Vous êtes sur la page 1sur 13

Numerical Simulation of Impinging Jets

Robert Wilke and Jörn Sesterhenn

Abstract This report concentrates on the investigation of heat transfer of a confined


round impinging jet. A direct numerical simulation was performed at a Reynolds
number of Re D 3;300 using a grid size of 512  512  512 points. It is shown that
the dissipative scales are well resolved. This enables the examination of the impact
of the jet’s turbulent flow field on the heat transfer of the impinged plate. In this
study the distribution of the local Nusselt number is presented and related to the
instantaneous flow field of the jet. First results of turbulent statistics are shown.

1 Introduction

Impinging jets provide an effective cooling method for various applications such as
the cooling of turbine blades of aircraft. An increase of efficiency not only reduces
the required cooling air mass flow and consequently the fuel consumption, but also
enables new combustion concepts with even higher cooling demands to be applied
in the future.
Heat transfer due to forced convection of a jet impinging on a flat plate has
been studied for decades. General information including schematic illustrations
of the flow fields as well as distributions of local Nusselt numbers for plenty
of different geometrical configurations and Reynolds numbers Re can be found
in several reviews, such as [7, 8, 12, 13] based on experimental and numerical
results. Since experiments cannot provide all quantities of the entire flow domain
spatially and temporally well resolved, the understanding of the turbulent flow
field requires simulations. Existing publications of numerical nature use either
turbulence modelling for the closure of the Reynolds-averaged Navier-Stokes
(RANS) equations, e.g. [15], or large eddy simulation (LES), e.g. [4]. Almost all
available direct numerical simulations (DNS) are either two-dimensional, e.g. [3],
or do not exhibit an appropriate spatial resolution in the three-dimensional case,

R. Wilke () • J. Sesterhenn


Fachgebiet Numerische Fluiddynamik, Technische Universität Berlin, Müller-Breslau-Str. 11,
10623 Berlin, Germany
e-mail: robert.wilke@tnt.tu-berlin.de; joern.sesterhenn@tu-berlin.de;
http://www.cfd.tu-berlin.de/

© Springer International Publishing Switzerland 2015 275


W.E. Nagel et al. (eds.), High Performance Computing in Science
and Engineering ’14, DOI 10.1007/978-3-319-10810-0_19
276 R. Wilke and J. Sesterhenn

e.g. [6]. Recent investigations come from Dairay et al [5]. He conducted a DNS
of a round impinging jet. Since those computations require tremendous amounts
of computing time, their presented results are preliminary and consider the Nusselt
number and the friction coefficient only.
This study deals with a direct numerical simulation of a turbulent round imping-
ing jet. According to Hrycak [7], for free jets as well as for impinging jets four
different states dependent on the Reynolds number exist. Those states characterise
the jet with regard to turbulence, ranging from dissipated laminar (Re < 300) to
fully turbulent (Re > 3;000). To ensure simulating a fully turbulent impinging jet, a
Reynolds number of Re D 3;300 is chosen.

2 Numerical Method

The governing Navier-Stokes equations are formulated in a characteristic pressure-


velocity-entropy-formulation, as described by Sesterhenn [11] and solved directly
numerically. This formulation has advantages in the fields of boundary conditions,
parallelization and space discretisation. No turbulence modelling is required since
the smallest scales of turbulent motion are resolved. The spatial discretisation uses
6th order compact central schemes of Lele [9] for the diffusive terms and compact
5th order upwind finite differences of Adams et al. [1] for the convective terms. To
advance in time a 4th order Runge-Kutta scheme is applied.
The present simulation is conducted on a numerical grid of size 512  512  512
points for the computational domain sized Lx  Ly  Lz D 12D  5D  12D,
where D is the inlet diameter, see Fig. 1. A confined impinging jet is characterised

Lx Lz
y
D

Ly

y
z
x

Fig. 1 Computational domain with instantaneous iso-surface of Q-criterion QD 12 P 2 C !ij !ij 
 @ui
 @ui @u   @ui @u 
Sij Sij ; P D  @xi
; !ij D 12 @x j
 @xji ; Sij D 12 @x j
C @xji at Q D 1;000 m2 s4
Numerical Simulation of Impinging Jets 277

