You are on page 1of 31

In: Cooling Systems: Energy, Engineering and Applications

ISBN 978-1-61209-379-6
Editor: Aaron I. Shanley
2011 Nova Science Publishers, Inc.
The exclusive license for this PDF is limited to personal website use only. No part of this digital document
may be reproduced, stored in a retrieval system or transmitted commercially in any form or by any means.
The publisher has taken reasonable care in the preparation of this digital document, but makes no expressed
or implied warranty of any kind and assumes no responsibility for any errors or omissions. No liability is
assumed for incidental or consequential damages in connection with or arising out of information contained
herein. This digital document is sold with the clear understanding that the publisher is not engaged in
rendering legal, medical or any other professional services.

Chapter 2

APPLICATIONS OF IMPINGEMENT JET


COOLING SYSTEMS
Hyung Hee Cho1, Kyung Min Kim
and Jiwoon Song
Department of Mechanical Engineering, Yonsei University,
Seoul, 120-749, Korea

ABSTRACT
This book chapter discussed impingement jet cooling. Jet impingement achieves
locally high heat transfer on an interested surface. For these reasons, the impinging jet
cooling technique has been widely used in many industrial systems such as gas turbine
cooling, rocket launcher cooling and high-density electrical equipment cooling in order to
remove a large amount of heat. In this chapter, experimental and numerical investigations
are reviewed on flow and heat transfer characteristics of impinging jets. The review
included the general single jet and jet impingement; their active controls; high speed jet
flows and jet impingement; liquid impinging jet cooling; array jet impingement; array jet
impingements on effusion surface. In detail, to enhance the heat transfer in single and
array jets, there is discussion on design and control of nozzle geometry, nozzle insertion,
jet vibration, secondary injection, and suction flow. In addition, effects of various factors
have been concerned with different operating conditions (crossflow, rotation,
compressible flow, working fluid, etc.) and combined techniques (rib turbulator, pin fin,
dimple, effusion hole, etc.).

1. INTRODUCTION
Impinging jets, which are used in many applications, can produce high heat/mass transfer
rates. It is easy to adjust the location of interest and to remove a large amount of heat on the
impingement surface. For these reasons, the impinging jet technique has been widely used in
cooling, heating, and drying systems for many engineering applications such as gas turbine
1

Corresponding author. Tel.: +82 2 2123 2828, Fax.: +82 2 312 2159. E-mail address: hhcho@yonsei.ac.kr.

38

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

cooling, drying of paper or textiles, processing of steel or glass, high-density electrical


equipment cooling, freezing of cryogenic tissue among others. In addition to their
applications, jet-impingement cooling has been of interest to many fields of fluid dynamics
and thermal engineering Figure 1.

Figure 1. Various applications of impinging jets.

Numerous experimental and numerical investigations have been conducted on heat


transfer and cooling of impinging jets in order to control heat transfer and improve cooling
performance in single and array jets with subsonic to supersonic jet velocities. Many
researchers have performed a multitude of experiments by considering the nozzle geometry,
impinging surface, and active methods such as jet vibration, secondary injection, and suction
flow. In addition, various factors concerned with operating conditions (crossflow, rotation,
mass flow rate, jet velocity, working fluid, etc.) and combined techniques (rib turbulator, pin
fin, dimple, effusion hole, etc.) have been investigated in order to identify the optimum
condition and geometry.
A large number of experimental methods including the liquid crystal method, IR camera
technique, and naphthalene sublimation method have been used to obtain detailed heat
transfer distributions. Given the improvements in computational performance, many
numerical calculations have been performed to obtain local heat transfer distributions in
complex geometries, because the numerical results are still disagreement with the
experimental result in local distribution.
There have been many studies and literature reviews on impinging jets. Martin [1]
reviewed investigations on the hydrodynamics of impinging flow along with the variables and
boundary conditions of heat and mass transfer, local variations in transfer coefficients for
single nozzles and nozzle arrays, integral mean transfer coefficients, and the influence of
outlet flow conditions on transfer coefficients for nozzle arrays. Viskanta [2] reviewed heattransfer characteristics of single and multiple isothermal turbulent air and flame jets
impinging on surfaces. The review focused on applications to materials or comparisons of
theory and experiments. The stagnation point heat transfer in this paper was described in a
similar way, despite the many differences in the jet characteristics (i.e., axial velocity and

Applications of Impingement Jet Cooling Systems

39

turbulence intensity) of isothermal and flame jets. Han and Goldstein [3] presented a review
of jet-impingement heat transfer in gas turbine systems. They reviewed characteristics of the
different flow regions for submerged jets such as the free jet, stagnation flow, and wall-jet
regions. In addition, they discussed heat-transfer characteristics of both single and multiple
jets considering the effects of parameters important to gas turbine systems including the
curvature of surfaces, crossflow, angle of impact, and rotation. Weigand and Spring [4]
summarized relevant experimental and numerical results on multi-jet-impingement heat
transfer including the latest developments in the literature. They provided profound
knowledge on the design of such configurations complemented by a structured listing of
factors influential for heat transfer. This paper included the physics of multiple jet
configurations; an introduction to the characteristics of flow field and heat transfer in multiple
jets with regard to the effects of jet pattern, jet diameter, or open area; crossflow effects,
separation distance, jet-to-jet spacing, and optimization of impingement arrays.
In this chapter, recent research trends on impinging jets are reviewed. We focus on
summarizing single impingement jets and impingement jet arrays in which air/liquid will be
issued into air/liquid. We also consider the effects of air jet velocities of low to high Reynolds
number on flow and heat-transfer characteristics in a single air jet.

2. HEAT TRANSFER VARIABLES IN IMPINGING JETS


The Nusselt number is widely used in the heat transfer as a dimensionless form of heat
transfer coefficient. The variable is defined as
Nu =

(1)

where h is the heat transfer coefficient, D is the nozzle diameter, and k is the thermal
conductivity of the fluid. Here,
=

(2)

where qw is the wall heat flux, Tw is the wall temperature, and Tref is a reference temperature.
Normally, Tref is either the total temperature of jet flow (Tf) or the adiabatic wall temperature
(Taw). The latter is obtained from a recovery factor (r) defined by a non-dimensional form.
=


2 /2

(3)

where Uf is the jet velocity and Cp is the specific heat of the fluid at constant pressure.
The Nusselt number changes depending on whether Tf or Taw is chosen as the Tref. For a
low Reynolds number flow, two results are the same. However, the results show discrepancy
for high Reynolds numbers. Therefore, the reference temperature must be carefully chosen.
Detailed explanations are provided by Goldstein et al. [5-6].

40

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

3. SINGLE JET FLOW AND JET IMPINGEMENT


In general, the flow structure of an impinging jet consists of the free jet, stagnation flow,
and the wall-jet regions as illustrated in Figure 2. The free jet region is divided into the flow
development and fully developed regions. A potential core in the flow development region is
observed when the height (H) to nozzle diameter (d) is less than about 6. The width of the
potential core in free jet flow is reduced moving downstream by the shear flow. In the fully
developed jet flow, the velocity profile is approximated as a Gaussian distribution. The jet
flow impinging on a surface forms the stagnation flow region in which the radial velocity
increases rapidly as the surface is approached. Subsequently, the outward radial flow
generates the wall-jet region.
The heat transfer distributions are non-uniform due to these complex flow patterns, such
as flow deceleration, acceleration, transition, and etc. Energy separation of temperature
field appears in the impinging jet due to a strong coherent structure of ring vortices around the
jet [3]. Basic to advanced experiments and simulations have been conducted to understand the
heat transfer characteristics on impinging jets.
Liquid

Nozzle

V
Free jet region

Potential core

Boundary
layer

Stagnation region

Wall jet region

x or r

Impingement surface (target surface)


Figure 2. Schematic diagram of flow region around an impinging jet.

Hwang [7] showed the effect of jet spacing on the radial distribution of the Nusselt
numbers. For small nozzle-to-plate spacings of 2 and 4, there are two peaks in the Nusselt
number distribution as shown in Figure 3. The first peak at an r/D of less than 0.5 is
generated by the effect of stagnating flow acceleration. This acceleration reduces the
boundary layer thickness, resulting in an increase in heat transfer. The secondary peak
appears at an r/D of approximately 2. Generally, it is thought that the secondary peak in the
Nusselt number is induced by the flow transition from laminar to turbulent and by secondary
vortices. The vortices near the wall disturb the boundary layer flow and enhance the mixing
of ambient fluids, creating the peak in the Nusselt number. As the gap distance increases and
the location of the impingement plate moves outside of the potential core, there is only one
peak at the stagnation point. The reason is that the jet flow is fully developed and the

Applications of Impingement Jet Cooling Systems

41

turbulence intensity at the stagnation region is sufficiently high. In addition, Lee and Lee [8]
measured local heat transfer in a stagnation region on a heated flat plate for an axisymmetric
impinging jet. They reported the Nusselt number correlation at the stagnation point consisted
of Reynolds numbers and nozzle-to-plate spacings.
170

x/ D=2
x/ D=4
x/ D=6
x/ D=8
x/ D=12
x/ D=16

150

Nu

130

110

90

70

50
0

r /D

Figure 3. Local heat transfer distributions in different impinging jet spacing [7]. (D=nozzle exit
diameter, r=radial direction, x=nozzle-to-plate distance).

