You are on page 1of 13

Applied Thermal Engineering 25 (2005) 409–421

www.elsevier.com/locate/apthermeng

An experimental and numerical investigation on a


confined impinging air jet at high Reynolds numbers
E. Baydar *, Y. Ozmen
Department of Mechanical Engineering, Karadeniz Technical University, 61080 Trabzon, Turkey
Received 17 September 2003; accepted 29 May 2004
Available online 6 July 2004

Abstract
An experimental and numerical study is carried out to investigate flow field of a confined jet issuing from
the lower surface and impinging normally on the upper surface. The mean velocity, turbulence intensity and
pressure distributions in the impingement region were obtained for Reynolds numbers ranging from 30,000
to 50,000 and a nozzle-to-plate spacing range of 0.2–6. The effects of Reynolds number and nozzle-to-plate
spacing on the flow structure are examined. A subatmospheric region occurs on the impingement plate at
nozzle-to-plate spacings up to 2 for Reynolds numbers studied and it moves radially outward from the
stagnation point with increasing nozzle-to-plate spacing. It is concluded that there exists a linkage among
the subatmospheric region, turbulence intensity and heat transfer coefficients. The numerical results ob-
tained using the standard k–e turbulence model are in agreement with the experimental results except for
the nozzle-to-plate spacings less than one.
 2004 Elsevier Ltd. All rights reserved.

Keywords: Jet impingement; Subatmospheric region; Pressure coefficient; Turbulence; k–e model

1. Introduction

A single air jet or an array of jets, impinging normally on a surface, may be used to achieve
enhanced coefficients for convective heating, cooling or drying. Applications include tempering of
glass plate, annealing of metal sheets, drying of textile and paper products, deicing of aircraft

*
Corresponding author. Tel.: +90-462-3772962; fax: +90-462-3255526.
E-mail address: baydar@ktu.edu.tr (E. Baydar).

1359-4311/$ - see front matter  2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.applthermaleng.2004.05.016
410 E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421

Nomenclature

Cp pressure coefficient
D nozzle diameter (m)
H nozzle-to-plate spacing (m)
Nu Nusselt number
DP difference between the pressure on the impingement surface and the atmospheric
pressure (N/m2 )
r radial distance measured from the jet axis (m)
Re Reynolds number, U0 D=m
u local radial velocity (m/s)
v local axial velocity (m/s)
u0 turbulence velocity (rms) in the radial direction (m/s)
v0 turbulence velocity (rms) in the axial direction (m/s)
U0 nozzle exit velocity (m/s)
z axial distance measured from the nozzle exit (m)
Greeks
m kinematic viscosity of air (m2 /s)
q density of air (kg/m3 )

systems and cooling of heated components in gas turbine engines, computers and electronic
instruments.
The flow and heat transfer of an impinging jet depend on many parameters such as the non-
dimensional nozzle-to-plate spacing (H =D), Reynolds number (Re), Nusselt number (Nu), Prandtl
number (Pr), and the non-dimensional distance from the stagnation point (r=D). In addition, the
effect of the nozzle geometry, flow confinement, turbulence, recovery factor, and the dissipation of
the jet temperature have all been shown to be significant [1].
It was concluded that the impinging jet flow is turbulent for the Reynolds numbers larger than
2000 [2–4]. Goldstein and Behbahani [5] reported that at large nozzle-to-plate spacings the cross
flow diminishes the peak heat transfer coefficient and at small spacings the cross flow can increase
the peak heat transfer coefficient for an impinging air jet. The convective heat transfer by the
impingement of laminar and turbulent circular liquid jets has been studied analytically and
experimentally by Liu et al. [6]. Lytle and Webb [7] experimentally investigated the flow structure
and heat transfer characteristics of air jet impingement for nozzle-plate spacings less than one
nozzle diameter in the range of 3600 < Re < 27; 600. Huber and Viskanta [8] examined the
influence of spent air exits located between the jets on local heat transfer coefficient for a confined
impinging array of air jets. The effect of nozzle geometry on the local convective heat transfer to a
confined impinging air jet was experimentally studied at various nozzle-to-plate spacings by
Colucci and Viskanta [9]. San et al. [10] measured the local Nusselt number of a confined
impinging jet. They found that the Nusselt number is affected due to the flow recirculation and
mixing on the impingement surface. Saad et al. [11] compared the turbulence, mean flow, and heat
E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421 411

