You are on page 1of 6

International Journal of Heat and Mass Transfer 83 (2015) 21–26

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Heat transfer in turbulent planar offset attaching jets with small initial
offset distances
N. Gao ⇑, D. Ewing, C.Y. Ching
Department of Mechanical Engineering, McMaster University, Hamilton, Canada

a r t i c l e i n f o a b s t r a c t

Article history: Measurements of the mean heat transfer coefficient from the wall below turbulent offset wall attaching
Received 23 November 2013 planar jets were performed for jets with Reynolds numbers (Rej ) of 21,800–54,500. The jets were initially
Received in revised form 10 November 2014 offset 0.2–1 times the initial jet height (Hj ) above the wall. The Reynolds number dependence of the
Accepted 26 November 2014
maximum Nusselt number varied with offset distance from Re0:64 j for the flows with an offset distance
of 0.2Hj to Re0:77
j for flows with initial offset distances of Hj . The distance from the jet exit to the location
of the maximum heat transfer coefficient also increased with Reynolds number for the jets with
Keywords:
smaller offset distances. The maximum heat transfer coefficient for the different offset jets could be
Turbulent planar offset jet
Heat transfer
correlated by hX m =ka ¼ cðU j X m =mÞ0:74 where X m is the distance to the maximum heat transfer location
and the parameter c varies with the initial offset distance from 0.08 to 0.1. The profiles of the Nusselt
number approach the results for a wall jet downstream of x=Hj  8 and collapsed when scaled by
Re0:77
j in all cases.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction the inner shear layer of the jet causes a recirculating flow region
below the jet in all cases; and the pressure differences associated
Turbulent offset wall attaching planar jets, formed when a pla- with this region causes the jet to approach the wall similar to flows
nar jet exits an outlet near an adjacent surface, are used in a range over a backward facing step [7–11] or through a sudden expansion
of cooling and drying applications. The heat transfer coefficient [12,13]. The nature of this process differs in planar offset jet flows
from the wall to the jet increases to a maximum where the jet with small and large initial offset distances. For jets with small ini-
attaches to the wall before decreasing as the jet recovers and tran- tial offset distances, the flow attaches to the wall before the jet
sitions to a wall jet flow [1]. The heat transfer coefficient where the becomes fully developed so that the inner shear layer of the jet
jet attaches to the wall decreases if the initial offset distance of the interacts with the wall before the outer shear layer of the jet. For
jet is increased for a given jet Reynolds number. For example, Kim jets with large offset distances, the jet is fully developed before
et al. [1] found the maximum Nusselt number decreased by the jet interacts with the wall. The trajectory of the jet towards
approximately 50% when the initial offset distance (Hs ) was chan- the wall also changes with offset distance. The mean attachment
ged from 0.5 to 5 times the initial jet height size (Hj ) for planar jets distance appears to be proportional to the initial offset distance
initially parallel to the wall, of interest here. The maximum Nusselt of the jet for flows with Hs K 0:2Hj [6], but the distance to the
number for the flows with Hs =Hj ¼ 5 varied as Re0:49 j for Reynolds mean attachment point of the jet decreases relative to Hs for larger
number based on the jet height (Rej ) between 6000 and 30,000 offset distances [3,6] causing the curvature of the jet toward the
in this case [1]. The change in the heat transfer coefficient with wall to change.
Reynolds number for other initial offset distances was not exam- The effect that these changes in the development of offset
ined and does not appeared to have been examined in detail attaching jets have on the Reynolds number dependence of the
elsewhere. heat transfer coefficient has not been considered in detail. The heat
The development of the flow field in planar offset jets changes transfer produced by planar offset jets with Reynolds numbers of
with the initial offset distance of the jet [1–6]. Entrainment into 21,800 to 54,500 exiting a long channel initially offset 0.2Hj to Hj
above the wall was considered here. The heat transfer coefficients
⇑ Corresponding author at: School of Aeronautics and Astronautics, Dalian were measured using a thin electrically heated foil as the heat
University of Technology, China. transfer surface and a thermal camera to measure the surface tem-
E-mail address: gaonan@dlut.edu.cn (N. Gao). perature of the foil. The experimental facility and methodology