Table 1 Physical parameters of the simulation. po ; p1 ; To ; TW ; Re; P r; ; R denote the total


pressure, ambient pressure, total temperature, wall temperature, Reynolds number, Prandtl number,
ratio of specific heats and the specific gas constant
po =p p1 To TW Re Pr  R
1:5 105 Pa 293:15 K 373:15 K 3;300 0:71 1:4 287 J/(kg K)

by the presence of two walls, the impinging plate and the orifice plate. The grid
is refined in those wall-adjacent regions in order to ascertain a maximum value in
time and space of the dimensionless wall distance y C of the closest grid point to the
wall smaller than one for both plates. For the x- and z-direction a slight symmetric
grid stretching is applied which refines the shear layer of the jet. The refinements
use hyperbolic tangent respectively hyperbolic sin functions and lead to minimal
and maximal spacings of 0:017D  x D z  0:039D and 0:0017D  y 
0:016D. The maximum change of the mesh spacing is 0:5 % respectively 1 %. For
the circumferential averaging an equidistant grid of size n  nr D 1;024  512 is
applied for each slice at constant height. Table 1 shows the physical parameters of
the simulation.
The computational domain is delimited by four non-reflecting boundary condi-
tions, one isothermal wall which is the impinging plate and one boundary consisting
of an isothermal wall and the inlet. The walls are fully acoustically reflective. The
location of the nozzle is defined using a hyperbolic tangent profile with a disturbed
thin laminar annular shear layer as described in [14].
A sponge region is applied for the outlet area r=D > 5, that smoothly forces the
values of pressure, velocity and entropy to reference values. This destroys vortices
before leaving the computational domain. The reference values at the outlet were
obtained by a preliminary large eddy simulation of a greater domain.

3 Results and Discussion

3.1 Kolmogorov Scales

The scales of turbulent motion span a huge range from the size of the domain to the
smallest energy dissipating ones. Since the turbulent kinetic energy is transferred
downwards to smaller and smaller scales, the smallest ones have to be resolved by
the numerical grid in order to obtain a reliable solution of the turbulent flow. They
are given with the kinematic viscosity  and dissipation rate  by

  14
3
l  (1)

278 R. Wilke and J. Sesterhenn

 1=3
Fig. 2 Grid spacing related to Kolmogorov length scale hx hy hz = l . The solid line represents
the value 1

and are valid for isotropic turbulence, that occurs at sufficiently high Reynolds
numbers. The Reynolds number of 3,300 can be considered as low or not sufficiently
high to obtain isotopic turbulence. Therefore the Kolmogorov microscales provide
a conservative clue. The ratio of the mesh width to the Kolmogorov length scale
 1=3
hx hy hz = l is shown in Fig. 2 and reaches a maximum value of 1.6 at the lower
wall. For supersonic turbulent boundary layers, Pirozzoli et al. [10] showed that
the typical size of small-scale eddies is about 5::6l . A strongly different behaviour
for the present boundary layer of high subsonic Mach number is not expected. The
 1=3
maximum ratio in the area of the free jet is hx hy hz = l D 1:3 and of the wall
 1=3
jet is hx hy hz = l D 1:5.

3.2 Mean and Instantaneous Flow Field

In addition to the criterion due to turbulent motion of the jet, the boundary layer
including the viscous sub-layer also has to be resolved appropriately in order to
achieve reliable results of the heat transfer at the impinging plate. The maximum
dimensionless wall distance
u y
yC D (2)


of the present simulation occurs at r=D D 0:46 and reaches a value of y C D 0:64.
The minimum number of points in the viscous sub-layer y C  5 is 7 for the entire
domain. Figure 3 shows the velocity- and temperature boundary layer profile for
different distances from the stagnation point. uC ; u and W are the dimensionless
radial velocity, the friction velocity and the wall shear stress:
Numerical Simulation of Impinging Jets 279

a b

Fig. 3 Mean boundary layer profiles for different radial positions. : r=D D 0:5, :
r=D D 1:0; : r=D D 1:5; : r=D D 2:0; : r=D D 3:0; : r=D D 4:5;
: uC D y C , uC Dln.y C /=0:41 C 5:1 respectively T C D P r y C . (a) Dimensionless
velocity. (b) Dimensionless temperature

r  
C u W @u
u D ; u D ; W D : (3)
u @y W

The dimensionless temperature is given by

TW  T qW
TC D ; T D ; (4)
T cp u

where T is the friction temperature and qW the wall heat flux.