Measurements of the velocity and turbulence distributions are important in the design of
impinging jets because local heat-transfer distributions change with the velocity and the
turbulence of the impinging jet. Gardon and Akfirat [9] concluded that heat-transfer
phenomena can be explained as effects of the intense and spatially varying turbulence in jets.
Popiel and Trass [10] showed ring-shaped wall eddies induced by large-scale toroidal vortices
from an impinging round jet using the smoke-wire flow-visualization technique. They
reported that these wall eddies for low nozzle-to-plate separations, induced consecutively at a
transition zone, were responsible for the additional enhancement of local momentum and heat
or mass transfer. Jambunathan et al. [11] investigated the parameters affecting flow and heattransfer characteristics of impinging jets. They reported that heat transfer by jet flow on an
impinged surface is affected by nozzle geometry, flow confinement, turbulence intensity,
recovery factor, and dissipation of jet temperature, as well as many other parameters (Re, Pr,
non-dimensional nozzle-to-plate spacing, and non-dimensional displacement from the
stagnation point). Chung and Luo [12] and ODonovan and Murray [13] have studied
unsteady heat-transfer characteristics using transient measurements and calculations of flow
velocity and intensity.
Many researchers who study heat-transfer improvement in impinging jets have used an
effective passive-control technique. For example, Hwang and Cho [14] installed multi-tube
insertion in the nozzle and examined the vortex flow interaction between inserted multi-tube
jet flows (Figure 4). Bilen et al. [15] studied differences about the round jets with different
swirling inserts. Lee and Lee [16] studied heat-transfer enhancement using an elliptic jet for

42

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

engineering applications having a small nozzle-to-surface spacing less than the potential core
length. They reported that the heat transfer rate was larger than that for the axisymmetric jet
(AR = 1) in the stagnation region at smaller nozzle-to-surface spacings (H/D<4) at higher
nozzle aspect ratios (ARs) for the elliptic jet. Zhou and Lee [17] installed a mesh screen in
front of the jet nozzle and performed experiments varying the solidity of the screen. The
results showed that for nozzle-to-surface spacings of H/D<4, the turbulence intensity
increased with mesh solidity, causing an increase in the local heat transfer rate. Gau et al. [18]
added arrays of triangular tabs to the exit of a turbulent round impinging jet issuing from a
long pipe. They found that local heat transfer increased by more than 25% in a series of
distinct regions surrounding the impingement region at small nozzle-to-surface distances.

(a)

(b)

(c)

(d)

Figure 4. Multi-tube insertion in a jet nozzle exit: (a) schematic; (b) no insertion; (c) l=3 and d=1; (d)
l=3 and d=4 [14].

In order to apply impinging jets to cooling systems, experimental studies have been
performed on various impinging surface conditions as illustrated in Figure 5. Yan and Saniei
[19] and Beitelmal et al. [20] investigated the effect of an inclined flat surface. They showed
that the region of maximum heat transfer shifted towards the uphill side of the plate and the
maximum Nusselt number decreased with decreasing inclination angle.
Single round jet

Confined jet

Single slot jet

Jet with cross flow

Oblique jet

Concave/convex surface

Figure 5. Configurations of jet impingement.

The location of the maximum heat transfer region was found to be insensitive to the
Reynolds number in the range used in their studies. Nakabe et al. [21] conducted flow-

Applications of Impingement Jet Cooling Systems

43

visualization and heat transfer experiments for a duct flow including longitudinal vortices in a
conventional impinging jet with crossflow. Their experiment demonstrated that jet
impingement is effective for heat transfer augmentation even in a case with crossflow, and
confirmed the generation of longitudinal vortices by the inclined jet.
Cornaro et al. [22] investigated impinging jets on both concave and convex surfaces
using smoke-wire flow visualization. They showed that flow over a concave surface was less
stable than that over a convex surface, because the flow upstream of the concave surface was
strongly affected by the spent fluid from the surface into recirculation. This spent fluid
entrained in the primary jet flow (re-entrainment), reducing the likelihood of stable ring
vortices and the ability of cooling performance. Heat transfer characteristics of a heated slot
jet impinging on a semicircular convex surface were investigated by Chan et al. [23].
Meanwhile, Choi et al. [24] studied the flow and heat transfer characteristics of a slot jet
impinging on a semicircular concave surface. They laid emphasis on interpreting the heat
transfer data in association with the measured mean velocity and velocity fluctuations of the
impinging and evolving wall-jet region along the concave surface in particular. Lim et al. [25]
and Lee et al. [26] investigated the heat transfer characteristics on convex and concave
surfaces and changed the surface angle between impinging jet and concave surface from 0 to
40. They reported that the displacement of the maximum Nusselt number from the
stagnation point increased with increasing surface angle or decreasing nozzle-to-surface
distance.
Gau and Lee [27] performed experiments on slot air-jet impingement-cooling along
triangular rib-roughened walls. The geometric shape of the triangular ribs was more effective
in deflecting the wall jet away from the wall than in the case with rectangular ribs. Pyo et al.
[28] examined an impinging jet on a flat surface, which had a set of hybrid rods with a semicircular cross section. They reported that the installed hybrid rods enhanced heat transfer in
the entire surface. Lee et al. [29] carried out experiments to investigate heat transfer
enhancement by a perforated plate installed between an impinging jet and the target plate. In
their study, the average Nusselt number was about two times higher in the presence of the
perforated plate than in its absence at the stagnation region. Nakoda et al. [30] experimentally
investigated the effect of the finned surfaces and surfaces with vortex generators on the local
heat transfer coefficient between an impinging circular air jet and a flat plate using thermal IR
imaging on the flat plate. They observed that augmentation of heat transfer for the surfaces
with vortex generators was higher than that for the finned surfaces.

4. ACTIVE CONTROLS IN SINGLE JETS


The flow characteristics of an impinging jet are affected by ejected jet-flow conditions.
This is because vortices are generated at the free shear layer by the initial KelvinHelmholtz
instabilities and the vortex ring grows; moreover, a vortex pairing process occurs as the jet
flow develops. Therefore, the flow and heat transfer characteristics on the impingement
surface can be changed by the controlling the vortices. Several techniques are used to control
the vortices such as adding secondary flows to the jet periphery, changing nozzle exit shapes
and conditions, and acoustic excitation. The acoustic excitation method forces sound waves
into the jet flows, resulting in flow and heat-transfer characteristics being affected by the

44

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

forcing frequencies. Thus, the heat transfer rates can be enhanced or reduced on the
impingement surface by forcing the flows with the proper excitation frequency.
Cho et al. [31] and Lee et al. [32] investigated heat transfer characteristics with additional
secondary flows to the jet periphery for different nozzle types such as circular contoured and
circular pipe nozzles. Crow and Champagne [33] forced the flow with specific frequencies
and amplitudes. They found that the jitters generated at the shear layer decreased when the
flow was excited with a frequency of St = 0.3, and referred to this excitation condition as a
preferred mode. Zaman and Hussain [34-35] found that the turbulence intensity of the flow
with an excitation St of 1.6 or 2.4 is lower than that of the non-excited jet. Gau et al. [36]
showed that flow structure was significantly altered by acoustic excitation with an inherent or
non-inherent frequency. At an inherent frequency, that is, a half or a quarter of the
fundamental frequency, the turbulence intensity increased and heat transfer was enhanced by
acoustic excitation. Liu and Sullivan [37] showed that enhancement and reduction of local
heat transfer in the wall-jet region could be attained by forcing the impinging jet near the
natural and subharmonic frequencies, respectively. Hwang et al. [38] conducted an
experimental study to investigate the flow and heat transfer characteristics of an impinging jet
with acoustic excitation. Two acoustic excitation methods differing in the location of the
actuatormain jet excitation and shear layer excitationwere tested and compared. The
effects of excitation level on the heat-transfer and flow characteristics were also investigated.
Despite differences in the basic flow scheme and location of the forcing actuator, the main jet
excitation method showed heat transfer and flow characteristics similar to those with shear
layer excitation. As shown in Figure 6, the effects of acoustic excitation on the jet increase
with the excitation level. Therefore, both the frequency and level of excitation are important
factors in acoustic excitation.

a)

b)

Figure 6. (a) Schematic of acoustic excitation system. (b) Experimental results. Comparison of Nusselt
number at the stagnation point with increasing dB with the shear layer excitation [38]. (D=nozzle exit
diameter, H=nozzle-to-plate distance, St=Strouhal number,

).

