You are on page 1of 12

International Journal of Heat and Mass Transfer 53 (2010) 56295638

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


j o u r n a l h o m e p a ge : w w w.e l s e v i e r.c o m / l o c a t e / i j h m t

Flat plate heat transfer with impinging axial fan ows


Jason Stafford , Ed Walsh, Vanessa Egan, Ronan Grimes
Stokes Institute, Mechanical & Aeronautical Engineering Department, University of Limerick, Limerick, Ireland

a r t i c l e

i n f o

Keywords:
Impingement heat transfer
Flat plate
Axial fan
Infrared thermography
Flow eld

a b s t r a c t
Axial ow fans are widely used for the augmentation of heat transfer. One such use is an axial ow fan
impinging air on to a heated at plate which has many practical applications and has been experimentally investigated to characterize local heat transfer distribution. Using infrared thermography, a twodimensional prole of the heat transfer coefcient on a at plate is quantied for a range of fan speeds,
from 2000 to 6000 rpm, and fan to plate distances of 515 mm. The inuence of air ow interaction
with motor supports on the exit ow plane of an axial fan was shown to result in substantial gradients in
heat transfer coefcient in addition to non-uniform uctuations in heat transfer coefcient on the plate
sur- face. Local maxima in the heat transfer distribution were identied, and shown in some cases to
be a function of fan to plate distance, promoting heat dissipation signicantly for discrete heat sources.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
The axial ow fan is widely used in many engineering applications as its versatility has resulted in implementation into large
scale systems such as industrial dryers and air conditioning units,
to automotive engine cooling and in-cabin air recirculation systems. At the smaller scale, the use of axial ow fans for the
purpose of augmenting heat transfer is particularly evident in
electronics due to the possibility of relatively compact designs
which can be easily accommodated. The extended use of axial ow
fans for uid movement and heat transfer has resulted in detailed
research into the performance attributes of many designs.
However, the practi- cal example of impingement cooling of a
heated at plate using an axial ow fan has received less
attention despite possible uses such as surface drying, and also
cooling of discrete heat sources, common in printed circuit board
designs.
The performance aspects of axial ow fans have been examined
in the literature, and particularly applicable for the current study
are works which analyze the air ow distribution from axial fan
designs. The exit ow from axial fan outlets and at locations
axially downstream of the fan outlet have been investigated
through the use of Particle Image Velocimetry (PIV) [1,2]. Yen and
Lin [1] exam- ined three different designs including shrouded,
shroudless, and a shrouded winglet-blade design to conrm
relative performance enhancements of the design choices
mentioned. Each design also had varied geometric parameters
such as blade angle, number of blades, fan speed and hub to tip
ratios. Consequently, common trends in air ow distribution
revealed that such air ow patterns

Corresponding author.
E-mail address: jason.stafford@ul.ie (J. Stafford).

0017-9310/$ - see front matter


2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijheatmasstransfer.2010.08.020

are likely to exist for the majority of axial fan designs. Air ow
which expanded downstream of the fan exit ow plane was
observed, and related to the discontinuity of the shroud which
con- ned the ow to the axial direction.
Yoon and Lee [2] measured the velocity distribution in a similar
region for a forward swept design axial ow fan. Using stereoscopic PIV, a three-dimensional prole of the ow structures was
created using axial, tangential, and radial velocity components. In
plane velocity vectors were presented with out of plane velocity
contours, revealing the highly three-dimensional ow patterns
that are produced from an axial fan in operation. The expanding
ow pattern, apparent in the study by Yen and Lin [1], continued
to diverge at a constant angle after the fan outlet. The conical
shape of the jet results from centrifugal forces within the jet
inducing a radial ow component; and based on the literature
reviewed is synonymous with axial fan designs.
Estevadeordal et al. [3] also used the PIV technique for investigating the instantaneous and time-averaged velocity eld around
the fan blade pressure and suction surfaces. This was achieved by
synchronising the digital PIV system components to the blade
position. By throttling the inlet ow to the fan, various fan operating points could be examined to determine the effects of system
resistance on the local velocity eld. Both ow visualization and
velocity eld data highlighted unsteady aerodynamic excitations
such as parallel blade wakes, axial streaks, in addition to pressure
and suction side ow separation. At the trailing edge of the fan
blade, increasing pressure rise across the fan promoted ow
unsteadiness. Increased pressure rise also changed the stagnation
point on the blade from the leading edge for a recommended
oper- ating condition, to the pressure side of the blade. This in
turn can contribute to the generation of small eddies on the
suction side

J. Stafford
J. Stafford
et al.
et / al.
International
/ International
Journal
Journal
of Heat
of Heat
andand
Mass
Mass
Transfer
Transfer
53 (2010)
53 (2010)
56295638
56295638

563

563

Nomenclature
A
AC
AF
Cp
cmid
D
d
dh
E_
H
h
I
k
n
Nu
DP
Q
q
q00
r
Ra
Re
T
t
Ub
U
V
VA
Vabs
VT
W
dW
x, y

at plate area (m )
2
circular area of impingement (m )
2
fan area (m )
specic heat capacity (J/kg K)
chord length at blade midspan (m)
fan diameter (m)
annular slot width (m)
hub diameter (m)
energy rate (J/s)
plate distance from fan outlet (m)
heat transfer coefcient (W/m2 K)
current (A)
thermal conductivity (W/m K)
no. of recordings
Nusselt number
static pressure difference (Pa)
volumetric ow rate (m3/s)
heat transfer rate (W)
heat ux (W/m2)
radial direction from fan center (m)
Rayleigh number
Reynolds number
temperature (K)
thickness of layer (m)
blade velocity at midspan (m/s)
mean uid velocity (m/s)
voltage (V)
axial velocity (m/s)
absolute velocity (m/s)
tangential velocity (m/s)
width of plate (m)
uncertainty in the result
cartesian coordinates (m)

of the blade, leading to the unsteadiness observed at the trailing


edge [4].
Velarde-Suarez et al. [5] experimentally investigated the
unsteadiness at various locations near the inlet and outlet of a
var- iable pitch axial ow fan with a 600 mm tip and 380 mm
hub diameter. This type of design offered an increase in the range
of operation while maintaining fan efciency. High levels of
unstead- iness were observed at the hub and housing regions,
which were promoted for off-design operating conditions.
Understanding the uid ow patterns and locations of high
uid velocity and turbulence intensities in the exit ow of an axial
fan is also benecial for the design of thermal management
solutions. Previous work on heat transfer enhancement using low
prole ra- dial fan ows [6,7] provides an insight to the
improvements which can be achieved using various fan heat
sink arrangements. Egan et al. [6] presented the uid ow prole
produced by a miniature radial fan within a miniature heat sink.
By aligning the uid ow with the heat sink channels through
the use of a diffuser, an in- crease in the heat dissipation rate
was achieved. Regions of low velocity air ow were also
alleviated thus reducing the possibility of local temperature rises
on the surface. Similarly, Stafford et al. [7] determined that the
ow patterns produced from a radial fan within novel nless heat
sink designs enhanced heat transfer over that predicted by
theory.
In an experimental study of axial fan and heat sink thermal
solutions, Lin et al. [8] investigated two axial fan designs, with
high pressure low ow rate and low pressure high ow rate
charac- teristics, and also the thermal performance of two
heat sink