The main flow characteristics can be compared with experimental results from
literature. Figure 4 shows the mean distributions of the radial velocity and tempera-
ture for different radial positions and Fig. 5 for different distances from the wall.
The maximal mean radial velocity increases from the stagnation point, reaches
its maximum value at r=D D 0:8 and then decreases. The maximal temperature
difference T  To increases monotonously with greater radii as a consequence of
turbulent mixing of the cold impinging jet and the hot environment. For the same
reason, the profile at r=D D 0:5 which is located in the shear layer of the jet is at
lower temperatures than the other ones.
The heat transfer at the impinging plate is strongly related to the vortical
structures of the turbulent flow field. In the shear layer of the jet (primary)
ring vortices develop and grow until they collide with the wall and then stretch
and move in radial direction. As soon as the primary toroidal vortex passes the
deceleration area of the wall jet r=D > 0:8 the flow separates and forms a
new secondary counter-rotating ring vortex that enhances the local heat transfer,
directly followed by a likewise annular area of poor heat transfer due to separation.
280 R. Wilke and J. Sesterhenn

a b

Fig. 4 Mean distributions for different radial positions r=D : r=D D 0:5, : r=D D 1:0;
: r=D D 1:5; : r=D D 2:0; : r=D D 3:0; : r=D D 4:5. (a) Radial velocity.
(b) Temperature difference T  To relative to T D TW  To

Fig. 5 Mean distributions for different distances from the lower wall of the radial ur and wall
normal velocity component v. : ur at y=D D 0:01, : ur at y=D D 0:05; : ur at
y=D D 0:1; : r=D D 2:0; : v at y=D D 0:01; : v at y=D D 0:05; : v at
y=D D 0:10
Numerical Simulation of Impinging Jets 281

Travelling downstream the vortex pair becomes unstable and breaks down into
smaller structures that rise. As a consequence the two rings at the wall of very high
and low heat transfer vanish and the cycle restarts.
Figures 6 and 7 show the birth of those vortex rings on a x-y plane through the
center of the jet. The background pictures the Mach number. Thereon contour lines
of same values of the Q-criterion are drawn. The two vertical lines show the position
of the maximal radial velocity. In addition, the Nusselt number is represented on the
x-z plane at the wall. The first picture of the series is taken at a time step where
the primary ring vortex reaches the deceleration area of the wall jet. Beside the
stagnation area no strong peak in heat transfer is present. In the next three pictures,
the secondary counter-rotating ring vortex is born and travels downstream. A strong
annular peak in the Nusselt number can be observed which travels together with the
vortex pair until they collapse and the strong heat transfer disappears.
The investigated life cycle leads to a secondary area located at r=D D 1::1:4 of
high values of the radial Nusselt number
ˇ
D @T ˇ
N u.r/ D  .r/ˇˇ (5)
T @y W

that is shown in Fig. 8a. The primary maximum exists at r=D D 0:18.
Buchlin[2] reported the inner or primary maximum at r=D D 0:7 and the outer
or secondary at r=D D 2:4 for experimental data at Re D 60;000; h=D D 1
with the remark, that the strength of both maxima decrease and the inner one
moves to the stagnation point with increased nozzle to plate distances h=D and
decreased Reynolds number. Since direct numerical simulations presently cannot
reach high Reynolds numbers like Re D 60;000 of the experiments, no direct
comparison between the numeric values can be performed. Nevertheless Buchelins
conclusions concerning the phenomenon of secondary vortex rings, the movement
of the secondary maximum towards the stagnation point and the loss of strength of
both maxima match the results of the present simulation well.