transfer characteristics of an array of confined impinging slot jet with those of a single jet. Several
experimental investigations to determine velocity and turbulence characteristics of impinging jets
were made [12–16]. The interaction between impingement jets on flat plate are studied with several
experimental flow visualization techniques by Carcasci [17]. Ichimiya and Yamada [18] have
visualized the flow field of an impinging jet for low spacings and Reynolds numbers using a
thermosensitive liquid crystal. They have reported that, with the increase in Reynolds number and
nozzle-to-plate spacing, the recirculation flow on the impingement surface moves downstream
and its volume increases correspondingly. Fairweather and Hargrave [19] have obtained the mean
and fluctuating velocities and shear stresses in an air jet impinging on a flat plate surface by
particle image velocimetry. They have observed that the existence of a recirculation zone which
carries material from the periphery of the wall jet to its initial region. A combined approach to
characterize the flow field and local heat transfer in jet impingement configurations, featuring a
mass transfer experiment and a digital visualization technique was employed by Angioletti et al.
[20]. They concluded that local heat transfer is strongly influenced by impingement structures.
Narayanan et al. [21] experimentally investigated the flow field, surface pressure and heat transfer
rate of a slot impinging jet. They reported that the interaction between correlated motion in the
outer region and near-wall turbulence causes the rise in Nusselt number towards a secondary
peak.
Many numerical studies have been made on a jet impinging on a flat plate using various tur-
bulence models. Craft et al. [22] used the impinging jet for evaluating Reynolds averaged tur-
bulence models but none of the models was able to yield satisfactory results. Dianat et al. [23]
made predictions of axisymmetric and two-dimensional impinging turbulent jet with a standard
k–e turbulence model. Behnia et al. [24] numerically studied the jet impingement at high Reynolds
number flows with turbulence models. They addressed the importance of near-wall modelling.
Some of the recent numerical studies of turbulent impinging jets have used turbulence models
such as cubic k–e, k–w, DNS or LES [25–28].
An increase in local heat-transfer coefficients was seen as secondary peak at the impinging jet
flows at low nozzle-to-plate spacings. This phenomenon was explained with the transition from
laminar to turbulent flow in wall jet region [7,14,29]. This was also clarified with the presence of a
recirculation region which was supported by the subatmospheric pressure data [9]. It was con-
cluded that the local heat transfer coefficient was enhanced at both the leading and trailing edges
of the recirculation region [30]. Only a few studies, however, paid attention to the subatmospheric
region. Baydar [31] carried out an experimental investigation for low Reynolds numbers up to
10,000 at various nozzle-to-plate spacings. A subatmospheric region was observed on the
impingement plate for Re > 2700 and the nozzle-to-plate spacings less than 2 and that there ex-
isted a linkage between the subatmospheric region and the peaks in local heat-transfer coefficients.
In the existing literature, larger values of nozzle-to-plate spacings (H=D > 1) have been gen-
erally considered. However, the low nozzle-to-plate spacings (H =D < 1) are commonly used in
material processing applications. The present study is concerned with the experimental and
numerical investigation of the confined impinging jet flow fields at various spacings. The main
objective of this study is to investigate the presence of subatmospheric regions in impinging jet
applications with high Reynolds numbers. The relation of this subatmospheric region with the
affecting parameters of the jet flow, such as nozzle-to-plate spacing, H =D, and resulting flow
parameters including pressure distribution and the turbulence intensity are examined. There is a
412 E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421

little numerical study on impingement jets for low nozzle-to-plate spacings. Another purpose of
this study is to asses the standard k–e turbulence model, particularly at lower spacings (H=D < 1).