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.11.086
0017-9310/Ó 2014 Elsevier Ltd. All rights reserved.
22 N. Gao et al. / International Journal of Heat and Mass Transfer 83 (2015) 21–26

Nomenclature

A area of the heat transfer surface, m2 Tj jet temperature, K


Hj height of the jet, m Tw foil temperature, K
Hs offset distance from lower edge of the jet to the wall, m T1 ambient temperature, K
h convective heat transfer coefficient, W/m2 K U mean stream-wise component of the local velocity, m/s
I electric current passing through the stainless steel foil, U cl mean stream-wise velocity at the centerline of jet exit,
amp m/s
ka thermal conductivity of air, W/m K U1 free stream velocity, m/s
L length of the channel forming the jet, m Uj flow rate averaged stream-wise velocity at the jet exit,
Nuj Nusselt number based on the jet height, hHj =ka m/s
Num Nusselt number based on the location of maximum V voltage drop across the stainless steel foil, volt
Nusselt number, hX m =ka W width of the channel forming the jet, m
Nur Nusselt number based on the re-attachment location, Xm distance from jet exit to the location with the maximum
hX r =ka Nusselt number, m
Pr Prandtl number Xr distance from jet exit to the time-averaged reattach-
qconv convective heat flux to the flow over the foil, W/m2 ment location, m
qelec total Joulean heating, W/m2 x; y; z spatial coordinates in the stream-wise, vertical and
qrad radiation heat loss from the foil, W/m2 cross-stream directions, m
ReH Reynolds number based on the step height, U 1 Hs =m e emissivity
Rej Reynolds number based on the jet height, U j Hj =m m kinematic viscosity of air, m2/s
Rem Reynolds number based on the location of maximum r Stefan–Boltzmann constant, 5:67  108 W/m2 K4
Nusselt number, U j X m =m
Rer Reynolds number based on the re-attachment location,
U 1 X r =m