Greek symbols
e
emissivity
l
dynamic viscosity (kg/m s)
fan rotation
q
density (kg/m3)
r
StefanBoltzmann constant (W/m2 K4)
rh
normalized uctuation in heat transfer coefcient
s
time interval between frames (s)
Subscripts
aw
adiabatic wall
b
blade midspan conditions as reference
c, U
unsteady conduction
c
conduction
C
circular area of impingement
d
annular slot width as characteristic length
f
thin-foil (SS304)
F
fan
fc
forced convection
gen
input
i
instantaneous
in
into control volume
max
maximum
min
minimum
nc
natural convection
out
out of control volume
p
paint
r
radiation
st
stored
W
plate width as characteristic length
1
ambient

designs. A conventional CPU heat sink with a vertical planar n array was compared to a heat sink design which consisted of oblique
planar ns. It was concluded that the oblique nned design provided enhanced thermal performance due to the larger surface
area and ow acceleration between the ns. The high pressure fan
was also determined as the appropriate choice for the cooling
assembly which exhibited high system resistance.
The referenced studies by Egan et al. [6], Stafford et al. [7], and
Lin et al. [8] considered global measurements from a heat sink surface in terms of average heat transfer coefcient, or bulk thermal
resistance values. Up to recently, local heat transfer distributions,
resulting from fan velocity patterns, have lacked documentation.
An axial fan impinging air on a heated at plate can provide information on high heat transfer regions that correspond to the exit
ow distributions of the axial fan. This can be critically important
for electronic systems where parallel ow is the common initial
assumption. Sui et al. [9] investigated the exit ow patterns of an
axial ow fan operating in standalone, and also in the presence
of a at plate positioned downstream and normal to the fan outlet.
The existence of two shear layers was noted, using PIV in the
downstream region. The outer shear layer resulted from the interaction of the fan exit ow with the surrounding ambient uid. An
inner shear layer was also observed due to uid interaction with
the zero ow region beneath the fan hub. In the presence of a at
plate, ow recirculation occurred beneath the hub which reduced
the static pressure distribution at a location on the plate corresponding to the fan central axis. The lowest static pressure, and
greatest ow recirculation, was noted at a height to diameter

distance from the plate of H/D = 0.6. This closed recirculating ow


is also apparent for annular jet impingement as noted by Ichimiya
[10]. Interestingly, it was found that for an equivalent H/D = 0.65,
the recirculating effects produced a maximum in heat transfer in
the central region of the annulus. In a separate study on at plate
heat transfer performance, Sui et al. [11] examined the static pressure and heat transfer distributions at local radial points along the
impingement plate from the central axis of the fan. This onedimensional analysis was considered for two locations at a
distance from the impingement plate, and also for two different
designs of axial fan. Geometric dimensions of both fans were similar; however one design included additional hub ns to investigate if air ow structures beneath the hub could be promoted.
The static pressure coefcient was at a maximum corresponding
to the high velocity air ow which leaves the blade tip as noted
by [1,2]. The resultant stagnation point on the plate surface was
also the location of the maximum local heat transfer, often noted
in studies of an impinging annular jet on a heated at surface
[10,12].
A shrouded axial fan has an exit ow area which resembles a
circular annulus due to the zero ow condition at the fan hub.
Con- sequently, some similarities in local heat transfer may exist
with the case of an annular jet impinging on a heated at plate,
which has been analysed previously [10,12]. It should be noted
however, that for these studies the uid leaving the nozzle of the
annular jet and approaching the heated plate solely consists of an
axial veloc- ity component. As previously mentioned, the uid
leaving the annular area of an axial fan is highly threedimensional, and stud- ies by Grimes and Davies [13] and Grimes
et al. [14] show that the ow is unsteady and swirling, with a
dependency on static pres- sure rise. This type of uid motion has
been represented in imping- ing jets by the introduction of swirl
generators [1517], and also in TaylorCouette ow with axial
ows [18,19].
Lee et al. [15] experimentally investigated a turbulent swirling
round jet impinging on a at surface for various swirl rates and
plate distances from the nozzle exit. Heat transfer characteristics
were acquired using a thermosensitive liquid crystal sheet,
sprayed onto a gold-coated polyester substrate sheet which was
electrically heated. At low nozzle to plate spacing, the inclusion of
a swirl gen- erator provided increased heat transfer over the case
with no swirl generator. This can be accredited to the angle of
the approach velocity of the uid relative to
the plate
orientation, as regions of high shear are extended in the radial
direction over that created by a predominantly axial uid
velocity with no swirl. However, at large nozzle to plate
spacing, inducing swirl no longer proved
benecial in dissipating heat as the uid velocity had decreased
substantially before interaction with the plate surface. Alekseenko
et al. [16] examined the velocity eld downstream of an annular
swirling jet at a constant distance from the impingement plate,
for various swirl rates. Increasing swirl rate resulted in a reduction
of the recirculation zone beneath the jet centreline. The conventional jet with zero swirl rate also achieved a higher velocity magnitude near the impingement plate over all swirling jets. This
effect on the surface heat transfer measurements by Huang and
El-Genk [17] was also noted.
In the current paper, the inuences of velocity distributions
downstream of axial fan blades on local heat transfer are discussed
for an axial fan impinging air on a heated at plate. Infrared measurements of the heated plate are used to obtain two-dimensional
heat transfer proles over a wide range of fan rotational speeds. A
fan with an outlet diameter of 48.5 mm was employed and the fan
to plate distance was varied between 5 and 15 mm. This results in
a height to fan diameter ratio, H/D = 0.1030.309, respectively.
Although practical applications in electronic cooling with heat
sinks are typically for H/D 0.40.5 [8], the current study
examines thermal performance on a at surface absent of heat