Hwang et al. [39] investigated the flow and heat transfer characteristics of an impinging
jet by comparing the results of controlling vortex pairing, secondary flows, and acoustic
excitation. The resultant flow patterns with secondary flow are interesting for jet-control

Applications of Impingement Jet Cooling Systems

45

design, as shown in Figure 7. For the free jet without secondary flow in Figure 7(a), a ring
vortex, created around the jet by virtue of the instability of the mixing layer, moves
downstream. For the co-blowing flow in Figure 7(b), there are two shear layers between the
main jet and the ambient air because the secondary flow is ejected around the former. One is
the shear layer between the secondary flow and the ambient air; the other is between the main
flow and the secondary flow. The development of the main jet flow is delayed by the
secondary flow. The reason is that the co-rotating outer vortices inhibit vortex pairings of the
inner vortices and the entrainment of ambient air. As a result, the jet potential core extends far
downstream. Although there are two shear layers for the suction case in Figure 7(c) as well,
the characteristics of the vortices are different from those in the other cases. With suction
flow, development of the main jet is enhanced because of the relatively high shear between
the main and secondary flows. The Nusselt number values at the stagnation point for various
spacings are shown in Figure 8. As mentioned above, for cases of secondary flow, the Nusselt
number for small spacings is slightly lower than that of the conventional jet. As the distance
increases, this trend is reversed and higher heat transfer rates are obtained. For suction cases,
the Nusselt numbers hardly differ at small nozzle-to-plate distances. However, they differ to
some extent at larger nozzle-to-plate distances.
(a)

(b)

(c)

Figure 7. Flow visualization results of free jet with the secondary flows: (a) normal jet; (b) co-blowing
jet; (c) suction jet [39].

Figure 8. Effect of secondary flow rate on variation of Nusselt number at the stagnation point with
various spacing: (a) Co-blowing; (b) suction [39]. (D=nozzle exit diameter, H=nozzle to plate distance,
R=velocity ratio).

46

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

5. HIGH VELOCITY JET FLOW AND JET IMPINGEMENT


When a high-pressure gas expands through a nozzle into the atmosphere, the state of a jet
varies depending on the pressure ratio through the nozzle. An under-expanded jet is obtained
if the pressure at the exit plane is greater than the ambient pressure. The typical shock
structure of a highly under-expanded jet is reported by Kim et al. [40] and illustrated in
Figure 9. The expansion waves from the nozzle lip reflect off the constant pressure boundary
as compression waves. Subsequent coalescence of these waves results in the appearance of
shock waves referred to as intercepting or barrel shocks.

Figure 9. Schematic diagram and visualization of under-expanded free jet [40].

For moderate under-expansion ratios, the intercepting shock may reflect regularly at the
centerline (regular reflection, RR), after which the reflected shock is reflected again off the jet
boundary as expansion waves. This process is repeated downstream, establishing the wellknown pattern of shock diamonds. As the under-expansion ratio is increased, a normal shock
known as a Mach disk occurs in the cell (Mach reflection, MR). Behind the Mach disk, a
region of subsonic flow is formed, which is separated from the supersonic flow behind the
reflected shock by the slip line.
When a flat plate is inserted in the near field of a free jet, the flow structure is
significantly disturbed [41]. Figure 10 shows the shock structure in the near field of the
under-expanded impinging jet. A stand-off shock occurs ahead of the plate, and the flow
decelerates through the shock. Owing to the interaction of the stand-off shock with the
intercepting shock in a free jet, a reflected shock is formed. The subsonic region behind the
stand-off shock is separated from the supersonic flow behind the oblique reflected shock by
the slip surface, and a mixing zone develops along the slip surface. As the shock-layer fluid
moves radially outward, the flow re-accelerates and the wall-jet region starts to develop. A
shear layer grows along the constant-pressure upper boundary of this wall jet and a boundary
layer along the surface of the plate. The wall-jet velocity is reduced to a subsonic value by the
gradual growth and merging of these viscous layers.

Applications of Impingement Jet Cooling Systems

Ma<1

1.

Intercepting shock

2.

Constant pressure jet boundary

3.

Mach disk

4.

Reflected shock

5.

Slip line

47

Ma>1

Figure 10. Schematic diagram and visualization of under-expanded impinging jet [41]. (Ma=Mach
number).

When a supersonic jet impinges on a solid surface, dynamic energy is converted to


thermal energy and high temperature is induced on the surface. Therefore, severe aerodynamic and thermal loads can be produced on the impinged surface. Impingement of a
supersonic jet on solid objects can be found in various situations such as the launch of a
rocket, the takeoff and landing of a vertical/short takeoff and landing aircraft, jet-engine
exhaust impingement, the thrust-vector control system of a solid rocket motor, and so on. A
supersonic jet may also be used for testing heat shields such as shuttle tiles. Supersonic jet
impingement includes very complicated phenomena such as a multiple shock system
consisting of barrel, plate, and reflected shocks, as well as the instability of a stagnation
bubble.
Ginzburg et al. [42] investigated the interaction between a supersonic under-expanded jet
and an obstacle mounted at a various distances from the nozzle exit. The jet used by the
authors had an exit Mach number of about 2.2 and an under-expansion ratio of 1.9. Potential
explanations were given for the occurrence of reverse flow pertinent to the peripheral
pressure maximum and unstable modes of flow for a certain nozzle-to-plate distance.
Gubanova et al. [43] studied an under-expanded impinging jet with an exit Mach number of
2. They also confirmed the presence of reverse flow on the plate using total-head tubes
installed parallel to the flow. Gummer and Hunt [44] measured surface pressure distributions
and shock shapes. Four jets were used with Mach numbers in the range of 1.642.77. They
compared the experimental results with theoretical predictions based on the method of
integral relations and polynomial approximation. Donaldson et al. [45] conducted an
extensive experimental study with three types of jets. Among these, a slightly under-expanded
jet (Pe/Pa = 1.42) and a highly under-expanded jet (Pe/Pa = 3.57) were included. They
correlated the radial velocity gradient at the impingement stagnation point with the free jet
centerline velocity and half radius (i.e., the radius at which the velocity drops to one-half its
maximum value). They also reported the existence of a stagnation bubble for the case of a
highly under-expanded jet. Carling and Hunt [46] investigated the wall jet from four
supersonic jets used in previous investigations by Gummer and Hunt [44]. They showed that

48

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

the mechanism, which mainly determines the supersonic near-wall jet, is jet edge expansion
and its reflections from the sonic line and the wall-jet boundaries. The near-wall jet is found
to consist of an alternating series of expansion and re-compression regions whose strengths
depend on the jet Mach number and decay with distance. Kalghatgi and Hunt [47] reported an
experimental investigation on the occurrence of bubbles in the shock layers of supersonic
impinging jets. They ascribed the occurrence of the bubbles to a mechanism different from
that proposed by Ginzburg: the interaction of the plate shock with very weak shock waves
produced in the jet. The weak shock waves in the jet may originate from small imperfections
in the nozzle wall or slight inaccuracies in the design or production of the nozzle contour.
Zien et al. [48] conducted a study of the flow field associated with a two-dimensional
supersonic jet impinging normally on a flat plate. They used two wedge nozzles manufactured
at the design Mach numbers of 1.75 and 2.75 to provide the jet in the experiment. By
comparing the results of theoretical computations based on the method of integral relations
with experiments, they showed that the approximate theory adequately predicted the surface
pressure distribution on the plate surface.
In contrast to numerous investigations related to flow structures, very few data are
currently available on heat transfer characteristics of under-expanded impinging jets. Most
studies of heat transfer between impinging jets and a surface have been conducted for the case
of subsonic jets. Gardon and Cobonpue [49] reported an interesting result related to the
present issue; they observed periodic variation of the heat transfer rate at the stagnation point
with a nozzle-to-plate distance for the case of Re = 112000, which corresponds to an underexpanded supersonic jet. However, as subsonic jets were their main concern, no further
explanation of under-expanded cases was given. Fox et al. [50] introduced a conceptual
model for the separation of total temperature to explain the experimental data taken at high
subsonic Mach numbers, and ascribed the region of lower-wall temperature to 'vortexinduced cooling'the total energy separation by the secondary vortices generated near the
plate surface. In their extended study on energy separation, Fox and Kurosaka [51] also
proposed a mechanism of 'shock-induced cooling' for supersonic jets. They showed that
unsteady movements of shock structure disrupted by the formation and movement of vortices
cause total temperature separation. For the under-expanded case, they used a jet issuing from
a convergent nozzle, and with a convergent-divergent nozzle, they obtained the jet data under
over-expanded conditions at the upper limit allowed by their experimental rig. Although they
mentioned the shock-induced cooling on the impinging plate, they presented only the results
for total temperature distributions in the free jets.
Kim et al. [41] and Yu et al. [52] investigated the flow and heat transfer characteristics of
under-expanded sonic jet impingement on a flat plate as shown in Figure 11. Heat transfer
coefficients at small nozzle-to-plate distances within the initial expansion length increase
because of the turbulence induced by mixing near the slip-line, which originates from the
intersection of the stand-off shock and the intercepting shock of a free jet. As the underexpansion ratio is increased, the increment in heat transfer coefficients persists beyond the
initial expansion length given the marked subsonic core behind the strong Mach disk and the
interaction of the subsonic region with the reflected expansion waves. As a result, the
maximum Nusselt number at the stagnation point for an under-expansion ratio of Pe/Pa = 3.5
is approximately twice that for Pe/Pa = 1.5. As the nozzle-to-plate distance is increased, heat
transfer coefficients increase given the encroachment of the mixing layer into the center of the
jet. Thus, a secondary local maximum Nusselt number occurs where the periodic variation of

Applications of Impingement Jet Cooling Systems

49

the pitot pressure in the jet almost decays. This maximum increases by up to 30% with
increasing under-expansion ratio within the present experimental range.