2. Experimental study

Fig. 1 shows a schematic of the experimental apparatus. Air is discharged from a shaped brass
nozzle which is mounted with its axis vertical in the center of a confined plate. The air supplied to
the nozzle by a centrifugal fan mounted below the confined plate by way of a plenum chamber
containing a gauze screen to ensure a uniform flow towards to nozzle. The air flow is regulated by
a valve. The nozzle is D ¼ 25 mm diameter and 12.7 mm length, of which outlet is slightly
rounded. The impingement plate made of plexiglass was 420 mm diameters and 5 mm thickness,
and it was mounted as perpendicular to the jet axis at various spacings according to the con-
finement plate. The jet issuing from the nozzle, with a velocity U0 , impinges vertically onto the
impingement plate at a distance H from the nozzle and enters the surrounding room air.
Mean velocity and turbulence measurements were obtained together with TSI IFA-100 hot-
wire anemometer interfaced to a data acquisition system using TSI 1220-20 hot film probe. The

Fig. 1. Schematic of the experimental set-up.


E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421 413

probe was calibrated with TSI Model 1125 calibration apparatus. Mean and turbulence velocities
were evaluated from 4096 samples gathered at 1024 samples per second by used a computer for
collecting and analyzing all data. The sampling rate is 1000 Hz, and a low pass filter of 300 Hz is
used for all these procedures. The probe and its support was positioned parallel to the two plates,
and moved in the vertical and lateral directions using a three-dimensional traversing unit with an
accuracy of 0.1 mm. The nearest distance that the probe can be positioned to the walls is 2.3 mm,
since the diameter of the support is 4.6 mm. Velocity measurements were made for H =D spacings
of 1, 2 and 4 and for Reynolds number of 30000. Along the jet axis, centerline mean velocity, v,
and turbulence velocity, v0 , were measured. The measurements of the mean velocity, u, and tur-
bulence velocity, u0 , in the radial direction were obtained. The jet exit velocities were up to 30 m/s
ðM < 0:088Þ, so compressibility effects were negligible. Turbulence intensity at the exit of the
nozzle was less than 0.6%.
The pressure measurements at the impingement surface were made for the nozzle-plate spacings
of 0.2–6 for Reynolds numbers of 30,000, 40,000 and 50,000. The pressure distributions at the
impingement surface were obtained by moving the impingement plate with respect to the jet, using
a traversing unit with an accuracy of 0.1 mm. The surface pressures were measured with a pressure
tapping 0.5 mm in inner diameter placed at the centre of the impingement plate. This tapping was
connected to a micronanometer giving pressure reading with an accuracy of 0.01 mm H2 O was
used to determine the surface pressures. The Reynolds number depends on the nozzle diameter, D,
and the nozzle exit velocity, U0 . The pressure coefficient, Cp , is defined as DP =ðqU02 =2Þ, where DP
represents the difference between the local pressure on the impingement surface and the atmo-
spheric pressure while qU02 =2 represent the dynamic pressure.

2.1. Experimental uncertainty

In order to question the validity of the results obtained experimentally, an uncertainty analysis
is made [32]. The uncertainty for the Reynolds numbers is ±2%, and the uncertainty in the
measurements of the axial velocity was estimated to be less than ±5%. Turbulence velocities, u0
and v0 , have a corresponding estimated uncertainty of ±6% and ±4%, respectively. The last
location near the impingement plate, the uncertainty is found to be largest. Uncertainty in
pressure measurements at the impingement surface was about ±1%. The experimental results were
reproducible within these uncertainty ranges.

3. Numerical study

Numerical study is made to examine the performance of the standard k–e turbulence model,
particularly at low spacings (H =D < 1). The turbulent flow field of an impinging jet is assumed to
be axisymmetric. The governing equations written in cylindrical coordinates for a steady, tur-
bulent and incompressible flow are given in Table 1.
( "   2   #  2 )
2
ou ov v 2 ou ov
lt ¼ Cl qk 2 =e; leff ¼ l þ lt ; G  leff 2 þ þ þ þ
oz or r or oz
414 E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421

Table 1
Governing equations
! !
o 1 o o o/ 1 o o/
ðqu/Þ þ ðrqv/Þ  C/  rC/ ¼ S/
oz r or oz oz r or or

/ C/ S/
   
u leff op o ou 1 o ov
 þ leff þ rleff
oz oz oz r or oz
   
v leff op o ou 1 o ov
 þ l þ rleff
or oz eff or r or or
k leff =rk G  qe
e leff =re e e2
C1 G  C2 q
k k