conducting nned geometries which increase convective surface


area with increasing H/D. As the at plate surface retains a constant convective surface area independent of H/D, it is anticipated
that positioning the fan at lower H/D, without adversely effecting
fan performance above the recommended operating range, will
be benecial on heat transfer. The objective therefore is to determine the effect of fan to plate distance and fan rotational speed
on local heat transfer distributions.
2. Fan design
The heat transfer performance of an axial fan jet impinging a
heated at plate is investigated using infrared thermography and
a heated-thin-foil approach. The geometric specications of this
fan are included in Table 1. A nominal operating speed of
4000 rpm is recommended by the manufacturers, hence two additional rotational speeds above and below this nominal value are
examined. The fan outlet is shown in Fig. 1, indicating the
locations of the supports for the motor, which are positioned 120
apart, and also the fan hub and housing.
Fan performance was characterized experimentally using a test
facility designed in accordance with BS848 [20]. A detailed description of this measurement facility is provided in [21]. Measured static pressure and ow
rate data for each rotational speed
considered in
this study is
presented in
Fig. 2.
The
corresponding operating points with the inclusion of a at plate
downstream are also shown for each rotational speed. Operating
at the closest fan to plate dis- tance of 5 mm provides the greatest
resistance to air ow over the remaining heights of 10 and 15 mm.
For the latter fan to plate dis- tances, the inclusion of the at
plate has a minor inuence on fan ow rate. Static pressure is
reduced as the resistance to the uid movement in the axial
direction is attenuated.
Quin and Grimes [22] observed a sensitivity of Reynolds number on fan performance for Reb < 1980. Below this Reynolds number the increased ratio of viscous to momentum forces begins to
adversely affect fan performance. Fan performance was shown to
progressively reduce with further reductions in Reynolds number,
and conventional scaling laws [4] were no longer satised. A similar Reynolds number effect was noted for miniature centrifugal
fan designs [23]. In the current study the heat transfer analysis is
considered for values above the critical Reb. The range of Reynolds
number examined for axial fan impinging air ow on a at plate,
based on the denition in Eq. (1), is 2800 < Reb < 8500.

qU b cmid
Reb

where Ub and cmid are the blade velocity and chord length at the
blade mid radius.
3. Experimentation
The experimental approach for characterizing convective heat
transfer on a heated at plate as a consequence of the impinging
air ow directed from an axial fan is presented in the following
sections. Heat transfer measurement procedures are discussed

Table 1
Fan specications.
Diameter (mm)
Nominal rotor speed (rpm)
Hub-tip ratio
No. of blades
Motor support location

48.5
4000
0.51
7
Fan outlet

Motor support
(carrying electric cables)

Motor supports

120

Fan housing

Fan hub
(containing motor)
Fig. 1. Axial fan used to augment heat transfer from a heated at plate.

60
Fan curves (2000rpm, 4000rpm, 6000rpm)
H = 5mm
Operating points
H = 10mm
H = 15mm

50

P (Pa)

40

30

20

10

0
0

3
4
3
3
Q (m /s) x 10

Fig. 2. Fan and system characteristics due to the inclusion of a at plate


downstream.

proceeded with uncertainty analysis information on the measured


quantities.
3.1. Heat transfer analysis
The local heat transfer performance on the heated at plate was
quantied using infrared thermography and a heated-thin-foil
technique. A schematic of the experimental apparatus is presented
in Fig. 3, where the heated-thin-foil is representative of a heated
at plate being cooled by an axial ow fan. A stainless steel 304
grade foil with a measured thickness of 14.7 lm is clamped and
tensioned using copper busbars and a tensioning mechanism that
prevents deection of the at plate under the impinging forces of
the fan air ow. An electric current is passed through the electrically resistive thin-foil resulting in heating of the plate by Joule effect to produce a constant heat ux condition. The air leaving the
fan outlet is conned to exit in the radial direction, preventing
recirculation of heated air back into the fan inlet. The thermal
images of the at plate were acquired using a ThermaCam Merlin
camera with an InSb detector operating in the 35 lm MWIR spectral range. A 25 mm lens was used giving a eld of view of
22
16 and providing a temperature resolution of 312.5 lm. A
calibration of the infrared camera was

conducted to ensure

Fig. 3. Experimental conguration for heat transfer measurements with details of


heat transfer modes.

accurate temperature measurements [24]. A Type-K thermocouple


was used to obtain the ambient air temperature and was positioned 200 mm upstream of the fan inlet. Sixty thermal images
were recorded at 1 frame per second once the foil reached a quasi-steady state. In this state, time-varying uctuations in temperature were noted due to the turbulent and unsteady uid ow
impinging the thin-foil surface. In the time-averaged analysis, the
images were averaged to reduce noise and time-varying uctuations in the temperature prole that were a magnitude of 10 3 of
the averaged temperature map.
On the camera observation side, the thin-foil is coated with an
opaque matt black paint to provide a high emissivity surface. The
mean thickness of this coating was also measured and found to
be 21.8 lm. Both foil and paint thicknesses were measured to account for the contribution of tangential conduction in the energy
balance of Eq. (2) which has been shown by Stafford et al. [24] to
produce signicant errors in the forced convection heat transfer
coefcient if ignored at this scale. In Fig. 3, a control volume is
shown for the foil and paint layers indicating the modes of heat
transfer that exist in the current experimental arrangement. The
analysis is based around an energy balance where

2E

in

_ gen

_ out _ st

The time-averaged heat transfer coefcient is calculated from a

@2 T

@ T

@T

time-averaged temperature prole, therefore ltering any time-

q U kf tf kp t p @x2
@y2
c;

varying uctuations into the mean. Hence, the storage term


E_ st 0. In the analysis of unsteady heat transfer coefcients,
this term must be considered in the energy balance.
Eq. (3) denes the forced convection heat transfer coefcient
derived from the energy balance

where q and Cp are the material density and specic heat


capacitance.
The uctuations in heat transfer coefcient are presented relative to the time-averaged heat transfer coefcient over the record-

hfc

q00

00 q
nc

gen

00
where qgen

00

qr 00

qc

T aw

00

q Cp tf q Cp t p
@
f
f
p
p

ing time interval i = 1:n.

is the input heat ux, q00nc is the heat ux by natural con-

rh

q
Pn
1
hi
hfc
2n 1
i1

hfc

vection from the camera viewing surface, q00r is the radiation heat
ux [25], and q00c is the contribution of the conductive heat ux in
the foil and paint layers [24]. Eqs. (4)(7) were used to calculate