Figure 11 Stagnation Nusselt number distribution for different under-expansion ratios [52]. (Pe= static
pressure at the nozzle exit plane, Pa= ambient pressure, Zp= axial distance from the nozzle exit plane,
DNE= nozzle exit diameter ).

Kim et al. [40] presented the flow and heat transfer characteristics for the impingement of
under-expanded, axisymmetric, and supersonic jets on a flat plate. Figure 12 shows the
recovery factor distributions as a function of nozzle-to-plate distance. The results show that
the interaction of shock waves in the free jet with a plate shock formed in front of the
impingement plate depends strongly on the distance of the plate from the nozzle exit.
Consequently, the surface pressure and adiabatic wall-temperature distributions on the
impinging plate change significantly with the nozzle-to-plate distance. The temperature
recovery on the plate, where a supersonic jet impinges, is characterized by 1) the consistent
existence of a cooled region near the jet periphery and 2) a central cooled region that appears
occasionally. The former is due to the vortex-induced total temperature separation, and the
latter is mainly attributable to the vortex-shock interaction. However, the surface pressure
distribution, which has a peripheral maximum, implies that the cooling in the central region is
linked to the existence of a stagnation bubble: the cooling effect in the central region may be
enhanced by the stagnation bubble, which traps the fluids with a lower total temperature.
When a jet impinges on a plate at an angle of inclination, the jet has three-dimensional
rather than axisymmetric characteristics, and the surface phenomena on the target plate
become more complex than those for perpendicular impingement. Kim et al. [40] also
investigated the flow and heat transfer characteristics of inclined impinging of underexpanded jets on a flat plate. Yu et al. [52] performed a similar study using an IR camera. For
this reason, Yu et al. [52] presented two-dimensional distributions that provide better spatial
resolution of data.
Along with these under-expanded jet-impingement studies, Shin et al. [53] and Kim et al.
[54] have conducted heat transfer studies on perpendicular jet impingement with various
subsonic jet flows using IR thermography methods, as shown in Figure 13.

50

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song


Recovery Factor = 1

Tr Tjo
u2j / 2C P

1.2
1.5

2.0

2.5

1.0
0.5
1.0
2.5
0.8

2.0
1.5

-6

-4

-2

1.0
0

R/DNE

ZP/DNE

0.5

exit Mach number Ma=1.8


exit pressure ratio Pe/Pa=1.15

nozzle pressure ratio P0/Pa=1.15


Figure 12. Recovery factor distribution of Supersonic jet impingement for Ma=1.8 and Pe/Pa=1.15 [40].
(DNE= nozzle exit diameter, R=radial distance, Zp= axial distance from the nozzle exit plane).

Figure 13. Recovery factor obtained by IR-camera and thermocouple for Ma=0.54 [53]. (D= nozzle exit
diameter, R=radial distance, L=lateral distance).

6. LIQUID JET IMPINGEMENT


The liquid impinging jet has been regarded as a solution to critical cooling requirements,
particularly in cases where very high rates of heat removal are necessary. Liquid-jet
impingement-cooling is employed in numerous industrial applications such as annealing of
hot steal sheets and cooling of electronic devices. The limited range of cooling capacity

Applications of Impingement Jet Cooling Systems

51

afforded by air jet impingement is extended by employing liquid jets that, through convective
nucleate boiling, yield higher cooling performance. Jet impingement enhances the convective
heat transfer coefficient in the nucleate boiling regime and increases the CHF, or maximum
heat flux, thereby extending the wall superheat range of stable operating conditions.
Both air and liquid jet impingements have been applied to electronics cooling
applications. In 1983, IBM shipped its first system using array jets that cooled the 64 mm by
64 mm multi-chip module (containing 36 chips) used in the IBM 4381 processor with a direct
thermal path from the back of each chip to the ceramic cap via thermal paste. This cooling
scheme [55-56] employed what is known as highly parallel impingement. A liquid array-jet
impingement cooling system using fluorocarbons [57] was utilized in the SSI supercomputer
SS-1 to dissipate 40 W on 6.5 mm chips.
Confinement is inevitable in many applications of impinging jets because of spatial
constraints or installation limitations. Jet confinement in the stagnation and/or wall-jet
region(s) may alter the flow field and heat transfer to be different from those of the
unconfined jet, as reviewed by Garimella [58]. Such is also the case when heat transfer
augmentation techniques are imposed on the target surface, thus necessitating a thermal
analysis that considers the degree of confinement. The confinement of liquid jets arises in
even more diverse situations than that of air jets. Moreover, jet-impingement boiling can
seriously deteriorate with confinement because of the considerable volumetric expansion by
phase changes that increase the size of the pressure drop and can decrease the maximum heat
flux.
Despite the importance of considering confinement in the design of liquid-jet
impingement-cooling systems, there has been little study of confined liquid impinging jets.
The partial boiling regime is another topic that has received less attention than its usefulness
would dictate. This regime has been regarded as a transition stage between single-phase
convection and fully developed nucleate boiling. However, the high convection of jet
impingement commonly extends the heat flux range where partial boiling is available. Partial
boiling of a water jet may cover a heat flux range of 30300 W/cm2, which corresponds to the
heat generation densities of many industrial applications where jet-impingement boiling can
be applied as an effective cooling method. Therefore, the effects of confinement on jetimpingement heat transfer are carefully investigated over the various heat transfer regimes of
single-phase convection, partial boiling, and fully developed nucleate boiling.
Many studies on heat transfer from liquid-jet impingement have been conducted for
single-phase convection, nucleate boiling, and CHF. The high transport of jet impingement
substantially extends the heat flux range corresponding to the partial-boiling regime, thus
covering the diversity of the heat removal rates required in industrial applications. Although
there is great potential for applying partial boiling to cooling applications, few investigations
on the matter have been reported in the literature. Vader, Wolf and their coworkers [59-61]
have been conducted and reported comprehensive investigations of partial boiling.
Conversely, the onset of nucleate boiling has attracted the attention of some researchers [6265], and fully developed nucleate boiling and CHF have been extensively investigated.
Local measurements are essential to understanding the relationship between jetimpingement heat transfer and other parameters. For single-phase convection, local
measurements of heat transfer are commonly employed as module-averaged measurements.
However, it can be difficult to install as many sensors as are required to obtain fine-resolution
data when high-heat flux is involved. Heater material in the form of shim stock or thin foil,

52

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

which is commonly used in local measurements, is too delicate to apply to the high-heat flux
conditions associated with nucleate boiling let alone CHF. Thus, very few of the previous
investigations on jet-impingement boiling have included local measurements [59-61].
In applying liquid jets, the confined jet is very useful. Wu et al. [65] compared the
convection coefficient distributions of confined and unconfined single-circular jets as
illustrated in Figure 14. Confining the single-circular jets forces the radial flows to be twoway channel flows, enhances mixing and turbulence, and increases the effective flow rate
over the heated surface. The precipitated transition to turbulence elevated convection
coefficient distributions for the confined single jet at x/d<6 with V = 9.5 m/s and at x/d<4
with V = 3.3 m/s, indicating that as the velocity increased, the region where the confined jet
outperformed the unconfined jet expanded. Enhancement of the stagnation convection
coefficient also increased with increasing velocity, which is attributed to the increase in
mixing and turbulence associated with increasing velocity.
a)

b)

Unconfined

Figure 14. Continued on next page.

Applications of Impingement Jet Cooling Systems

53

c)

Unconfined

Figure 14. (a) Schematic of confined and unconfined slot jet. (b) V=3.3 m/s. (c) V=9.5 m/s. Effects of
confinement on convection heat transfer coefficient distributions for jet impingement at z=0.0 [65]. (d=
nozzle exit diameter, H=nozzle-to-plate distance, x=streamwise direction, z=lateral direction, h=heat
transfer coefficient).