used in this investigation are outlined in the next section. The IV


q_ elec ¼ ð2Þ
measurements of the heat transfer coefficients are then reported A
and discussed.
and the local radiation heat transfer from the foil was estimated by
2. Experimental methodology  
q_ rad ¼ er T 4w  T 41 : ð3Þ
The heat transfer measurements were performed using the
facility shown in Fig. 1 used by Gao and Ewing [6]. Air from a var- The current through the foil (I) was measured using a meter on
iable speed blower flowed through the upper settling chamber the power supply, while the voltage drop across the foil (V) was
(122 cm  72.4 cm  45.7 cm) to the upper channel with a height measured using a multi-meter connected to the mounts. Here, A
(Hj ) of 3.8 cm, a length (L) of 81 cm (or L=Hj ¼ 21:3), and a width is the area of the foil between the mounts. The jet temperature
into the page (W) of 74.3 cm (or W=Hj ¼ 19:5). The exit of the (T j ) was measured using a thermocouple located in the upper set-
lower channel was sealed during these experiments. The flow exit- tling chamber. The foil temperature distribution (T w ðxÞ) was mea-
ing the upper channel attached to a wall parallel to the channel sured using an infrared camera (FLIR THERMACAM SC3000)
walls with a length of 1.8 m in the flow direction (x). The height located approximately 1 m directly above the foil. The thermal
of the wall could be adjusted to change the offset distance of the camera has a resolution of the 0.02 °C, a reported accuracy of
bottom wall of the channel (Hs ). The facility also had a back wall ±1 °C, and integration time of 3 ms. The spatial resolution of the
above the channel and side walls that extended approximately image was 0.1 mm (0:027Hj ).The top of the foil was coated with
1 m above the bottom wall. candle soot black paint with an emissivity e of 0:96  0:01, that
The heat transfer coefficient for the offset attaching jet was resulted in an uncertainty in the measured temperature of approx-
measured using the electrically heated surface shown in Fig. 1, imately 0.2 °C [14,15]. The mean foil temperature distribution was
similar to [14,15]. The heat transfer surface was a 45.7 cm by determined by averaging the temperature from 100 thermal image
10 cm (12.0Hj by 2.6Hj ) stainless steel foil mounted over a gap samples measured at 1 Hz. The standard deviations of the local
between plexiglass plates. The 0.025 mm thick foil was attached temperature fluctuations computed from these realizations were
to adjustable mounts used to tension the foil. The foil was clamped less than 0.3 °C, so the uncertainty in the mean temperature due
to the mounts with machined square aluminum bars that were to sample size should be less than 0.03 °C. The temperature
connected to a regulated DC power supply (INSTEK GW GPR- distributions were uniform from z=Hj ¼ 0:8 to 0.8 to within
1820HD). A 1.5 mm thick plexiglas plate was positioned 1.6 cm experimental uncertainty. The non-uniformity in the temperature
below the foil to form a closed cavity between the plate and the distribution near the edges was consistent with conduction to
foil. There was a 3 cm deep open cavity below this plate to further the neighbouring plates. The temperature profiles were averaged
reduce the natural convection through the bottom. The heat over the central region of the foil (from z=Hj ¼ 0:8 to 0.8) to
transfer surface was positioned along the centerline of the facility. reduce uncertainty. The radiation heat transfer was up to 14% of
Following [14,15], the local heat transfer coefficient for the jet the local electrical heating below the recirculating flow region
was computed using and up to 7% elsewhere.
The lateral heat conduction in the central region of the foil,
q_ conv q_ elec  q_ rad computed from the gradient in the surface temperature, was less
h¼ ¼ ; ð1Þ than 0.5% of the total electrical heating. The conduction heat loss
T w ðxÞ  T j T w ðxÞ  T j
through the cavities below the air foil was approximately 1% of
where the local electrical heating was estimated by the local heating, estimated using the temperature on the bottom
N. Gao et al. / International Journal of Heat and Mass Transfer 83 (2015) 21–26 23

to PC
Infrared
camera
Back Side walls Hj
wall Hs

Jet exit
y
L
457mm Bottom wall
Thermocouple 1.5mm 16mm 30mm

Foam Aluminum
to PC frame

DC power
supply

Fig. 1. Schematic of the experimental facility for the heat transfer measurements.

of the plexiglas plate. The electrical resistivity of the foil varied up 3. Results
to 2.5% due to the temperature variations. These effects and the
uncertainty in the measurements were considered to compute Typical profiles of the Nusselt number for offset jets with Rey-
the overall uncertainty using the approach outlined by Coleman nolds number of 44,400 are shown in Fig. 3. Here, the Nusselt num-
and Steele [16]. The uncertainty of the Nusselt number was esti- ber and the Reynolds number are based on the initial jet height
mated to be less than 7% for a 95% confidence interval. Full (Hj ). The Nusselt number increased along the stream-wise direc-
details of the uncertainty analysis can be found in Gao [17]. tion below much of the recirculating flow region, reaching a max-
Measurements were performed for offset jets with Hs of 0.2Hj to imum near the attachment point of the flow where the inner shear
1Hj with Reynolds number in the range 21,800 to 54,500. Velocity layer of the offset jet interacted with the wall [6]. The Nusselt num-
profiles measured near the jet exit using hot-wire anemometry ber downstream of the attachment point decreased as the attach-
showed that the offset distance and Reynolds number did not have ing jet recovers along the wall, and the structures from the inner
a significant effect on the profile on the inside of the jet near the shear layer interacted with the wall [6]. The Nusselt number pro-
exit as shown in Fig. 2. The flow averaged velocity U j for the differ- files approached the results for the planar wall jet as the flow
ent profiles was determined by integrating the curve fit to the evolved downstream. There was evidence of a minimum in the
lower half of the profile. The differences of U j =U cl computed using Nusselt number below the recirculating flow region for the larger
the lower half of the profiles and those using the full profiles [6] for offset distances similar to the results of Kim et al. [1]. This region
jets with a Reynolds number of 44,400 and offset distances of could not be resolved for the smallest offset distance examined
Hs =Hj ¼ 0.2, 0.6 and 1.0 were less than 1.5%. The centerline velocity here. The maximum Nusselt number decreased as the initial offset
(U cl ) during the experiments was measured using a pitot tube. The height of the jet increased similar to the results of Kim et al. [1].
uncertainty in the resulting Reynolds number was less than 2%. A comparison of the Nusselt number profiles in the attaching
Velocity profiles measured at different positions across the offset region is shown in Fig. 4. The stream-wise distance has been nor-
jet with a Reynolds number of 44,400 collapsed to within experi- malized by the distance to attachment location X r for each Hs =Hj
mental uncertainty for 5 6 z=Hj  5 at x=Hj ¼ 12 [18]. that was measured in the same facility using oil film visualization