The normalized maximum and minimum uctuations in heat transfer coefcient on the at surface were also considered.

all of the above with the exception of q00nc which was measured
experimentally. In this experimentation, the foil surface was insulated on the fan outlet side, which is subject to forced convection

rh;max max h i

for experiments with the axial fan operational. The resistance to


heat ow was two orders of magnitude lower at the surface subject

rh;min min hi

hfc

fch i1:n

10
11

hfc
h
fc

i1:n

to natural convection than in the direction normal to the insulation


layer. The input heat ux, q00gen , was varied resulting in examination
5
6
of a range of Rayleigh numbers (10 10 ) similar to that experienced on the camera viewing side of the at plate during fan operational experiments. Temperature proles were recorded using an
infrared camera and the relationship between heat ux by natural
convection and Rayleigh number was evaluated. This relationship
is provided in Eq. (5) for the current experimental investigations,
where W represents the plate width. This agrees with previous
nd- ings in the literature for the fundamental case of a heated at
plate facing upwards [26,27].

where hi is the instantaneous heat transfer coefcient, and hfc is the


time-averaged heat transfer coefcient.
The Nusselt number is dened in Eq. (12) based on the characteristic length (D dh), or the slot width of the annular ow leaving the fan outlet. This characteristic length scale was chosen as it
represents the hydraulic diameter of the annular outlet ow which
impinges the at surface. Although fan to plate distance inuences
heat transfer, it is this length scale which primarily governs both
the area subjected to forced convection cooling and the outlet ow
dynamics of the annular jet.

00
qgen
VI
A

Nud hfcaird
k

0:2

q00nc

0:766RaW kair T

q00
c

where d is the slot width of the annular fan outlet, and kair is the
1T

q00r erT4

T 41

6
2

12

@ 2 @
! T
T

kf tf kp t p
@ x2 @ y2

Eq. (7) solves for the contribution of both the foil and high emissivity paint layer on tangential conduction. This form of the twodimensional conduction equation arises from an energy balance of
an innitesimal control volume and has been described in detail
by Stafford et al. [24]. The relevant modes of heat transfer into
and out of the control volume is illustrated in Fig. 3. As the Biot
number of both layers is low (Bi < 0.1), it can be assumed that the
same temperature eld exists throughout the stainless steel and
paint layers. Consequently, tangential conduction is described as
in Eq. (7) and normal conduction in the z-direction is negligible.
The maximum contributions of each mode of heat transfer dened
00
in Eq. (3) relative to the input heat ux, qgen
were: 10.5% for q00nc , 19%
for q00r , and up to 95% for qc00 .
The time-averaged heat transfer coefcient was solved using a
time-averaged temperature prole. The root-mean-square of the
uctuations in heat transfer coefcient was calculated to determine the effect of turbulence and uid unsteadiness on the surface
heat transfer distribution. This was achieved by solving for the
instantaneous heat transfer coefcient using the energy balance

uid thermal conductivity.


The relationship between heat transfer and uid dynamics can
be examined through the scaling of non-dimensional Nusselt number with Reynolds number. The previously dened Reb is inappropriate for this as it only characterizes fan aerodynamics and is
independent of the at plate heat transfer. Therefore, the Reynolds
number which characterizes the annular outlet ow impinging the
heated at plate can be dened as:

Red

q Ud

13

where U is the mean uid velocity at the fan exit, calculated using
the measured ow rate, Q. This provided a range 600 < Red < 3000.
The fan to plate distance, dened H in Fig. 3, was varied between 5 and 15 mm which results in 0.103 6 H/D 6 0.309.
The range of fan speeds examined for each fan to plate distance
was
20006000 rpm.
3.2. Uncertainty
The inuence of measurement uncertainties on the calculated
forced convection heat transfer performance has been accounted
for using an uncertainty analysis [28]. This uncertainty analysis is
based around a worst case combination approach, evaluated
through the use of Eq. (14)
n

o1=2

of Eq. (2) over the recording interval. The storage term E_ st can
no longer be neglected when determining the instantaneous
heat
transfer and an unsteady state conduction equation must be
imple- mented to replace Eq. (7). This contains the additional
effect of heat ow over time interval, s.

dW
XN

W X1dX1

W X1 W X2dX2

W X2

W XNdXN

W
14

where each WX+dX WX term denes the result as a function of an


independent variable. The uncertainties in the measured values of
voltage, current, temperature and area were estimated as 5 mV,

hfc (W/m K)
120

a
0.07

100

0.06
80

0.05
Y (m)

5 mA, 0.1 K and 1 10 6 m2. Uncertainties in the measurement of


pressure and volumetric ow rate were 5% and 2.8%, respectively.
The maximum uncertainty in the heat transfer coefcient, Nusselt
and Reynolds numbers were estimated at 10.9%, 11.2% and 3.1%,
respectively. The optical tachometer used for measuring fan rotor
speed has an accuracy related to the resolution limit of 1 rpm. For
the fan speeds considered however, uncertainty was noted as
approximately 20 rpm, due to variations in speed monitored over
the test duration. Experimental uncertainty bands have been neglected when presenting the data for clarity.