In studies on liquid jets, the effects of working fluid are intensely studied issues.
Nakayama et al. [66] investigated boiling heat transfer from forced convection in a planar
impinging flow using fluorocarbon FX3250 (PF5050). They simulated a chip array with five
(4 mm wide) strip heaters made of constantan foil with a thickness of 10 m. They observed
that the transition from single-phase heat transfer to partial boiling was accompanied by a
temperature overshoot from the center strip. The observed transitions and CHF were
dependent on the jet velocity and subcooling. In their experiments, the local heat transfer was
only measured at a stagnation point because only one thermocouple was installed on the
center strip with a width of 4 mm (x/W = 2). They did not investigate measurements recorded
in the lateral direction in the stagnation region, x/W~2, in which the heat transfer rate varied
greatly. Inoue et al. [67] investigated the CHFs of pure water in the confined flow of a twodimensional jet on flat and concave surfaces under various flow conditions and measured the
CHF at different downstream positions. They found the local CHF profile of the confined jet
flow cooling system on the surface curvature heater. To evaluate the transition from local jetimpingement heat transfer to boiling heat transfer for a free-surface, a planar jet of water with
a thin-plate heater (thickness = 0.2970.635 mm) was also employed by Wolf et al. [60-61,
68] and Vader et al. [69].
Shin et al. [70] investigated the effects of nozzle-to-plate spacing in a confined liquid jet
using fluorocarbon PF5050. They showed that, at the nozzle-to-surface spacing of H/W = 4.0,
the fully developed boiling regime is initiated at a heat flux of q = 14.8 W/cm2. As shown in

54

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

Figure 15(a), the vapor bubble on the heated surface at a heat flux of q = 16.0 W/cm2 is
distributed across the entire surface including the stagnation point. However, at a spacing of
H/W = 1.0, as shown in Figure 15(b), the bubble at the stagnation point appears intermittently,
where the bubble is absent partially at the stagnation zone despite a fully developed boiling
regime for a high heat flux of q = 22.0 W/cm2. It is noted that the critical heat flux is 29.1
W/cm2 in this case. The vapor bubble near the stagnation point is eliminated quickly by
liquid-jet impingement, although the confined channel is occupied fully by bubbles at the
downstream region.
a)

b)

Figure 15. (a) Re=2,847, H/W=4.0, q=16.0 W/cm2. (b) Re=2,995, H/W=1.0, q=22.0 W/cm2.
Photographs in the fully developed boiling regime at Re 3,000 for H/W=1.0 and 4.0 [70].

The critical heat flux data obtained in the experiments with a heated length of 10 mm are
plotted for various nozzle-to-surface spacings in Figure 16. A critical heat flux correlation
induced from the experimental data is expressed to second-order function of impinging jet
height-to-width and mass velocity.
The critical heat flux increases as a function of the mass flow rate (jet velocity), and that
is consistent with typical trends for CHF. In the present experiments, the critical heat flux at a
spacing of H/W = 1.0 is lower than that at a spacing of H/W = 0.5 or 4.0. This trend is
observed in tested whole-mass flux conditions. Considerable heat transfer enhancement of jet
impingements with low spacing resulted from significant increases of the turbulence intensity
by acceleration of the impinged flow and the velocity discrepancy between the accumulated

Applications of Impingement Jet Cooling Systems

55

and impinged flows. The mechanism results in high heat transfer in a wide wall-jet region.
This is called a secondary peak, a well-known heat transfer characteristic induced by an
impinging jet with low spacing. However, as the nozzle-plate spacing increases, the heat
transfer rate of the impinging jet is reduced, and the high heat transfer region becomes
narrow. Thus, at large nozzle-plate spacing, the heat transfer rate is related to the mass flow
rate. Given that the CHF is also enhanced by the heat transfer rate, a high CHF at H/W = 0.5
is observed with the high heat transfer rate of impinging jet flow, whereas a low CHF at H/W
= 1.0 appears because of the reduced flow velocity near the surface and resulting the low heat
transfer rate. However, the CHF increases at a larger nozzle-plate spacing of H/W = 4,
especially at a high velocity of 0.52 m/s (Re = 5000). The high flow velocity generates the
high heat transfer, and the large spacing helps to remove the generated bubbles from the
boiling surface for a confined channel, so that CHF increases at a high flow velocity and large
nozzle-plate spacing, culminating in a delayed starting point for film boiling. This effect is
enhanced largely with increasing mass flow rate. Therefore, there is a nozzle-to-surface
spacing with a minimum CHF for a certain mass flow rate (or jet velocity). The reason is that
the boiling heat transfer at the stagnation region increases with decreasing spacing because of
the high jet-impingement effect, and conversely the boiling heat transfer on the downstream
region increases with increasing spacing because of sufficient spacing for removing bubbles.
In an actual design, the geometry of the impinging jet avoids the H/W with the lowest critical
heat flux because it causes low system performance. Therefore, the optimum nozzle-plate
spacing should be selected for enhancing performance. Thus, they have found the impinging
jet nozzle-to-surface spacing (H/W) with the lowest CHF at each flow velocity (V).
45.0
Re
2000
3000
5000

Exp. Data

Correlation

q"CHF [W/cm ]

40.0

V [kg/m
[m/s]2s]
G
0.21
336
0.30
493
0.52
842

35.0
The lowest q"CHF

30.0

25.0
0.0

1.0

2.0

3.0

4.0

H/W

Figure 16. Effect of nozzle-to-surface spacing and jet velocity on critical heat flux [70]. (H= nozzle-tosurface distance, W=nozzle width,

=critical heat flux).


qCHF

7. ARRAY JET IMPINGEMENT


Since multiple jets cover heating or cooling in large areas, the key factors are both a high
average heat transfer coefficient and uniformity of heat transfer over the target surface.
Therefore, studies on multiple jets have been mainly focused on identifying the optimum

56

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

multiple jet condition by considering interactions among the jets. Recently, the majority of
these studies are related to complicated effects of applied conditions (crossflow, rotation,
geometry, etc.), or combined techniques (rib turbulator, pin fin, dimple, effusion hole, etc.)
instead of simple multiple jets.
San and Lai [71] investigated the effects of jet-to-jet spacing on the local Nusselt number
for confined circular air jets. They suggested the optimum jet-to-jet spacing for Re in the
range of 10,00030,000 from their experimental results. Brevet et al. [72] studied
optimization in the case that impingement was confined by the test section and spent air was
constrained to exit in only one direction. The heated-thin-foil technique was used jointly with
IR thermography. Ekkad and Kontrovitz [73] obtained detailed heat transfer distributions on a
jet-impingement target surface with dimples. Their results showed that the presence of
dimples on the target surface, in-line or staggered with respect to jet location, produced lower
heat transfer coefficients than in their absence. They reported that bursting phenomena
associated with flow over dimples produces disturbances of the impingement jet structures
and results in lower levels of heat transfer coefficients on the target surface. Kanokjaruvijit
and Martinez-botas [74] performed experiments on an array-jet impinging onto a staggered
array of dimples at a Reynolds number of 11,500. Two dimple geometries, hemispherical and
cusped elliptical, were examined. Moreover, the effect of crossflow on heat transfer was
investigated using one-, two-, and four-way spent air exits. Nakabe et al. [75] studied the
interactions of longitudinal vortices made by two inclined impinging jets in in-line and
staggered arrays using thermochromic liquid crystal, fluorescence dyes, and particle image
velocimetry (PIV). Taslim et al. [76] measured the heat transfer coefficient of impingement
for different target wall-roughness geometries on a curved surface: a smooth wall, a wall with
high surface roughness, a wall roughened with conical bumps, and a wall roughened with
tapered radial ribs. They showed the averaged Nusselt number at various operating conditions
using thermocouples embedded in the brass test piece. They reported that although surface
roughness increases the impingement heat transfer coefficient, the driving factor in heat
transfer enhancement is the increase in surface area. Haiping et al. [77] studied heat transfer
on a rib-roughened surface with arrays of circular jets and suggested correlation equations.
Son et al. [78] presented detailed heat transfer coefficient distributions measured on
downstream surface and target-surface of the impingement plate. It was found that the
average heat transfer coefficient on the impingement downstream surface is about 50% of the
average target-surface heat transfer coefficient. Gao et al. [79] tested impingement heat
transfer for linearly stretched arrays of holes. Two different arrays were investigated, with
holes of uniform diameter through the array for one case and holes of varying diameter at
every row location for another. Rhee et al. [80] investigated the effects of spent air flows with
and without effusion holes on the impingement plate. The results showed that for a multiple
jet without effusion holes, the spent air of the injected jets forms a crossflow within the
confined space and significantly affects the downstream jet flow. However, uniform
distributions and enhancements of heat transfer coefficients were obtained by installing the
effusion holes. The understanding of heat transfer in a cooling system with crossflow is
necessary because the crossflow exists in the internal passages in actual combustor-wall or
turbine-blade cooling operations. The studies related to the crossflow effect are mainly
concerned with the array-jet impingement on the solid plate. Huber and Viskanta [81]
examined the effects of spent air exit in the orifice plate on local and average heat transfer for
an array of impinging jets. Their results show that the interaction of adjacent impinged jets is

Applications of Impingement Jet Cooling Systems

57

reduced by the spent air, which results in enhanced heat transfer on the target plate. Huang et
al. [82], Metzger and Korstad [83], and Florschuetz et al. [84] investigated the effects of
crossflow on heat transfer/flow characteristics of array-jet impingement, and reported that
heat transfer is reduced at upstream regions because of initial crossflow. Bailey and Bunker
[85] and Gao et al. [79] investigated heat transfer characteristics for impinging jet arrays with
crossflow. Andrews et al. [86] investigated the effects of the direction of transverse ribs with
respect to the crossflow for impingement cooling, and reported that the rib turbulator changes
the effects of crossflow and enhances heat transfer, especially with ribs normal to the
crossflow. Yoon et al. [87] showed detailed flow and heat transfer information using
numerical simulations and the naphthalene sublimation method. As shown in Figure 17, they
reported that the local maximum heat transfer coefficients move further in the downstream
direction due to the increase of crossflow velocity. In addition, the effect of the crossflow
occurs strongly at the small gap.