0.6 0.6
Hs/Hj=0.2 Rej=21800
Hs/Hj=0.4 Rej=32500
H /H =0.6
(a) Re =44400
(b)
s j j
H /H =0.8 Re =54500
s j j
0.4 0.4
H /H =1.0
s j
(y−Hs)/Hj

j
(y−H )/H
s

0.2 0.2

0 0
0 0.2 0.4 0.6 0.8 1 1.2 0 0.2 0.4 0.6 0.8 1 1.2
U/Uj U/Uj

Fig. 2. Distributions of the stream-wise mean velocity measured at x=Hj ¼ 0:1 for (a) jets with the offset distances of Hs =Hj ¼ 0:2 to 1.0 and Rej ¼ 44; 400 and (b) jets with
Hs =Hj ¼ 0:4 and Rej ¼ 21; 800 to 54,500.
24 N. Gao et al. / International Journal of Heat and Mass Transfer 83 (2015) 21–26

300
200

160
200

120

Nu j,max
j
Nu

0.64
80 0.209 Re j
0.68
0.144 Re j
0.73
40 0.079 Re j
100
0.77
0.051 Re j
0
0 2 4 6 8 10
x/H
j
10000 50000
Fig. 3. Distribution of the Nusslet number for offset jets with Hs =Hj ¼  0.2, 4 0.4,
Rej
O 0.6,  0.8 and  1 for Rej ¼ 44; 400.
Fig. 5. Change in the maximum Nusselt number with the jet Reynolds number for
offset jets with Hs =Hj ¼  0.2, 4 0.4, O 0.6,  0.8 and  1.