The following values are used for the coefficients in the above equations, as described by Launder
and Spalding [33].
Cl ¼ 0:09; C1 ¼ 1:44; C2 ¼ 1:92; rk ¼ 1; re ¼ 1:3:
At the jet exit, measured values of velocity and turbulence intensity were specified. At the solid
surface, a logarithmic wall function was used. At the outflow boundary, a constant static pressure
and zero radial gradients of all properties are assumed.
In the numerical study, a computer program developed in the finite difference technique was
used. The basic equations are solved numerically by SIMPLEC method of Van Doormaal and
Raithby [34] based on finite volume technique on staggered grids. Hybrid difference scheme is
adopted for the convective terms in the momentum equations. The discretised equations are
solved iteratively using the tridiagonal matrix algorithm line solver. A mesh structure which is
uniform in the axial direction and non-uniform in the radial direction is used. The mesh used is
refined for each value of H =D until negligible differences are obtained. For instance, the geometry
of the mesh is found to be 22 · 142 for H=D ¼ 0:2 while it is 142 · 142 for H=D ¼ 6. Iterations
were continued until the residual for each equation was less than 1 · 105 .

4. Result and discussion

Experimental and numerical investigations were carried out for the nozzle-to-plate spacings of
0.2–6 and Reynolds numbers of 30,000, 40,000 and 50,000 in confined impinging jet. Mean and
turbulence velocities have been non-dimensionalized by the nozzle exit velocity, U0 . The axial
distance from the nozzle, z, has been normalized by the nozzle-to-plate spacing, H .
Fig. 2(a) illustrates the measured and computed axial velocity profiles along the jet centerline
for Re ¼ 30; 000 and H =D ¼ 1, 2 and 4. It is seen from the figure, the experimental and numerical
data are in good agreement. The presence of the impingement plate gives rise to the deceleration
and deflection of the flow. Increasing H =D broadens the velocity profile as a result of enhanced
diffusion of the jet flow. The potential-core length can be quantified by the distance from the jet
E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421 415

0.8

0.6
v/Uo
H/D=1, Num.
0.4 H/D=2, Num.
H/D=4, Num.
H/D=1, Exp.
0.2
H/D=2, Exp.
0 H/D=4, Exp.
(a)

0.05
H/D=1
H/D=2
0.04
H/D=4
v /Uo

0.03

0.02

0.01

0 (b)
0 0.2 0.4 0.6 0.8 1
z/H

Fig. 2. (a) The axial velocity profiles along the jet centerline for Re ¼ 30; 000. (b) The variations of turbulence intensity
along the jet centerline for Re ¼ 30; 000.

exit where the jet velocity retards to 95% of its exit velocity [35]. In the absence of the impingement
plate, that is in the case of free jet, the potential core length was measured approximately as 3:2D.
This result agrees well with the results of previous studies in which this length is given between 3D
and 6D for a free jet [36]. In existence of impingement plate, the potential core lengths are
approximately 0:3H , 0:6H and 0:8H for H =D ¼ 1, 2 and 4, respectively. As expected, increased
values of H =D resulted in increased values of the potential core lengths. For spacing of H =D ¼ 4,
the potential core length reaches to the largest value which is the potential core length of the free
jet. It is seen from the figure that although the numerical predictions have the same trend with the
experimental data, the potential core lengths obtained from numerical predictions are quite
smaller than those obtained from experimental data.
For the same values, the variation of turbulence intensity measured along the jet centerline are
shown in Fig. 2(b). The turbulence intensity remains nearly constant within the potential core [12].
It is seen that the turbulence intensity remains nearly constant up to z=H ¼ 0:3 and 0.6 for
H =D ¼ 1 and 2, respectively. These axial distances also correspond to the potential core lengths
for each spacing shown in Fig. 2(a). As the impingement surface is approached, in contrast to the
velocities, centerline turbulence levels increase throughout the development of the jet. Down-
stream of the potential core region, the variation of turbulence intensity with the axial distance
416 E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421

exhibits different characteristics depending on the value of H =D. There is no significant variation
in turbulence intensity for H =D ¼ 1 and 2 whereas it increases strongly with increasing z=H for
H =D ¼ 4. This is due to the jet spreading with increasing spacing.
The effect of Reynolds number on pressure coefficient at a specific value of H =D is shown in a
separate figure starting from Fig. 3(a)–(d). It is observed from the figures, in general, that the
pressure distributions on the impingement surface are independent of the Reynolds number, while
they depend strongly on the nozzle-to-plate spacing. The region near the impingement plate is
often referred to as the deflection zone where there is a rapid decrease in axial velocity and a