60

0.04
0.03

4. Results and discussion

0.02

Fig. 4 presents local heat transfer coefcient distributions on a


heated at plate due to the impinging air ow of an axial ow
fan, which operates at a nominal rotational speed of 4000 rpm.
Fan to plate distances of 5, 10, and 15 mm are presented with
the corresponding H/D also labeled in the gure captions. The fan
outlet geometry is superimposed on the local heat transfer maps,
and is centered over the measurement region which spans an area
approximately 2 1.5D. With reference to Fig. 1, the location of
outlet ow regions, motor supports, and fan hub are all discernible.
The local heat transfer coefcient maps of Fig. 4 distinguish the
complex patterns which emerge as a result of the three
dimension- ality of the impinging air ow combined with a ow
interaction with the motor supports positioned on the fan exit
ow plane. Six regions of increased heat transfer are evident
which can be attributed to the fan blade motor support
interaction. The effect of fan blade motor support interactions
have been investigated in axial fan designs, however mainly in
relation to acoustic emissions [29,30]. As the fan blades pass over a
motor support, the uid from the pressure side of the blade is
forced to divert around the support obstruction. Consequently,
streamlines converge at either side of the support, resulting in
local maxima in the exit air velocity approaching the heated
plate further downstream. Hennissen et al. [31] also observed
this local rise in velocity magnitude when analyzing the exit ow
prole of a standalone axial ow fan with three motor supports
on the exit ow plane. As expected, a sudden drop in exit velocity
to zero was noted directly beneath the motor supports. In the
current study, this is reected in the heat transfer measurements
particularly at the closest fan to plate distance in Fig. 4(a), as
three regions of decreased heat transfer coefcient ap- pear
between the local maxima and resemble the support geometries. These three regions are at an angular offset to the position
of the motor supports upstream of the at plate. This is caused by
the swirl angle of the exit air ow, as the offset is in the fan
rotational direction and increases with increasing H/D.
Furthermore, the po- sition of the six peaks in heat transfer
coefcient on the at plate is also inuenced by the swirling air
ow with 50% of them situated directly beneath the motor support for the range of H/D
exam- ined. As H/D is increased, these discrete regions of
enhancement in heat transfer become easily dened as they are
further isolated from each other due to the expanding air ow
downstream of the axial fan. The position of these peaks in heat
transfer coefcient however, do not change substantially over
the fan to plate dis- tances examined which is possibly due to
the relatively small span of H/D considered. In particular, the
three peaks that are situated directly beneath the motor supports
are at almost equal locations, independent of H/D. As the motor
support thickness in the axial direction is approximately one
third of the width of the largest mo- tor support, the air ow
leaving the fan blade may be deected off the supports towards
the axial direction. Such a deection would reduce the swirl
angle of the exit ow in this local region. This hypothesis is
illustrated in Fig. 5. Consequently, the location of these peaks
remains unchanged for the range of H/D considered.

0.01

40
20
0.02

0.04

0.06

0.08

0.1

X (m)
hfc (W/m2K)
120

b
0.07

100

0.06
80

Y (m)

0.05
0.04

60

0.03

40

0.02
20

0.01
0.02

0.04

0.06

0.08

0.1

X (m)
hfc (W/m2K)
120

c
0.07

100

0.06
80

Y (m)

0.05
0.04

60

0.03

40

0.02
20

0.01
0.02

0.04

0.06

0.08

0.1

X (m)

Fig. 4. Forced convection heat transfer coefcient for 4000 rpm and a fan to plate
distance of (a) 5 mm (H/D = 0.103), (b) 10 mm (H/D = 0.206), and (c) 15 mm (H/
2

D = 0.309). Contour level: 5 W/m K.

This is reinforced when analyzing the change in position of the


remaining three peaks in heat transfer, which are located beneath
the outlet ow area of the fan. These peak locations are no longer
independent of H/D as the absolute velocity entails a greater swirl
angle.
A numerical study on the acoustic emissions due to fan blade
motor support interaction for circular motor supports by Lu et al.
[29] supports this hypothesis. Pressure contours around the motor
support indicate a stagnation point on the support surface corresponding to the left side of the motor support as shown in Fig. 5.
This suggests that uid impingement and therefore deection ex-

Fan hub

Motor support

VA

Vabs
VT

Impingement
zone

40

10

Flat plate
Fig. 5. Section view of an axial fan above a at plate with hypothesized
interaction of the absolute velocity eld with a motor support.

0.1

0.2

0.3

0.4
r/D

0.5

0.6

0.7

0.8

Fig. 6. Mean radial distribution of Nusselt number for 4000 rpm and 0.103 6 H/
D 6 0.309.

60

2000 rpm
Fan hub

Fan blade

4000 rpm
6000 rpm

50
40

Nud

occur beneath the three motor supports, as opposed to directly beneath the outlet ow areas. In fact, for each fan to plate distance
examined, the maximum peak in forced convection heat transfer
is a consequence of the blade-support interaction for the largest
motor support which carries electric wiring to the motor. Through
these ndings, benets in positioning of discrete heat sources for
maximum heat transfer can therefore be attained.
Overall, the region directly beneath the fan hub provides the
lowest heat transfer performance with local heat transfer coef2
cients as low as 6 W/m K at H/D = 0.103 for the lowest fan speed
of 2000 rpm examined. As there is no air ow directly impinging
this region, the majority of heat is dissipated through mixed con0
vection, as low velocity air ows exist in this region. In a study
of
the plate
velocity
eld between
an axial Sui
fan et
impinging
air normal
to
a at
positioned
downstream,
al. [9] concluded
that
as H/D was increased to 0.6, the strength of uid recirculation beneath the fan hub also increased. Similarly, a recent experimental
study of the velocity eld downstream of a swirling annular jet
by Yang et al. [34] also indicated strong recirculation of uid below
the centerline region of the jet. This is also reected in the current
heat transfer study as strengthening ow recirculation improves
the heat transfer performance to approximately 15 W/m2 K on
the at plate directly beneath the hub centre for H/D = 0.309 and
2000 rpm. The substantial degradation of heat transfer in this
region again emphasizes the importance in positioning of discrete
heat sources previously discussed. As an example, if a 10
10
mm heat source was centrally positioned directly beneath the fan
cen- tre on the at plate, an average heat transfer coefcient of
19 W/ m2 K would be achieved at the nominal speed of 4000 rpm
and a H/D = 0.103. Alternatively, if this heat source was
positioned at the peak in heat transfer beneath the widest motor
support, as shown in Fig. 4(a),
a mean heat transfer
2
coefcient of 104 W/ m K could be achieved. This equates to
over vefold increase in the amount of heat which can be
dissipated for the same chip to ambient temperature difference by
positioning the heat source just
20 mm away from the original position under the fan centre. As
the fan to plate distance increases to H/D = 0.309 this margin
reduces, however it is still a substantial factor of three.
Figs. 6 and 7 describe the radial heat transfer development by
computing the circumferential average of the local dimensionless
Nusselt number along radial points from the fan central axis. Sections directly beneath the fan hub and blade passage are also dened. In Fig. 6, the Nusselt number is presented for each H/D
examined and at the nominal fan rotational speed of 4000 rpm.

30
20

Impingement
zone

ists in this region. Beneath the support, the referenced authors


show a reduction in pressure, which is similar to that observed
in the study of bodies subjected to a cross ow. The shear layer
separates from the body and vortex shedding occurs resulting in
unsteadiness in the near wake region [32,33].
In Fig. 4(a and b), the largest peaks in heat transfer coefcient

Fan blade

50

Swirl angle

Nud

Stagnation
region

H/D = 0.103
H/D = 0.206
H/D = 0.309

60

Ub

30
20
10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

/D
Fig. 7. Mean radial distribution of Nusseltr number
for range of rotational speeds
examined and H/D = 0.309.