Figure 17. Heat transfer coefficients in array jet cooling with crossflow [87]. (d= nozzle exit diameter,
H=nozzle-to-plate distance, x=axial direction, y=lateral direction).

58

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

Compared to stationary impingement cooling, studies on rotating impingement cooling


have been shown in a few researches. Epstein et al. [88] investigated the heat transfer
characteristics in the leading edge of the turbine blade at the first time. They found that
rotation reduces heat transfer and changes local heat transfer distributions under high rotation
conditions. Analytic analysis was also performed; it was inferred that the observed heat
transfer reduction is due to a buoyancy effect. Aronstein [89] studied the rotating impinging
jet flow using flow visualization in a rotating water tank. It was shown that the main jet flow
and vortex structure change depending on the jet orientation relative to the rotating axis.
Mattern et al. [90] measured local heat/mass transfer coefficients on a curved impinging plate
with regard to various parameters such as hole-to-hole spacing, hole-to-plate spacing (H/d),
and jet orientation. They reported that the spanwise heat/mass transfer coefficient decreases at
high H/d, although the heat/mass transfer is slightly higher than that of the stationary case at a
small H/d. Glezer et al. [91] examined swirling flow using rotation focused on the leading
edge region of a turbine blade. They showed that Coriolis forces play an important role in
enhancing the internal heat transfer when their direction coincides with a tangential velocity
vector of swirling flow. Hsieh et al. [92] measured local heat transfer coefficients for a single
confined impinging jet using a thermocouple at various Reynolds and Rotation numbers. The
combined jet impingement and rotation effect were shown to affect the heat transfer response.
Yan et al. [93] obtained detailed heat transfer characteristics under elliptic multiple jets by a
transient liquid crystal technique. Elliptic jet holes of five different aspect ratios, AR = 4, 2, 1,
0.5, and 0.25, jet Reynolds numbers Re = 1500, 3000, and 4500, and three exit flow
conditions were considered. The results showed that the effects of aspect ratio and crossflow
had significant influences on the axial shift of the impingement/touchdown locations.
Hwang and Chang [94] studied the heat-transfer and pressure-drop characteristics in a
triangular duct cooled by an array of jets. The outflow orientation significantly affects the
local heat transfer characteristics by influencing the jet flow along with the crossflow in the
triangular duct. The triangular duct with two openings was recommended given that it had the
highest wall-averaged heat transfer and a moderate loss coefficient among the three outflow
orientations investigated. Iacovides et al. [95] reported an experimental study of impingement
cooling in a rotating passage with a semi-cylindrical cross section. They reported that very
strong rotation leads to the disappearance of all secondary peaks and some of the primary
peaks in array jets, and that the heat transfer distributions are caused by this flow behavior.
Hong et al. [96] investigated the heat transfer characteristics on a concave surface for
impinging jets in stationary and rotating domains and compared the results on a concave
surface with those on a flat surface as shown in Figure 18. They concluded that the heat
transfer on the concave surface is enhanced by increasing the spanwise direction because of
the curvature effect.

Figure 18. Local heat transfer distributions on the concave surface in array jet cooling [96].

Applications of Impingement Jet Cooling Systems

59

8. ARRAY JET IMPINGEMENT ON EFFUSION SURFACE


The previous studies are mainly concerned with impinging jets on a solid plate without
effusion holes on a target plate. Downs and James [97], Jambunathan et al. [11], Viskanta [2],
and Cho [98] reviewed the previous studies of jet impinging heat transfer. However, as shown
in Figure 19, studies on heat transfer characteristics for effusion cooling and
impingement/effusion cooling are quite limited. A hybrid cooling scheme uses the combined
mechanisms of impingement, transpiration, and film cooling. The spent impingement fluid is
vented through passages in the target surface into the hot gas or the outside.

Figure 19. Schematic diagram of impingement/effusion cooling [98].

Hollworth and Dagan [99] and Hollworth et al. [100] measured average and local heat
transfer coefficients of arrays of turbulent air jets impinging on perforated target surfaces, and
reported that arrays with staggered effusion holes consistently yielded higher heat transfer
rates than did impinging jets on solid plates. Nazari and Andrews [101] studied film-cooling
performance by examining the effects of number of holes for impingement/effusion cooling.
Cho and Goldstein [102] investigated the effect of hole arrangements on local heat/masstransfer characteristics inside the effusion plate. They found that the high transfer rate is
induced by strong secondary vortices and flow acceleration, and that the overall transfer rate
is approximately 45~55% higher than that with impingement cooling alone. Cho and Rhee
[103], Cho et al. [104], and Ekkad et al. [105] also investigated the heat/mass transfer and
flow characteristics of an impingement/effusion cooling system under various experimental
conditions such as gap distance, Reynolds number, hole arrangement, and size as shown in
Figure 20.

Figure 20. Heat/mass transfer coefficients on the effusion plate at different H/d [103].

60

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

For the impingement/effusion cooling system with crossflow, Rhee et al. [106]
investigated heat transfer characteristics for impinging jet arrays with initial cross flow. They
reported that the overall heat/mass transfer rates on the effusion (target) plate decreased as the
crossflow rate increased, showing a trend similar to that of array-jet impingement. In addition,
they found that locally low transfer regions are formed between the adjacent effusion holes.
Hong and Cho [107] and Hong et al. [108-109] conducted experimental investigations to
enhance the heat transfer for an impingement/effusion cooling system when the initial
crossflow was formed. To improve heat transfer, circular guides or turbulators are installed on
the injection holes, or inclined jets are applied to the impinging plates as shown in Figure 20.

Figure 21. Various Configurations of impinging surfaces for heat transfer enhancement. [107-109].

Applications of Impingement Jet Cooling Systems

61

To enhance the cooling performances in the jet impingement/effusion cooling systems,


the combined cooling schemes have been investigated. Funazaki et al. [110] and Yamawaki et
al. [111] investigated the effects of circular pin fins on impingement/effusion cooling but did
not consider crossflow. Rhee et al. [112] applied various rib configurations to
impingement/effusion cooling systems with initial crossflow. They reported that averaged
heat transfer increased 4~15% compared to cases without ribs (Figure 22). The increase of
heat transfer on the effusion plate with ribs decreases as the crossflow effect becomes strong.
There are only a few studies on combined cooling for impingement/effusion cooling to reduce
the adverse crossflow effect and to enhance cooling performance.

Figure 22. Heat/mass transfer coefficients in impingement/effusion cooling with rib turbulators [112].

Parsons et al. [113-114] investigated the averaged Nusselt number on array-jet cooling
with/without effusion holes in rotation conditions applied to the midcore of turbine blade. At
the leading and trailing jet orientations, the difference in heat transfer is attributed to
crossflow behavior induced by Coriolis forces. It is reported that as the wall-to-coolant
temperature difference ratio increases, the Nusselt number ratios decrease up to 10% for all

62

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

walls in both channels with all other parameters held constant. However, they performed heat
transfer experiments using thermocouples embedded in the copper block and suggested only
the bulked Nusselt number data. Therefore, it is not possible to accurately predict the detailed
heat transfer distributions on the rotating blade with effusion holes. Detailed information
about the heat transfer coefficient is required for the prevention of hot spots as well as a better
design of the rotating blade. Hence, it is required to obtain the local heat transfer coefficients
on the effusion surface for rotating impingement/effusion cooling system. Hong et al. [115117] measured local heat transfer coefficients using naphthalene sublimation method and
investigated the local heat transfer characteristics for rotating impingement/effusion cooling,
and devised design parameters such as jet orientation and surface geometry as shown in
Figure 23. They showed that the heat/mass transfer distributions in the axial orientation were
similar to those for the stationary cases, while the trailing orientation produced different heat
transfer distributions with divided high heat transfer regions and a low heat transfer region
around the stagnation area. They also showed that the concave surface provided better and
more uniform heat transfer than the flat surface.

Figure 23. Local heat transfer coefficients in rotating impinge/effusion cooling: (a) stationary; (b)
rotation-axial orientation; (c) rotation-trailing orientation [117].