3
Re=30000, Exp.
Re=40000, Exp.
2
Re=50000, Exp. 1.2
Re=30000, Num. Re=30000, Exp.
1 Re=40000, Num. 1 Re=40000, Exp.
Re=50000, Num. Re=50000, Exp.
0.8 Re=30000, Num.
0 Re=40000, Num.
Re=50000, Num.
0.6

-1
0.4
Cp

-2
Cp

0.2

0
-3

-0.2
-4
-0.4

-5
-0.6

-6 -0.8
0 1 2 3 4 5 0 1 2 3 4 5
(a) r/D (b) r/D

1.2
1
Re=30000, Exp. Re=30000, Exp.
Re=40000, Exp. 1 Re=40000, Exp.
0.8 Re=50000, Exp. Re=50000, Exp.
Re=30000, Num.
0.8 Re=30000,Num.
Re=40000, Num.
0.6 Re=50000, Num. Re=40000,Num.
Re=50000,Num.
0.6
Cp
Cp

0.4
0.4

0.2
0.2

0
0

-0.2
-0.2
0 1 2 3 4 5
0 1 2 3 4 5
r/D
(c) (d) r/D

Fig. 3. The radial distributions of the pressure coefficient. (a) H =D ¼ 0:2. (b) H =D ¼ 0:6. (c) H =D ¼ 1. (d) H =D ¼ 2.
E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421 417

corresponding rise in static pressure. At the stagnation point (r=D ¼ 0), the velocity is zero and
the pressure is a maximum, and then the pressure decrease as the flow accelerates along the
impingement surface. This case is consistent with global continuity considerations. There
observed a subatmospheric region on the impingement surface for spacings up to 2. The sub-
atmospheric region becomes stronger and more deductible with decreasing nozzle-to-plate spac-
ing. As nozzle-to-plate spacing increases the fluid velocity decreases due to the jet spreading, and
thus the strength of the subatmospheric region decreases. Similar findings were also obtained for a
confined impinging jet at low Reynolds numbers (Re < 10; 000) by Baydar [31]. It is clear from the
figures that stagnation pressure is independent of Reynolds number on the entire range of H =D.
While the nozzle-to-plate spacing increases, stagnation pressure decreases. The experimentally
obtained pressure coefficients at the stagnation point are greater than one for H =D < 1, remain
constant approximately at one in the range of 1 6 H=D 6 2. For low spacings (H =D < 1), near the
stagnation point, the flow rapidly decelerates and therefore the pressure increases significantly.
Although the measured and computed data agree well for H=D > 1, significant differences be-
tween these data are observed for H =D 6 1. The k–e model predicts too low local pressures for
H =D < 1. Discrepancies between calculated and measured results for spacings of H =D < 1 are
originated from turbulence model used which is insufficient when the flow is strongly curved.
It was shown [7,9,29] that the local Nusselt number depends strongly on H =D, and varies
significantly throughout the impingement plate at low values of H =D (H =D 6 1). The existence of
the secondary peak in Nusselt number data is mainly explained by transition from laminar to
turbulent flow in wall jet region [7,14,29]. Asforth-Frost et al. [12] and Cooper et al. [13] con-
cluded that the secondary peak in Nusselt numbers is associated with the increase of the wall-
adjacent turbulence levels. Colucci and Viskanta [9] measured the local pressures and the heat

Fig. 4. The variation of subatmospheric locations for Re ¼ 30; 000.