Beneath the fan hub, increasing H/D increases the heat transfer
performance, which has been previously discussed through local
heat transfer coefcient maps presented in Fig. 4. Near the fan
hub edge, a cross-over in the heat transfer performance occurs.
In the subsequent region beneath the fan blade, the Nusselt number increases as r/D = 0.5 is approached, where a peak in Nud is
apparent for the closest fan to plate spacing of H/D = 0.103. As H/
D increases, the Nusselt number decreases, and the peak in heat
transfer shifts in the direction of increasing r/D. This is due to the
diverging conical velocity prole synonymous with axial ow fans.
The air ow leaving the fan blades is no longer conned by the fan
hub or outer shroud downstream of the fan exit plane. Consequently, the centrifugal forces within the annular jet produce a radial component which expands the velocity distribution with
increasing distance downstream. Sui et al. [9] showed that by
increasing H/D, the location of peak axial velocity magnitude which
impinges the at plate increases to r/D > 0.45. The inuence of this
impingement is also noted on the heat transfer results in Figs. 6
and 7 where the peaks in heat transfer are similarly at r/D > 0.45
for the range of fan to plate distances examined. Finally, in the region outside the core ow from the axial fan, there is a second
cross-over at r/D 0.65 and the Nusselt number for the closest
fan to plate spacing drops below that of the larger distances. This
is due to the spreading of the air ow distribution leaving the
shrouded fan, which impinges a greater surface area of the at
plate for increases in H/D.

In Fig. 7, the inuence of fan rotational speed on the average


ra- dial distribution of the Nusselt number is presented for
H/ D = 0.309. The radial heat transfer distribution agrees for all
fan speeds examined. Therefore the discussion of the heat transfer
re- sults for the nominal rotational speed is an appropriate
generaliza- tion of the heat transfer distribution on the at plate
for the entire range of fan speeds considered.
In Fig. 8(a), the root-mean-square uctuations in heat transfer
coefcient over
the time-averaged data for the recording
frequency considered are shown to be a maximum of 40%. A
fan to plate

h 0.4

0.35

0.07

0.3

0.06

0.25

Y (m)

0.05

0.2

0.04

0.15
0.03

0.1

0.02

0.05

0.01

0
0.02

0.04

0.06

0.08

0.1

X (m)

h,max

1.2

0.07

0.06

Y (m)

0.05

0.8

0.04

0.6

0.03
0.4
0.02
0.2

0.01
0.02

0.04

0.06

0.08

0.1

X (m)

h,min

0.07

0.2

0.06

Y (m)

0.05

0.4

0.04

0.6

0.03
0.8
0.02
1

0.01
0.02

0.04

0.06
X (m)

0.08

0.1

1.2

spacing of H/D = 0.309 is presented, while the remaining fan to


plate distances examined also produced a similar magnitude and
pattern in heat transfer uctuations. Areas under the hub experience uctuations in heat transfer up to 80100% due to the
unsteadiness in air ow recirculation. The position of the motor
supports at the fan exit results in an increase in heat transfer coefcient uctuations on the at plate, caused by the blade-support
interaction. As the fan blade passes over a motor support, local
pressure variations result in increased velocity uctuations combined with wake shedding on the strut. As previously discussed,
similarities between the observations at the rear of a body in cross
ow and the region beneath a motor support exists. In this region,
small eddies are created which are initiated as the uid diverts the
motor support. These turbulent structures continue downstream in
the direction of the swirling ow until impingement on the at
surface. The maximum and minimum heat transfer uctuations
in Fig. 8(b and c) suggest that these unsteady structures drive
the heat transfer in this region. In contrast, heat transfer uctuations in the areas unaffected by the motor support interactions
have maximum uctuations of 20% with rh typically less than
5%. This implies that the effect of turbulence generated by the
aerofoil blades in the form of tip vortices, blade wakes and separation, has a relatively minor effect on the surface heat transfer
coef- cient. Of course this is presented for an acceptable fan
operating point, and based on the literature reviewed [3,5,13,14],
an increase in the magnitude of the heat transfer uctuations will
occur when operating outside the recommended design point. The
largest uc- tuations are a direct consequence of the interaction
of the outlet ow and the widest motor support carrying electrical
wiring. These uctuations are in a region of low heat transfer
coefcient, shown
in Fig. 4(c). Interestingly, the uid interaction with the motor support provides a local increase of heat transfer coefcient
hfc 120 W/m2 K, but also results in a wide band of low heat trans2
fer coefcient hfc 30 W/m K which is highly unsteady
( rh
25 40%). The distribution of the maximum and minimum
lo- cal time-varying uctuations in heat transfer coefcient is
presented in Fig. 8(b and c). The magnitude of the uctuations
reinforces the signicance of unsteadiness and turbulence in the
outlet ow on surface heat transfer. Often the uctuating nature
of unsteady fan ows is overlooked when analyzing heat transfer
performance using time-averaged information. The data presented
in Fig. 8 indicates regions to avoid in the positioning of discrete
heat sources, such as electronic components, with reliability
which may be adversely inuenced by cyclic thermal loading.
Although the primary practical application of these results is towards enhancing electronics cooling, the arrangement presented
also applies to many other engineering scenarios. Akturk et al.
[35] presented an experimental study on uninhabited aerial vehicles which implement ducted axial fan designs to produce thrust
to hover above ground level. In order to achieve this, the axial
fan forces air normal to the surface, similar to the experimental
arrangement examined in this study. Akturk et al. [35] concluded
that the introduction of a cross ow, representing the vehicle in
forward ight, resulted in excessive moment imbalance in the
axial jet. Consequently, vehicle instability is promoted with this
non- uniformity in exit ow. As seen in Figs. 4 and 8, a clear
moment imbalance is also evident due to the position and
geometry of the axial fan motor supports, and the exit ow is
asymmetric when considering the analogy between uid ow and
heat transfer. The proportionality between heat and mass transfer
coefcients is an- other analogy which extends the application of
this work. In drying processes, the drying rate is dependent on
uid velocity, tempera- ture and solvent content of the drying air
[36]. Stress sensitive surFig. 8. Local (a) root-mean-square, (b) maximum, and (c) minimum heat transfer
coefcient uctuations normalized with the time-averaged local heat transfer

coefcient at 4000 rpm and H/D = 0.309. Contour levels: (a) 0.05; (b) and (c) 0.2.

