical simulations of
the aerodynamic behavior of large horizontal-axis wind turbines".
International Journal of Hydrogen Energy, 35, pp. 6005-6011, 2010.
Z. Driss et al. / SATRESET, Vol. 1, No. 4, pp. 116-123, December 2011

123
[16] Z. Driss, S. Karray, H. Kchaou, M.S. Abid, "Caractrisation
arodynamique des oliennes axe horizontal quipes de rotors
sphrique et elliptique". Rcents Progrs en Gnie des Procds, 101,
pp. 1-6, 2011.
[17] F. Wang, L. Bai, J. Fletcher, J. Whiteford, D. Cullen, "The
methodology for aerodynamic study on a small domestic wind turbine
with scoop". Journal of Wind Engineering and Industrial
Aerodynamics, 96, pp. 1-24, 2008.
[18] F. Wang, L. Bai, J. Fletcher, J. Whiteford, D. Cullen, "Development of
small domestic wind turbine with scoop and prediction of its annual
power output". Renewable Energy, 33, pp. 1637-1651, 2008.
[19] Fluent Inc. "Users Guide", Centerra Resource Park, 2005.
[20] Z. Driss, M. Ammar, W. Chtourou, M.S. Abid, "CFD Modelling of
Stirred Tanks", Engineering Applications of Computational Fluid
Dynamics, Volume 1, Chapter 5, pp. 145-258, 2011.
[21] J.O. Hinze, "Turbulence", McGraw-Hill Publishing Co., New York,
1975.
[22] D. Choudhury, "Introduction to the Renormalization Group Method
and Turbulence Modeling". Fluent Inc., Technical Memorandum, TM-
107, 1993.

Vous aimerez peut-être aussi