418 E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421

transfer coefficients on the impingement surface for H =D ¼ 0:25, 1 and 6, and Re ¼ 10; 000,
30,000 and 50,000, using an orifice similar to that for present study. They stated that a subat-
mospheric region was observed at H=D ¼ 0:25 (not plotted in Fig. 4(a) of Ref. [9]) at which a
secondary peak in Nusselt number data is observed in Fig. 5(a). From the same figures (Figs. 4(a)
and 5(a) of Ref. [9]) it is clear that there is no subatmospheric region at H =D ¼ 6 and there are no
peaks in Nusselt number distribution either. For H=D ¼ 1, on the other hand, there is no indi-
cation of the existence of a subatmospheric region in Fig. 4(a), in spite of the existence of the
secondary peak appearing in Fig. 5(a). In this study, a subatmospheric region has been measured
for H =D ¼ 1. Thus, we may conclude that there exists a linkage between the secondary peak in
local Nusselt number and subatmospheric region.

Fig. 5. (a) The radial distributions of the pressure coefficient for Re ¼ 30; 000. (b) The radial distribution of the tur-
bulence intensity for Re ¼ 30; 000. (c) The radial distributions of the Nusselt number for Re ¼ 30; 000 [9].
E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421 419

As shown above, the local pressure on the impingement surface becomes subatmospheric at a
certain radial distance from the stagnation point for nozzle-to-plate spacings up to H =D ¼ 2. The
variation of the location at which local pressure becomes subatmospheric with H=D is presented
in Fig. 4. Since it has been already disclosed that the Reynolds number has no effect on the
pressure distribution at the impingement surface, for the brevity, only results for Re ¼ 30; 000 is
shown. As seen, with the increasing nozzle-to-plate spacing, the subatmospheric region moves
radially outward from the stagnation point. A similar behaviour has been also observed in the
experimental study of Ichimiya and Yamada [18]. Although there is a little difference between the
numerical and the experimental results, they show the similar trend.
Fig. 5 show the radial variation of the pressure coefficient, the turbulence intensity and the
Nusselt number for Re ¼ 30; 000. The values of Fig. 5(a) and (b) are experimental, which are
obtained in this study, while Fig. 5(c) represents the experimentally obtained Nusselt number
values reported by Colucci and Viskanta [9]. For H =D ¼ 1, the pressure becomes subatmospheric
at r=D  1:1 and remains as subatmospheric up to r=D  5. Turbulence intensity measured at
z=H ¼ 0:75 has a maximum value at r=D  1:7 within this subatmospheric region and it shows a
local increase about r=D ¼ 0:6 due to rapidly acceleration of the flow in the deflection region just
after the stagnation point. For the same spacing, Colucci and Viskanta [9] found two local peaks
occurring at locations of r=D  0:6 and r=D  1:6 in Nusselt number. The locations of the inner
and secondary peaks in Nusselt number correspond very well to the locations of the peaks in
turbulence intensity. Similar findings were also obtained by Lytle and Webb [7], and Fitzgerald
and Garimella [14]. Secondary peak in Nusselt number and maximum turbulence intensity occur
nearly at the same location. Thus, it may be concluded that the maximum turbulence intensity and
secondary peak in Nusselt number are in the subatmospheric region. Furthermore, it is seen that
the location of secondary peak measured by Colucci and Viskanta [9] for H=D ¼ 0:25 is in the
subatmospheric region observed for H=D ¼ 0:2 in the present study. For spacing of H =D ¼ 6, it is
clear that there is no subatmospheric region and there is no peak in Nusselt number. It can be seen
that there is a certain correlation among the location of peak turbulence intensity and the sub-
atmospheric region and a secondary peak in the Nusselt number.

5. Conclusions

Experimental and numerical investigations were carried out on the impinging jet flow for high
Reynolds numbers up to 50,000 at various distances between the jet exit and impingement plate.
The characteristics of an impinging circular jet in a confined region were found to be sensitive to
the nozzle-to-plate spacing. Due to the presence of the impingement plate causing the decelera-
tion, the flow deflects at about a jet diameter above the impingement plate. At H =D < 2, a
subatmospheric region occurs on the impingement plate in jet applications. The subatmospheric
region becomes stronger with decreasing nozzle-to-plate spacing and it moves radially outward
from the stagnation point with increasing nozzle-to-plate spacing. Turbulence intensity takes two
maxima within this region. When compared to the results associated with heat transfer [7–9], it is
seen that there exists a linkage among the subatmospheric region in pressure distributions, the
peaks in turbulence intensity and the peaks in heat transfer coefficients on the impingement plate.
420 E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421

The experimental data in this study could be useful in developing and assessing turbulence models
for impinging jet flows.
In the numerical part of the study, the standard k–e turbulence model is used to predict flow
field characteristics. From the comparisons of numerical results with the experimental ones, it is
found that the model is capable of predicting the flow characteristics correctly in moderate values
of nozzle-to-plate spacing.