faces often require geometrically different jet designs, whose


surface heat and mass transfer performance have been considered
in detail in the literature [37]. Figs. 4 and 8 highlight the possible

steady nature of axial fan ows over similar Reynolds number


annular jet ows. This is conrmed by the previous discussion on
the heat transfer uctuations which exist on the surface of the at
plate. The primary source of this unsteadiness has been accredited
to the interaction between the outlet ow and motor supports on
the exit plane. The inuence of low velocity uctuations beneath
the fan hub are also highlighted in Fig. 9, as the heat transfer data
deviates from this relationship with Reynolds number. This is noticed over the H/D range examined suggesting that buoyancy induced convection is evident in this region which is supported by
the local heat transfer coefcient proles of Fig. 4 and therefore
different scaling can be expected to represent this mixed ow
region.
Fig. 10 describes the mean Nusselt number distribution for a
range of impingement area to fan area conditions. The impingement area is expressed as a circular region centered on the fan
axis. The mean Nusselt number therefore, is the average nondimen- sional heat transfer within this impingement area. The
variation of the mean Nusselt number over 0.103 6 H/D 6 0.309
is greatest at the fan central axis. A cross-over appears at Ac/AF
0.5, where operating at H/D = 0.103 begins to provide higher area
average heat transfer performance. At Ac/AF < 0.5, operating at
the larger H/D provides increased thermal performance beneath
the fan hub through strengthened uid circulation [9]. However
the most prac- tical region of interest is for AC/AF P 1, where the
air jet impinges on an area equal to or greater than the fan area.
At AC/AF = 1, the maximum difference in the mean Nusselt
number between H/ D = 0.103 and H/D = 0.309 is approximately
7%. This difference de- creases to 5% for the remaining AC/AF
values, suggesting that the mean heat transfer performance is
less sensitive to variation in fan to plate distance for the range
examined here. This can be attributed to the velocity prole
expanding downstream of the fan outlet as H/D increases,
discussed previously through
the
local heat
transfer
measurements. Although peaks in heat transfer de- crease with
increasing H/D, the axial swirling jet is simultaneously expanding
and therefore cooling a larger surface area. At AC/AF = 2, the mean
Nusselt number for H/D = 0.103 begins to level at a con- stant
value of Nud
28. The other fan to plate distances appear to be
approaching a similar magnitude of mean non-dimensional heat

30
25

H/D = 0.103
H/D = 0.206
H/D = 0.309

20
Nud

issues with axial fan impingement if drying uniformity is essential


to the mass transport process.
The distribution of the Nusselt number with radial position for
annular jets has been shown to scale with Re0.55 [10]. In the current
study for impinging axial fan ows, the scaling relationship is
found to be Re0.6 as shown in Fig. 9 and applies to the entire H/D
range examined. The increase in the Reynolds number exponent
over the annular jet study by Chattopadhyay [12] reects the un-

15

10
5
0
0

0.5

1.5
A /A
C

2.5

Fig. 10. Variation of mean Nusselt number for scaled heights examined and a
nominal fan speed of 4000 rpm.

transfer also. For AC =AF


2:6, a decrease in Nud is anticipated as
the inuence of the axial fan air ow on augmenting heat transfer
will weaken for extended r/D locations. At the point of impingement, the majority of air is deected in a radial direction and
boundary layer growth is similar to that of the conventional case
of ow over a at plate. Consequently, the velocity gradient at
the surface and hence the mean heat transfer rate must begin to
decrease in the radial direction beyond the point of impingement.
This decrease in Nud is noted in Fig. 9 for the circumferentially
averaged data and r/D > 0.55. As Fig. 10 presents the area averaged
Nud , this decrease is not yet reected in the mean heat transfer.
It is shown to be more benecial for area average heat transfer
applications to operate at H/D = 0.103, particularly where the
application is within a space constrained environment. It should
be noted however that further reductions in H/D may not
necessar- ily be advantageous, as the operating point for the fan
will result in further decreases in ow rate (Fig. 2). The fan may
also be forced to operate in the upper half or stalling region of the
fan characteristic curve with the additional system resistance
produced for
H/ D < 0.103. In this region fan efciency is
reduced, noise increases, and non-uniform airow across the fan
blade surface exists [4,14].
Finally, the non-uniformities in surface heat transfer have been
attributed to the fan blade motor support interaction. Therefore
reducing motor support thickness would conversely result in a uniform heat transfer distribution approaching that of an annular jet.
5. Conclusions

0.45
2000 rpm
4000 rpm
6000 rpm

0.4
0.35
Nud / Re0.6
d

0.3
0.25
0.2
0.15
0.1
0.05
0
0

0.1

0.2

0.3

0.4
r/D

0.5

0.6

0.7

0.8

Fig. 9. Scaling of Nusselt number with Reynolds number at H/D = 0.309.

Heat transfer using an axial ow fan impinging air on a heated


at plate has been investigated experimentally. Infrared thermography and a heated-thin-foil technique were implemented to acquire full eld measurements of the heat transfer coefcient over
the fan impingement zone. The interaction of air ow leaving the
fan blades with the motor supports on the exit ow plane of an axial fan were shown to result in localized maxima in the heat transfer coefcient. In addition, uctuations in surface heat transfer
coefcient were observed, and directly linked to the unsteady uid
structures generated by this interaction. Subsequently, signicant
enhancements in heat transfer can be achieved with accurate heat
source positioning. An increase in fan to plate distance results in
local peaks in heat transfer moving outwards on the plate surface
from the fan central axis. This has been attributed to the conical
exit ow distribution. The inuence of fan rotational speed has
been shown to be minimal on the location of increased heat transfer peaks for the range examined in this work.