Applications of Impingement Jet Cooling Systems

63

9. CONCLUSION
In this book chapter, applications of impingement jet cooling systems are discussed.
Cooling and heating under an impinging jet is generally superior to that achieved with typical
convective heat transfer methods. With an impinging jet, it is easy to adjust the location of
interest and to remove a large amount of heat on the impingement surface. For these reasons,
the impinging jet technique has been widely used in many industrial systems such as gas
turbine cooling, rocket launcher cooling and high-density electrical equipment cooling.
Recently, impinging cooling schemes have been applied to improve the performance in micro
and nano systems.
For improving cooling performance, researchers have performed numerous experiments
to control and design single and array jets. They considered the nozzle geometry, impinging
surface, and active methods such as jet vibration, secondary injection, and suction flow. In
addition, various factors concerned with operating conditions (crossflow, rotation, mass flow
rate, working fluid, etc.) and combined techniques (rib turbulator, pin fin, dimple, effusion
hole, etc.) have been investigated in order to identify the optimum conditions and geometries.
There have been numerous experimental and numerical investigations on measurements of
flow and heat transfer characteristics on impinging jets. The liquid crystal method, infrared
(IR) camera technique, and naphthalene sublimation method have been used experimentally
to obtain detailed heat transfer distributions. Given the improvements in computational
performance, many numerical calculations have been carried out to obtain local heat transfer
distributions in complex geometries. Various turbulence models have been developed for use
in many applications, because the numerical results are still disagreement with the
experimental results in their local distribution.
This book chapter is discussed and summarized as following topics:
(a) Single jet impingement: Active controls such as acoustic excitation and jet with
secondary flow for controls of flow pattern and heat transfer
(b) Supersonic jet flow and jet impingement: Flow and heat transfer characteristics by
compressible flow jet flows and jet impingement
(c) Liquid impinging jet: Single-phase and boiling heat transfer characteristics in jet
impingement
(d) Array jet impingement: Optimal design of multiple jet under various operating
conditions
(e) Array jet impingements on effusion surface: Flow and heat transfer characteristics by
additional film cooling effect and deflation of the hot air impinged

REFERENCES
[1]
[2]
[3]
[4]

Martin, H. Advances in heat transfer 1977, 13, 1-60.


Viskanta, R. Experimental Thermal and Fluid Science 1993, 6, 111-134.
Han, B.; Goldstein, R. J. Annals New York Academy of Sciences 2001, 934, 147-161.
Weigand, B.; Spring, S. Int Symp on Heat Transfer in Gas Turbine Systems, Antalya,
Turkey, 2009.

64
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song


Goldstein, R. J.; Behbahani, A. I.; Heppelmann, K. K. Int. J. Heat Mass. Transfer 1986,
29, 1227-1235.
Goldstein, R. J.; Sobolik, K. A.; Seol, W. S. J. Heat Transfer 1990, 112, 608-611.
Hwang, S. D. Effects of Nozzle Exit Condition and Shape on Flow and Heat Transfer
Characteristics of Impinging Jet (Master Thesis), Yonsei University, Korea, 2000.
Lee, J.; Lee, S. J. Proceedings of 11th IHTC 1998, 5, 433-438.
Gardon, R.; Akfirat, J. C. Int. J. Heat Mass Transfer 1965, 8, 1261-1272.
Popiel, C. O.; Trass, O. Experimental Thermal and Fluid Science 1991, 4, 253-264.
Jambunathan, K.; Lai, E.; Moss, M. A.; Button, B. L. Int. J. Heat Fluid Flow 1992, 13,
106-115.
Chung, Y. M.; Luo, K. H. J. Heat Transfer 2002, 124, 1039-1048.
ODonovan, T. S.; Murray, D. B. Int. J. Heat Mass Transfer 2007, 50, 3291-3301.
Hwang, S. D.; Cho, H. H. Transaction of KSME (B) 2004, 28, 135-145.
Bilen, K.; Bakirci, K.; Tapici, S.; Yavuz, T. Int. J. Energy Research 2002, 26, 305-320.
Lee, J.; Lee, S.J. Int. J. Heat Mass Transfer 2000, 43, 555-575.
Zhou, D.W.; Lee, S.J. Int. J. Heat Mass Transfer 2004, 47, 2097-2108.
Gao, N.; Sun, H.; Ewing, D. Int. J. Heat Mass Transfer 2003, 46, 2557-2569.
Yan, X.; Saniei, N. Int. J. Heat Fluid Flow 1997, 18, 591-599.
Beitelmal, A.H.; Saad, M.A.; Patel, C. D. Int. J. Heat Fluid Flow 2000, 21, 156-163.
Nakabe, K.; Inaoka, K.; Al, T.; Suzuki, K. Energy Convers. Mgmt. 1997, 38, 11451153.
Cornaro, C.; Fleischer, A.S.; Goldstein, R.J. Experimental Thermal and Fluid Science
1999, 20, 66-78.
Chan, T.L.; Leung, C.W.; Jambunathan, K.; Ashforth-Frost, S.; Zhou, Y.; Liu, M.H.
Int. J. Heat Mass Transfer 2002, 45, 993-1006.
Choi, M.S.; Yoo, H.S.; Yang, G.; Lee, J.S.; Sohn, D.K. Int. J. Heat Mass Transfer
2000, 43, 1811-1822.
Lim, K. B.; Lee, C. H.; Sung, N. W.; Lee, S. H. Exp. Therm. Fluid Sci. 2007, 31, 711719.
Lee, C. H.; Lim, K. B.; Lee, S. H.; Yoon, Y. J.; Sung, N. W. Exp. Therm. Fluid Sci.
2007, 31, 559-565.
Gau, C.; Lee, I.C. Int. J. Heat Mass Transfer 2000, 43, 4405-4418.
Pyo, C.K.; Park, S.R.; Kim, D.C.; Kum, S. M.; Yim, J.S. Korean J. Air-Cond. Refr.
Eng. 2000, 12, 277-283.
Lee, D. H.; Lee, Y. M.; Kim,Y. T.; Won, S. Y.; Chung, Y. S. Int. J. Heat Mass Transfer
2002, 45, 213-217.
Nakoda, P. M.; Prabhu, S. V.; Vedula, R.P. Experimental Thermal and Fluid Science
2008, 32, 1168-1187.
Cho, H.H.; Lee, C.H.; Kim, Y.S. ASME Paper No. 98-GT-276, 1998.
Lee, C.H.; Kim, Y.S.; Cho, H.H. Trans. KSME (B) 1998, 22, 386-398.
Crow, S.C.; Champagne, F.H. J. Fluid Mech. 1971, 48, 547-591.
Zaman, K.B.M.; Hussain, A.K.M.F. J. Fluid Mech. 1980, 101, 449-491.
Zaman, K.B.M., Hussain, A.K.M.F. J. Fluid Mech. 1980, 101, 493-544.
Gau, C.; Sheu, W.Y.; Shen, C.H. J. Heat Transfer 1997, 119, 810-817.
Liu, T.; Sullivan, J.P. Int. J. Heat Mass Transfer 1996, 39, 3695-3706.
Hwang, S. D.; Cho, H. H. Int. J. Heat Fluid Flow 2003, 24, 199-209.