3.3 Turbulent Statistics

Figures 9 and 10 show first results of the Reynolds stresses as well as the turbulent
heat fluxes. Those results are not completely statistically converged and therefore
will be related to the heat transfer and flow phenomena in the final paper.
282 R. Wilke and J. Sesterhenn

Fig. 6 Impact of the secondary vortex on the local Nusselt number illustrated by contours of
equal values of Q above the Mach number M a on a x-y-plane through the center of the jet and the
Nusselt number at the impinging plate (1)
Numerical Simulation of Impinging Jets 283

Fig. 7 Impact of the secondary vortex on the local Nusselt number illustrated by contours of
equal values of Q above the Mach number M a on a x-y-plane through the center of the jet and the
Nusselt number at the impinging plate (2)

a b

Fig. 8 Mean radial distributions for the local Nusselt number Nu and the skin friction coefficient
2W
Cf D u 2 . (a) Nusselt number. (b) Skin friction coefficient
1
284 R. Wilke and J. Sesterhenn

a b

c d

Fig. 9 Reynolds stresses. : y=D D 0:01, : y=D D 0:05; : y=D D 0:1. (a)
Reynolds stress tensor RST   . (b) Reynolds stress tensor RST yy . (c) Reynolds stress tensor RST rr .
(d) Reynolds stress tensor RST yr

a b

Fig. 10 Turbulent heat fluxes. : y=D D 0:01, : y=D D 0:05; : y=D D 0:1. (a)
Turbulent heat flux normal to the wall. (b) Turbulent heat flux in radial direction
Numerical Simulation of Impinging Jets 285

Fig. 11 Scalability of the


code; simulations run with
5123 grid points on CRAY
XE6 (Hermit)

4 High Performance Computing

Investigating physics by means of direct numerical simulation require huge comput-


ing capacity, which can only be provided by the most powerful high performance
computers that are available nowadays. The Kolomgorov’s scales that need to be
resolved lead to those capacities in the order of much more than a million core
hours per computation.
The code is parallelized with MPI libraries. It has been successfully ported to
CRAY XE6 (Hermit) and is typically used with 32  16  16 D 8;192 cores. The
typical run times of the simulation are 24 h whenever available. The code uses auto-
vectorisation. The scalability plot is shown in Fig. 11 and shows nearly perfect linear
scaling for the given problem.

5 Conclusion

This paper presents a direct numerical simulation of a subsonic confined impinging


jet. The chosen grid size of 5123 points matches the requirements for reliable results
on turbulent heat transfer. The Nusselt number distribution is presented and features
a week primary and secondary maximum which is caused by the occurrence of
secondary vortex rings that locally increase the heat transfer in an annular shape.
Those first results agree with experimental data of [2]. In future work a DNS at
Re D 8;000 will be performed and directly compared to experimental results of our
partner project within the Collaborative Research Centre 1029.
286 R. Wilke and J. Sesterhenn

Acknowledgements The simulations were performed on the national supercomputer Cray XE6
at the High Performance Computing Center Stuttgart (HLRS) under the grant number GCS-
NOIJ/12993.
The authors gratefully acknowledge support by the Deutsche Forschungsgemeinschaft (DFG)
as part of collaborative research center SFB 1029 “Substantial efficiency increase in gas turbines
through direct use of coupled unsteady combustion and flow dynamics”.