on a unheated surface for the offset jets with Reynolds number of flows so the distance to the maximum heat transfer coefficient
44,400 by [6]. The uncertainty in X r was less than 5% for jets with has been normalized by the offset height here. The location of
Hs =Hj ¼ 0:2 and 3% for jets with Hs =Hj ¼ 1:0 [6]. The results show the maximum heat transfer coefficient changed with the Reynolds
that the maximum heat transfer coefficient occurred immediately number for the flows with Hs of 0:2Hj where the distance to the
upstream of the attachment point for the flows with Hs P 0:4Hj . maximum heat transfer coefficient varied from approximately
The location of the maximum moved toward the attachment loca- 6Hs to 7Hs . These results are consistent with those in the flow over
tion as the offset distance decreased and for Hs =Hj ¼ 0:2 it appears a backward facing step where X r =Hs increased with the Reynolds
to occur slightly downstream of the attachment point. The profiles number based on step height (ReH ) for ReH K 20; 000 in flows with
of the heat transfer coefficient below the recirculating flow region the same upstream boundary layer thickness [19]. The location of
and the recovery region downstream of the attachment location the maximum heat transfer coefficient did not appear to change
were similar in the flows with Hs =Hj ¼ 0:6 to 1.0. The Nusselt num- with the Reynolds number for the flows with larger initial offset
ber increased more rapidly below the recirculation region for the distances, particularly for Hs P 0:6Hj .
flows with initial offset distances of 0.2Hj and 0.4Hj and decreased Ota and Nishiyama [20] proposed that the maximum Nusselt
more slowly downstream of the attachment point, particularly in number in a range of attaching shear layer flows could be corre-
the flow with Hs of 0.2Hj . lated by
The change in the maximum Nusselt number based on the ini-
tial jet height Hj with the jet Reynolds number for the different Nur ¼ 0:192Re0:665
r Pr1=3 ; ð4Þ
flows is shown in Fig. 5. The results show the Reynolds number
dependence of the maximum Nusselt number changed with the where the Nusselt number and Reynolds number are based on the
initial offset distance of the jet Hs . The maximum Nusselt number distance to the attachment point and the free stream velocity. The
varied as Re0:64
j for the flows with Hs of 0.2Hj but as Re0:77
j for the distance to the location of the maximum heat transfer coefficient
flows with Hs of Hj . The scaling parameter appears to vary contin- was used by Ota and Nishiyama [20] in cases where the attachment
uously with the offset distance. The maximum heat transfer loca- point was not measured. The change in the maximum Nusselt num-
tion in the different flows is shown in Fig. 6. The attachment ber based on the distance to the location of the maximum heat
location was not measured for the different Reynolds number

7
200
6

150 5
s

4
X /H
Nuj

100
3

2
50
1

0 0
0 1 2 3 4 0 20000 40000 60000
x/X Re
r j

Fig. 4. Comparison of the Nusslet number below the attaching offset jets with Fig. 6. Change in the location of the maximum Nusselt number with the jet
Hs =Hj ¼  0.2, 4 0.4, O 0.6,  0.8 and  1 for Rej ¼ 44; 400. Reynolds number for offset jets with Hs =Hj ¼  0.2, 4 0.4, O 0.6,  0.8 and  1.
N. Gao et al. / International Journal of Heat and Mass Transfer 83 (2015) 21–26 25

0.08

0.07

3
10 0.06

0.05

0.77
m,max

Nuj/Rej
0.04
Nu

Ota & Nishiyama[20]

k=0.74 0.03

0.02
2
10
0.01

4 5 6 0
10 10 10 0 1 2 3 4 5 6 7 8 9 10
Rem x/Hj

Fig. 7. Change in the maximum Nusselt number based on X m with Reynolds based Fig. 9. Scaled profiles of the Nusselt number in the wall jet region for the different
on the same length for the offset jets with Hs =Hj ¼  0.2, 4 0.4, O 0.6,  0.8 and  1. offset jets with Hs =Hj ¼  0.2, 4 0.4, O 0.6,  0.8,  1 and Rej ¼ (blue) 54,500 (red)
The results for offset jet with j Hs =Hj ¼ 0:5 to 5 from Kim et al. [1] and    the 44,400 (magenta) 32,500 (black) 21,800. (For interpretation of the references to
correlation proposed by Ota and Nishiyama [20] are also included. colour in this figure legend, the reader is referred to the web version of this article.)