References
[1] S.C. Yen, F.K.T. Lin, Exit ow eld and performance of axial ow fans, J. Fluids
Eng. 128 (2006) 332340.
[2] J.-H. Yoon, S.-J. Lee, Stereoscopic PIV measurements of ow behind an isolated
low-speed axial-fan, Exp. Therm. Fluid Sci. 28 (2004) 791802.
[3] J. Estevadeordal, S. Gogineni, W. Copenhaver, G. Bloch, M. Brendel, Flow eld in
a low-speed axial fan: a DPIV investigation, Exp. Therm. Fluid Sci. 23 (2000)
1121.
[4] F.P. Bleier, Fan Handbook: Selection, Application and Design, McGraw-Hill,
1997.
[5] S. Velarde-Suarez, R. Bellesteros-Tajadura, C. Santolaria-Morros, E. BlancoMarigorta, Total unsteadiness downstream of an axial ow fan with variable
pitch blades, J. Fluids Eng. 124 (2002) 280283.
[6] V. Egan, J. Stafford, P. Walsh, E. Walsh, An experimental study on the
performance of miniature heat sinks for forced convection air cooling, J. Heat
Transfer 130 (7) (2009). 071402 (19) pp.
[7] J. Stafford, E. Walsh, V. Egan, P. Walsh, Y.S. Muzychka, A novel approach to low
prole heat sink design, J. Heat Transfer 132 (9) (2010). 091401 (18) pp.
[8] S.C. Lin, F.S. Chuang, C.A. Chou, Experimental study of the heat sink assembly
with oblique straight ns, Exp. Therm. Fluid Sci. 29 (2005) 591600.
[9] D. Sui, S.S. Wang, J.R. Mao, T. Kim, T.J. Lu, Exit ow behavior of axial fan ows
with/without impingement, J. Fluids Eng. 131 (2009). 061103 (17) pp.
[10] K. Ichimiya, Heat transfer characteristics of an annular turbulent impinging jet
with a conned wall measured by thermosensitive liquid crystal, Heat Mass
Transfer 39 (2003) 545551.
[11] D. Sui, T. Kim, M.L. Xu, T.J. Lu, Novel hub ns for axial ow fans for the
enhancement of impingement heat transfer on a at plate, J. Heat Transfer
131 (2009). 074502 (13) pp.
[12] H. Chattopadhyay, Numerical investigations of heat transfer from impinging
annular jet, Int. J. Heat Mass Transfer 47 (2004) 31973201.
[13] R. Grimes, M. Davies, Air ow and heat transfer in fan cooled electronic
systems, J. Electronic Packag. 126 (2004) 124134.
[14] R. Grimes, M. Davies, J. Punch, T. Dalton, R. Cole, Modeling electronic cooling
axial fan ows, J. Electronic Packag. 123 (2001) 112119.
[15] D.H. Lee, S.Y. Won, Y.T. Kim, Y.S. Chung, Turbulent heat transfer from a at
surface to a swirling round impinging jet, Int. J. Heat Mass Transfer 45 (2002)
223227.
[16] S.V. Alekseenko, A.V. Bilsky, V.M. Dulin, D.M. Markovich, Experimental study of
an impinging jet with different swirl rates, Int. J. Heat Fluid Flow 28 (2007)
13401359.
[17] L. Huang, M.S. El-Genk, Heat transfer and ow visualization experiments of
swirling, multi-channel, and conventional impinging jets, Int. J. Heat Mass
Transfer 41 (3) (1998) 583600.
[18] S.T. Wereley, R.M. Lueptow, Velocity eld for TaylorCouette ow with an
axial ow, Phys. Fluids 11 (12) (1999) 36373649.

[19] J.Y. Hwang, K.S. Yang, Numerical study of TaylorCoutte ow with an axial
ow, Computers Fluids 33 (2004) 97118.
[20] BS848: Fans for general purposes, Part 1: methods for testing performance,
1980.
[21] R. Grimes, E.J. Walsh, D. Quin, M. Davies, The effect of geometric scaling on
aerodynamic performance, AIAA J. 43 (11) (2005) 22932298.
[22] D. Quin, R. Grimes, The effect of Reynolds number on microaxial ow fan
performance, J. Fluids Eng. 130 (2008). 101101 (110) pp.
[23] P. Walsh, V. Egan, R. Grimes, E. Walsh, Prole scaling of miniature centrifugal
fans, Heat Transfer Eng. 30 (1) (2009) 130137.
[24] J. Stafford, E. Walsh, V. Egan, Characterizing convective heat transfer using
infrared thermography and the heated-thin-foil technique, Meas. Sci. Tech. 20
(10) (2009). 105401 (111) pp.
[25] F.P. Incorpera, D.P. DeWitt, Fundamentals of Heat and Mass Transfer, fourth
ed., John Wiley and Sons, 1996.
[26] W.M. Lewandowski, E. Radziemska, M. Buzuk, H. Bieszk, Free convection heat
transfer and uid ow above horizontal rectangular plates, Applied Energy 66
(2000) 177197.
[27] T. Fujii, H. Imura, Natural convection heat transfer from a plate with arbitrary
inclination, Int. J. Heat Mass Transfer 15 (1972) 755767.
[28] J.R. Moffat, Uncertainty analysis, in: K. Azar (Ed.), Thermal Measurements in
Electronics Cooling, CRC Press, 1997, pp. 4580.
[29] H.Z. Lu, L. Huang, R.M.C. So, J. Wang, A computational study of the interaction
noise from a small axial-ow fan, J. Acoustic Soc. Am. 122 (3) (2007) 1404
1415.
[30] L. Huang, J. Wang, Acoustic analysis of a computer cooling fan, J. Acoustic. Soc.
Am. 118 (4) (2004) 21902200.
[31] J. Hennissen, W. Temmerman, J. Berghmans, K. Allaert, Modelling of axial fans
for electronic equipment, in: Proceedings of EURO-THERM Seminar, vol. 45,
Leuven, Belgium, 2022 Sept, 1995, pp. 309318.
[32] M. Ozgoren, Flow structure in the downstream of square and circular
cylinders, Flow Meas. Inst. 17 (2006) 225235.
[33] A. Lankadasu, S. Vengadesan, Onset of vortex shedding in planar shear ow
past a square cylinder, Int. J. Heat Fluid Flow 29 (2008) 10541059.
[34] H.Q. Yang, T. Kim, T.J. Lu, K. Ichimiya, Flow structure, wall pressure and heat
transfer characteristics of impinging annular jet with/without steady swirling,
Int. J. Heat Mass Transfer 53 (2010) 40924100.
[35] A. Akturk, A. Shavalikul, C. Camci, PIV Measurements and Computational Study
of a 5-Inch Ducted Fan for V/STOL U AV Applications, in: 47th AIAA Aerospace
Sci. Meeting and Exhibit 2009, Orlando, Florida, 58 January, 2009.
[36] A. Avci, M. Can, The analysis of the drying process on unsteady forced
convection in thin lms of ink, App. Therm. Eng. 19 (1999) 641657.
[37] F. Peper, W. Leiner, M. Fiebig, Impinging radial and inline jets: a comparison
with regard to heat transfer, wall pressure distribution, and pressure loss, Exp.
Therm. Fluid Sci. 14 (1997) 194204.

You might also like