Acknowledgements

This work was sponsored by the Turkey-State Planning Organization (DPT) under Grant no.
2003.200.200.5. The authors wish to thank Dr. O. Aydin, Dr. Y.E. Akansu and Dr. A. Unal for
their valuable contributions.

References

[1] K. Jambunathan, E. Lai, M.A. Moss, B.L. Button, A review of heat transfer data for single circular jet
impingement, Int. J. Heat Fluid Flow 13 (1992) 106–115.
[2] V.A. Marple, B.Y.H. Liu, K.T. Whitby, On the flow fields of inertial impactors, J. Fluids Eng. Trans. ASME 96
(1974) 394–400.
[3] M.D. Deshpande, R.N. Vaishnav, Submerged laminar jet impingement on a plane, J. Fluid Mech. 114 (1982) 213–
236.
[4] S. Polat, B. Huang, A.S. Mujumdar, W.J.M. Douglas, Numerical flow and heat transfer under impinging jets: A
review, in: C.L. Tien (Ed.), Ann. Rev. Numer. Fluid Mech. and Heat Trans., vol. 2, Hemisphere Publishing Corp.,
1989, pp. 157–197.
[5] R.J. Goldstein, A.I. Behbahani, Impingement of a circular jet with and without cross flow, Int. J. Heat Mass
Transfer 25 (1982) 1377–1382.
[6] X. Liu, V.J.H. Lienhard, J.S. Lombara, Convective heat transfer by impingement of circular liquid jets, J. Heat
Transfer, Trans. ASME 113 (1991) 571–582.
[7] D. Lytle, B.W. Webb, Air jet impingement heat transfer at low nozzle-plate spacings, Int. J. Heat Mass Transfer 37
(1994) 1687–1697.
[8] A.M. Huber, R. Viskanta, Convective heat transfer to a confined impinging array of air jets with spent air exits,
ASME J. Heat Transfer 116 (1994) 570–576.
[9] D.W. Colucci, R. Viskanta, Effect of nozzle geometry on local convective heat transfer to a confined impinging air
jet, Exp. Thermal Fluid Sci. 13 (1996) 71–80.
[10] J.Y. San, C.H. Huang, M.H. Shu, Impingement cooling of a confined circular air jet, Int. J. Heat Mass Transfer 40
(1997) 1355–1364.
[11] N.R. Saad, S. Polat, W.J.M. Douglas, Confined multiple impinging slot jets without crossflow effects, Int. Heat
Fluid Flow 13 (1992) 2–14.
[12] S. Ashfort-Frost, K. Jambunathan, C.F. Whithney, Velocity and turbulence characteristics of a semiconfined
orthogonally impinging slot jet, Exp. Thermal Fluid Sci. 14 (1997) 60–67.
[13] D. Cooper, D.C. Jackson, B.E. Launder, G.X. Liao, Impinging jet studies for turbulence model assessment-I.
Flow-field experiments, Int. J. Heat Mass Transfer 36 (1993) 2675–2684.
[14] J.A. Fitzgerald, S.V. Garimella, Flow field effects on heat transfer in confined jet impingement, J. Heat Transfer
Trans. ASME. 119 (1997) 630–632.
[15] L.L. Dong, C.W. Leung, C.S. Cheung, Heat Transfer and pressure characteristics of a twin premixed butane/air
flame jets, Int. J. Heat Mass Transfer 47 (2004) 489–500.
E. Baydar, Y. Ozmen / Applied Thermal Engineering 25 (2005) 409–421 421