transfer with Reynolds number based on the same length scale for Reynolds number scaling of the Nusselt number (Nuj ) differed in
the offset jet flows is shown in Fig. 7. The correlation proposed in the regions upstream and downstream of the maximum heat trans-
[20] and the results for moderate offset distances from Kim et al. fer coefficient (Re0:49
j and Re0:56
j ) in flows with Hs =Hj ¼ 5.
[1], where the location of the maximum heat transfer coefficient The profiles of the Nusselt number downstream of the recovery
was given, are also shown. The variation of the maximum Nusselt region approached the results for the wall jet in all cases. The pro-
number with the Reynolds number still differed for the different files of the Nusselt number for the different cases collapsed in the
offset jets, but the results from the different offset jets could be well region x=Hj P 8 when they were scaled using Re0:77 j as shown in
correlated by Fig. 9. This was consistent with the flow measurements in the off-
set jets with Hs =Hj 6 1 that showed that the flow transitioned to a
Num ¼ cRe0:74
m ; ð5Þ wall jet flow in region x=Hj ¼ 6 to 10 [6]. There were differences in
where Num is the Nusselt number based on the location of the max- the location where this transition occurred for the different initial
imum heat transfer and Rem is the Reynolds number at the jet exit. offset distances. The similarity of the Nusselt number scaling of the
The Reynolds number dependence here differed from that proposed heat transfer suggests that the flows are approaching the wall jet
by Ota and Nishiyama [20] indicating the correlation proposed by flow.
Ota and Nishiyama may not be applicable in wall attaching offset
jet flows. The parameter c varied from 0.08 to 0.1 as shown in 4. Conclusions
Fig. 8. The Nusselt number below the attaching jets upstream and
immediately downstream of the maximum heat transfer point Measurements of the convective heat transfer are reported for
varied in a similar way to the maximum heat transfer point. This planar offset attaching jets exiting a long channel with Reynolds
differed from the result of Kim et al. [1] who found that the numbers of 21,800 to 54,500 and initial offset distances of 0.2 to
1.0 times the jet height. The heat transfer coefficient increased to
a maximum before decreasing as the flow developed downstream.
The location of the maximum heat transfer coefficient was in the
0.1 proximity of the attachment location for the flows with a Reynolds
number of 44,400. The maximum heat transfer occurred upstream
of the mean attachment point of the unheated surface for the jets
0.08
with initial offset distances of 0.4Hj or greater. The location shifted
0.74

downstream relative to the attachment point as the offset distance


m

0.06 decreased for this Reynolds number and appeared to occur imme-
Nu /Re

diately downstream of the attachment point for the jet with an ini-
m

tial offset distance of 0:2Hj .


0.04
The distance from the jet exit to the location of the maximum
heat transfer coefficient changed with the jet Reynolds number
0.02 in the jets with small offset distance. The variation of the Nusselt
number based on the jet height with Reynolds number in the
attaching region also changed with offset distance from Re0:64 j for
0 jets with Hs of 0:2Hj to Re0:77 for the jets with Hs of 1:0Hj . The scal-
0 0.5 1 1.5 2 2.5 3 j
ing parameter varied continuously with an increase in the initial
x/X
m offset distance. The maximum Nusselt number for the air jet flows
considered here were well correlated by Num ¼ c Re0:74 m , where the
Fig. 8. Scaled profiles of the Nusselt number in the attaching region for jets with
Hs =Hj ¼  0.2, 4 0.4, O 0.6,  0.8,  1 and Rej ¼ (blue) 54,500 (red) 44,400 (magenta)
Nusselt number and the Reynolds number are based on the dis-
32,500 (black) 21,800. (For interpretation of the references to colour in this figure tance to the location of the maximum heat transfer. The parameter
caption, the reader is referred to the web version of this article.) c varied between 0.08 and 0.1 for the jets with different offset
26 N. Gao et al. / International Journal of Heat and Mass Transfer 83 (2015) 21–26