Applications of Impingement Jet Cooling Systems

65

[39] Hwang, S.D.; Lee, C.H.; Cho, H.H. Int. J. Heat Fluid Flow 2001, 22, 293-300.
[40] Kim, B. G.; Yu, M. S.; Cho Y. I.; Cho, H. H. J. Thermophysics Heat Transfer 2002, 16,
425-431.
[41] Kim, B. G.; Yu, M. S.; Cho, H. H. J. Thermophysics Heat Transfer 2003, 17, 313-319.
[42] Ginzburg, I. P.; Semiletenko, B. G.; Terpigorev, V. S.; Uskov, V. N. J. Eng. Physics
1973, 19, 1081-1084.
[43] Gubanova, O. I.; Lucev, V. V.; Plastina, L. N. Fluid Dynamics 1973, 6, 298-301.
[44] Gummer, J. H.; Hunt, B. L. The Aeronautical Quarterly 1971, 22, 403-420.
[45] Donaldson, C. D.; Snedeker, R. S. J. Fluid Mech. 1971, 45, 281-319.
[46] Carling J. C.; Hunt, B. L. J. Fluid Mechanics 1974, 66, 159-176.
[47] Kalghatgi, G. T.; Hunt, B. L. The Aeronautical Quarterly 1976, 27, 169-185.
[48] Zien, T. F.; Chien, K.Y.; Driftmyer, R. T. AIAA Journal 1979, 17, 4-5.
[49] Gardon, R.; Cobonpue, J. International Developments in Heat Transfer1962, 454-460.
[50] Fox, M. D.; Kurosaka, M.; Hedges, L.; Hirano, K. J. Fluid Mech. 1993, 255, 447-472.
[51] Fox, M. D.; Kurosaka, M. J. Fluid Mech. 1996, 308, 363-379.
[52] Yu, M. S.; Kim, B. G.; Cho, H. H. J. Thermophysics and Heat Transfer 2005, 19, 448454.
[53] Shin, S.; Kim, B. S.; Lee, J. W.; Yu, M. S.; Cho, H. H. Proceedings of The Third
National Congress on Fluids Engineering 2004, 623-626
[54] Kim, B. S.; Shin S.; Yu M. S.; Cho H. H.; Lee J. W.; Bae J. C. Transaction of KSAS
2007, 11, 47-53
[55] Chu, R. C. Proceedings of the 5th International Conference on Solid-State and
Integrated Circuit Technology 1998, 559-562.
[56] Chu, R. C. Proceedings of the IEEE 15th Annual Semiconductor Thermal Measurement
and Management Symposium 1999, 151-165.
[57] Ing, P.; Sperry, C.; Philstrom, P.; Claybaker, P.; Webster, J.; Cree, R. Proceedings of
IEEE 43rd Electronic Components and Technology Conference 1993, 218-237.
[58] Garimella, S. V. Annual Review of Heat Transfer 2000, 11, 413-494.
[59] Vader, D. T.; Incropera, F. P.; Viskanta, R.. J. Heat Transfer 1992, 114, 152-160.
[60] Wolf, D. H. Turbulent development in a free-surface jet and impingement boiling heat
transfer (Ph.D. Thesis), Purdue University, West Lafayette, IN, USA, 1993.
[61] Wolf, D. H.; Incropera, F. P.; Viskanta, R. Int. J. Heat Mass Transfer 1996, 39, 13951406.
[62] Ma, C. F.; Bergles, A. E. J. Heat Transfer 1986, 29, 1095-1101.
[63] Miyasaka, Y.; Inada, S. J. Chemical Engineering of Japan 1980, 13, 22-28.
[64] Mudawar, I.; Wadsworth, D. C. Int. J. Heat and Mass Transfer 1991, 34, 1465-1479.
[65] Wu, S.-J.; Shin, C. H.; Kim K.M.; Cho, H. H. Int. J. Multiphase Flow 2007, 33, 12711283.
[66] Nakayama, W.; Behnia, M.; Mishima, H. J. Electronic Packaging 2000, 122, 132-137.
[67] Inoue, A.; Ui, A.; Yamazaki, Y.; Lee, S. Nuclear Engineering and Design 2000, 200,
317-329.
[68] Wolf, D.H.; Viskanta, R.; Incropera, F.P. J. Heat Transfer 1990, 112, 899-905.
[69] Vader, D.T.; Incropera, F.P.; Viskanta, R. Experimental Thermal Fluid Science 1991, 4,
1-11.
[70] Shin, C.H.; Kim, K.M.; Lim, S.H.; Cho, H.H. Int. J. Heat and Mass Transfer 2009, 52,
5293-5301.

66

Hyung Hee Cho, Kyung Min Kim and Jiwoon Song

[71] San, J.Y.; Lai, M.D. Int. J. Heat Mass Transfer 2001, 44, 3997- 4007.
[72] Brevet, P.; Dejeu, C.; Dorignac, E.; Jolly, M.; Vullierme, J.J. Int. J. Heat Mass Transfer
2002, 45, 4191-4200.
[73] Ekkad, S.V.; Kontrovitz, D. Int. J. Heat Fluid Flow 2002, 23, 22-28.
[74] Kanokjaruvijit, K.; Martinez-botas, R.F. Int. J. Heat Mass Transfer 2005, 48, 161-170.
[75] Nakabe, K.; Fornalik, E.; Eschenbacher, J.F.; Yamamoto, Y.; Ohta, T.; Suzuki, K. Int.
J. Heat Fluid Flow 2001, 22, 287-292.
[76] Taslim, M.E.; Setayeshgar, L.; Spring, S.D. J. Turbomachinery 2001, 123, 147-153.
[77] Haiping, C.; Jingue, Z.; Taiping, H. ASME Paper No.2000-GT-220.
[78] Son, C.; Gillespie, S.; Ireland, P.; Dailey, C.M. ASME Paper No. 2001-GT-154.
[79] Gao, L.; Srinath, V.E. ASME Paper No.GT2003- 38178.
[80] Rhee, D.H.; Yoon, P.H.; Cho, H.H. Int. J. Heat Mass Transfer 2003, 46, 1049-1061.
[81] Huber, A. M.; Viskanta, R. Int. J. Heat Mass Transfer 1994, 37, 2859-2869.
[82] Huang, Y.; Ekkad, S. V.; Han, J. C. J Thermophysics and Heat Transfer 1998, 12, 7379.
[83] Metzger, D. E.; Korstad, R. J. J. Engineering for Power 1992, 94, 35-41.
[84] Florschuetz, L. W.; Metzger, D. E.; Su, C. C., J. Heat Transfer 1984, 106, 34-41.
[85] Bailey, J. C.; Bunker, R. S. ASME Paper No. GT-2002-30473.
[86] Andrews, G. E.; Abdul Hussain, R.A.A.; Mkpadi, M. C. Proceedings of IGTC 2003,
Paper No. IGTC2003Tokyo TS-075.
[87] Yoon, P.-H.; Rhee, D.-H.; Cho, H. H. Transaction of KSME (B) 2000, 24, 195-230.
[88] Epstein, A. H.; Kerrebrock, J. L.; Koo, J. J.; Preiser, U. Z., GTL Report No. 184.
[89] Aronstein, D. C. Effects of rotation on impinging jets for turbine cooling (Ph.D.
Thesis), University of Washington, USA, 1994.
[90] Mattern, C.H.; Hennecke, D. K. ASME Paper No. 96-GT-161.
[91] Glezer, B.; Moon, H. K.; Kerrebrock, J. L.; Bons, J.; Guenette, G. ASME Paper No. 98GT-214.
[92] Hsieh, S. S.; Huang, J. T.; Liu, C. F. J Heat Transfer 1999, 121, 811-818.
[93] Yan, W.M.; Mei, S.C.; Liu, H.C.; Soong, C. Y.; Yang, W.-J. Int. J. Heat Mass Transfer
2004, 47, 5235-5245.
[94] Hwang, J.J.; Chang, C.S. Int. J. Heat Mass Transfer 2001, 44, 1053-1063.
[95] Iacovides, H.; Kounadis, D.; Launder, B.E.; Li, J.; Xu, Z. J. Turbomachinery 2005, 127,
222-229.
[96] Hong, S.K.; Lee, D.H.; Cho, H.H. J. Mechanical Science and Technology 2008, 22,
1952-1958.
[97] Downs, S. J.; James, E. H. ASME Paper No. 87-H-35.
[98] Cho, H. H. Heat/mass transfer flow through an array of holes and slits (Ph.D. Thesis),
University of Minnesota, USA, 1992.
[99] Hollwarth, B. R.; Dagan, L. J. Engineering for Power 1980, 102, 994-999.
[100] Hollwarth, B. R.; Lehmann, G.; Rosiczkowski, J. J. Engineering for Power 1983, 105,
393-402.
[101] Nazari, A.; Andrews, G. E. Proceedings of 7th IGTC 1999, 2, 638-648.
[102] Cho, H. H.; Goldstein, R. J. Proceeding of the 3rd KSME-JSME Thermal Engineering
Conference 1996, 71-76.
[103] Cho, H. H.; Rhee, D. H. J. Turbomachinery 2001, 123, 601-608.
[104] Cho, H. H.; Rhee, D. -H.; Goldstein, R. J. J. Turbomachinery 2008, 130, 041003.

Applications of Impingement Jet Cooling Systems

67

[105] Ekkad, S. V.; Huang, Y.; Han, J. C. J. Thermophysics and Heat Transfer 1999, 13, 522528.
[106] Rhee, D. H.; Choi, J. H.; Cho, H. H. J. Turbomachinery 2003, 125, 74-82.
[107] Hong, S.K.; Cho, H.H. Transaction of KSME (B) 2006, 30, 1147-1154.
[108] Hong, S.K.; Rhee, D.-H.; Cho, H.H. Transaction of KSME (B) 2008, 32, 283-289.
[109] Hong, S.K.; Lee, D.H.; Kim, Y.D.; Cho, H.H. Transaction of KFMA 2008, 11, 24-30.
[110] Funazaki, K.; Tarukawa, Y.; Kudo, T.; Mastsuno, S.; Imai, R.; Yamawki, S. ASME
Paper No. 2001-GT-0148.
[111] Yamawki, S.; Nakamata, C.; Imai, R.; Mastsuno, S.; Yoshida, T.; Mimura, F.; Kumada,
M. ASME Paper No. GT-2003-38215.
[112] Rhee, D. H.; Nam, Y. W.; Cho, H. H. J. Turbomachinery 2005, 127, 615-626.
[113] Parsons, J. A; Han, J.-C. Int. J. Heat Mass Transfer 1998, 41, 2059-2071.
[114] Parsons, J. A.; Han, J.-C., ASME Paper No. 96-WA/HT-9.
[115] Hong, S.K.; Lee, D.H.; Cho, H.H. Applied Thermal Engineering 2009, 29, 2914-2920.
[116] Hong, S.K.; Lee, D.H.; Cho, H.H., Int. J. Heat Mass Transfer 2009, 53, 1373-1379.
[117] Hong, S.K.; Lee, D.H.; Cho, H.H. J. Heat Transfer 2010, 132, 114501.

You might also like