[16] S.D. Hwang, H.H. Cho, Effects of acoustic excitation positions on heat transfer and flow in axisymmetric
impinging jet: main jet excitation and shear layer excitation, Int. Heat Fluid Flow 24 (2003) 199–209.
[17] C. Carcasci, An experimental investigation on air impinging jets using visualization methods, Int. J. Therm. Sci. 38
(1999) 808–818.
[18] K. Ichimiya, Y. Yamada, Three dimensional heat transfer of a confined circular impinging jet with buoyancy
effects, J. Heat Transfer, Trans. ASME 125 (2003) 250–256.
[19] M. Fairweather, G.K. Hargrave, Experimental investigation of an axisymmetric, impinging turbulent jet. 1.
Velocity field, Exp. Fluids 33 (2002) 464–471.
[20] M. Angioletti, R.M. Di Tommaso, E. Nino, G. Ruocco, Simultaneous visualization of flow field and evaluation of
local heat transfer by transitional impinging jets, Int. J. Heat Mass Transfer 46 (2003) 1703–1713.
[21] V. Narayanan, J. Seyed-Yagoobi, R.H. Page, An experimental study of fluid mechanic and heat transfer in an
impinging slot jet flow, Int. J. Heat Mass Transfer 47 (2004) 1827–1845.
[22] T.J. Craft, L.J.W. Graham, E.B. Launder, Impinging jet studies for turbulence model assessment-II. An
examination of the performance of four turbulence models, Int. J. Heat Mass Transfer 36 (1993) 2685–2697.
[23] M. Dianat, M. Fairweather, W.P. Jones, Predictions of axisymmetric and two-dimensional impinging turbulent
jets, Int. J. Heat Fluid Flow 17 (1996) 530–538.
[24] M. Behnia, S. Parneix, Y. Shabany, P.A. Durbin, Numerical study of turbulent heat transfer in confined and
unconfined impinging jets, Int. Heat Fluid Flow 20 (1999) 1–9.
[25] B. Merci, E. Dick, Heat transfer predictions with a cubic k–e model for axisymmetric turbulent jets impinging onto
a flat plate, Int. J. Heat Mass Transfer 46 (2003) 469–480.
[26] T.H. Park, H.G. Choi, J.Y. Yoo, S.J. Kim, Streamline upwind numerical simulation of two-dimensional confined
impinging slot jets, Int. J. Heat Mass Transfer 46 (2003) 251–262.
[27] M. Tsubokura, T. Kobayashi, N. Taniguchi, W.P. Jones, A numerical study on the eddy structures of impinging
jets excited at the inlet, Int. Heat Fluid Flow 24 (2003) 500–511.
[28] F. Beaubert, S. Viazzo, Large eddy simulations of plane turbulent impinging jets at moderate Reynolds numbers,
Int. Heat Fluid Flow 24 (2003) 512–519.
[29] J.D.A. Walker, C.R. Smith, A.W. Cerra, T.L. Doligalski, Jet impact of a vortex ring on a wall, J. Fluid Mech. 181
(1987) 99–140.
[30] D.M. Schafer, S. Ramadhyani, F.P. Incropera, Numerical simulation of laminar convection heat transfer from an
in-line array of discrete sources to a confined rectangular jet, Numer. Heat Transfer 22 (1992) 121–141.
[31] E. Baydar, Confined impinging air jet at low Reynolds numbers, Exp. Thermal Fluid Sci. 19 (1999) 27–33.
[32] Model 1125/1125R-1 Calibrator/Probe Rotator, Instruction Manuel, T.S.I. Incorporated, P/N 1990113 Revision F,
St. Paul, 1987.
[33] B.E. Launder, D.B. Spalding, The numerical computation of turbulent flows, Comput. Meth. Appl. Mech. Eng. 3
(1974) 269–289.
[34] J.P. Van Doormaal, G.D. Raithby, Enhancements of the SIMPLE method for predicting incompressible flows,
Numer. Heat Transfer 7 (1984) 147–163.
[35] J.N.B. Livingood, P. Hrycak, Impingement heat transfer from turbulent air stream jets to flat plates: A literature
survey, NASA TM X-2778, 1973.
[36] S. Maurel, C. Solliec, A turbulent plane jet impinging nearby and from a flat plate, Exp. Fluids 31 (2001) 687–696.

You might also like