References

1. Adams, N.: A high-resolution hybrid compact-ENO scheme for shock-turbulence interaction


problems. J. Comput. Phys. 127, 27–51 (1996). http://dx.doi.org/10.1006/jcph.1996.0156.
doi:10.1006/jcph.1996.0156
2. Buchlin, J.: Convective heat transfer in impinging-gas-jet arrangements. J. Appl. Fluid Mech.
4, 3 (2011)
3. Chung, Y.M., Luo, K.H.: Unsteady heat transfer analysis of an impinging jet. J. Heat Trans.
124, 12, (6), 1039–1048 (2002). http://dx.doi.org/10.1115/1.1469522. ISBN 0022–1481
4. Cziesla, T., Biswas, G., Chattopadhyay, H., Mitra, N.: Large-eddy simulation of flow and
heat transfer in an impinging slot jet. Int. J. Heat Fluid Flow 22(5), 500–508 (2001). http://
dx.doi.org/http://dx.doi.org/10.1016/S0142-727X(01)00105-9. doi:http://dx.doi.org/10.1016/
S0142--727X(01)00105--9. ISSN 0142–727X
5. Dairay, T., Fortune, V., Lamballais, E., Brizzi, L.E.: Direct numerical simulation of the
heat transfer of an impinging jet. In: 14th European Turbulence Conference, Lyon, Sept
2013. Department of Fluid Flow, Heat Transfer and Combustion, Institute PRIME. CNRS –
Universite de Poitiers ENSMA
6. Hattori, H., Nagano, Y.: Direct numerical simulation of turbulent heat transfer in plane
impinging jet. Int. J. Heat Fluid Flow 25(5), 749–758 (2004). http://dx.doi.org/http://dx.
doi.org/10.1016/j.ijheatfluidflow.2004.05.004. doi:http://dx.doi.org/10.1016/j.ijheatfluidflow.
2004.05.004. ISSN 0142–727X. Selected papers from the 4th International Symposium on
Turbulence Heat and Mass Transfer
7. Hrycak, P.: Heat transfer from impinging jets. A literature review. New Jersey Institute of
Technology, 1981. Forschungsbericht
8. Jambunathan, K., Lai, E., Moss, M., Button, B.: A review of heat transfer data for
single circular jet impingement. Int. J. Heat Fluid Flow 13(2), 106–115 (1992). http://
dx.doi.org/http://dx.doi.org/10.1016/0142-727X(92)90017-4. doi:http://dx.doi.org/10.1016/
0142-727X(92)90017-4. ISSN 0142–727X
9. Lele, S.K.: Compact finite difference schemes with spectral-like resolution. J. Comput. Phys.
103(1), 16–42 (1992). http://dx.doi.org/10.1016/0021-9991(92)90324-R. doi:10.1016/0021–
9991(92)90324–R
10. Pirozzoli, S., Bernardini, M., Grasso, F.: Characterization of coherent vortical structures in a
supersonic turbulent boundary layer. J. Fluid Mech. 613(10), 205–231 (2008). http://dx.doi.
org/10.1017/S0022112008003005. doi:10.1017/S0022112008003005. ISSN 1469–7645
11. Sesterhenn, J.: A characteristic-type formulation of the Navier-Stokes equations for high order
upwind schemes. Comput. Fluids 30(1), 37–67 (2001). http://dx.doi.org/DOI:10.1016/S0045-
7930(00)00002-5. doi:10.1016/S0045–7930(00)00002–5. ISSN 0045–7930
12. Viskanta, R.: Heat transfer to impinging isothermal gas and flame jets. Exp. Therm. Fluid
Sci. 6(2), 111–134 (1993). http://dx.doi.org/http://dx.doi.org/10.1016/0894-1777(93)90022-
B. doi:http://dx.doi.org/10.1016/0894-1777(93)90022-B. ISSN 0894–1777
13. Weigand, B., Spring, S.: Multiple jet impingement – a review. Heat Trans. Res. 42(2), 101–142
(2011). ISSN 1064–2285
Numerical Simulation of Impinging Jets 287

14. Wilke, R., Sesterhenn, J.: Direct numerical simulation of heat transfer of a round subsonic
impinging jet. In: Active Flow and Combustion Control 2014. Notes on Numerical Fluid
Mechanics and Multidisciplinary Design Conference. Springer, Cham (2014)
15. Zuckerman, N., Lior, N.: Impingement heat transfer: correlations and numerical modeling. J.
Heat Trans. 127(5), 544–552 (2005). http://dx.doi.org/10.1115/1.1861921. ISBN 0022–1481

Vous aimerez peut-être aussi