distances. The profiles of the Nusselt number below the attaching [6] N. Gao, D. Ewing, Experimental investigation of planar offset attaching jets
with small offset distances, Exp. Fluids 42 (2007) 941–954.
jet, both upstream and downstream of the maximum heat transfer
[7] J. Eaton, J. Johnson, A review of research on subsonic turbulent flow
location, collapsed when they were scaled by the Reynolds number reattachment, AIAA J. 19 (1981) 1093–1100.
determined using the distance from the jet exit to the location of [8] J.C. Vogel, J.K. Eaton, Combined heat transfer and fluid dynamic measurements
the maximum Nusselt number. The normalized profiles in these downstream of a backward-facing step, J. Heat Transfer 107 (1985) 922–
929.
regions were similar for the jets with initial offset distances of [9] D.M. Driver, H.L. Seegmiller, J.G. Marvin, Time-dependent behavior of a
0:6Hj to 1Hj . The Nusselt number profiles approached the results reattaching shear layer, AIAA J. 25 (1987) 914–919.
for the wall jet flow in the region x=Hj P 8 and the profiles in [10] S. Jovic, D.M. Driver, Reynolds number effects on the skin friction in separated
flows behind a backward facing step, Exp. Fluids 18 (1995) 464–467.
the different jets collapsed when scaled by Re0:77
j . [11] P.G. Spazzini, G. Iuso, M. Onorato, N. Zurlo, G.M. Di-Cicca, Unsteady behavior of
back-facing step flow, Exp. Fluids 30 (2001) 551–561.
[12] J.W. Baughn, M.A. Hoffman, R.K. Takahashi, B.E. Launder, Local heat transfer
Conflict of interest
downstream of an abrupt expansion in a circular channel with constant wall
heat flux, J. Heat Transfer 106 (1984) 789–796.
None declared. [13] J.W. Baughn, M.A. Hoffman, R.K. Takahashi, D.H. Lee, Heat transfer
downstream of an abrupt expansion in the transition Reynolds number
regime, J. Heat Transfer 109 (1987) 37–42.
Acknowledgment [14] N. Gao, H. Sun, D. Ewing, Heat transfer to impinging round jets with triangular
tabs, Int. J. Heat Mass Transfer 46 (2003) 2557–2569.
This work was funded by the Natural Sciences and Engineering [15] N. Gao, D. Ewing, Investigation of the effect of confinement on the heat
transfer to round impinging jets exiting a long pipe, Int. J. Heat Fluid Flow 27
Research Council of Canada. (2006) 33–41.
[16] H.W. Coleman, W.G. Steele, Experimentation and Uncertainty Analysis for
References Engineers, second ed., John Wiley & Sons, 1999.
[17] N. Gao, Offset attaching planar jets with and without a co-flowing jet (Ph.D.
thesis), McMaster University, Canada, 2006.
[1] D. Kim, S. Yoon, D. Lee, K. Kim, Flow and heat transfer measurements of a wall
[18] N. Gao, D. Ewing, Investigation of the three dimensional flow structures in a
attaching offset jet, Int. J. Heat Mass Transfer 39 (1996) 2907–2913.
planar offset attaching jet, in: 49th AIAA Aerospace Sciences Meeting and
[2] R.A. Sawyer, The flow due to a two-dimensional jet issuing parallel to a flow
Exhibit, Orlando, USA, AIAA2011-0043, 2011.
plate, J. Fluid Mech. 9 (1960) 543–560.
[19] E.W. Adams, J.P. Johnston, Effects of the separating shear layer on the
[3] T. Lund, Augmented thrust and mass flow associated with two-dimensional jet
reattachment flow structure, Part 2: reattachment length and wall shear
reattachment, AIAA J. 24 (1986) 1964–1970.
stress, Exp. Fluids 6 (1988) 493–499.
[4] J.R.R. Pelfrey, J.A. Liburdy, Mean flow characteristics of a turbulent offset jet, J.
[20] T. Ota, H. Nishiyama, A correlation of maximum turbulent heat transfer
Fluids Eng. 108 (1986) 82–88.
coefficient in reattachment flow region, Int. J. Heat Mass Transfer 30 (1987)
[5] A. Nasr, J.C.S. Lai, Turbulent plane offset jet with small offset ratio, Exp. Fluids
1193–1200.
24 (1998) 47–57.

You might also like