Sunteți pe pagina 1din 15

The inuence of nozzle-exit geometric prole on statistical properties

of a turbulent plane jet


Ravinesh C. Deo
*
,1
, Jianchun Mi
2
, Graham J. Nathan
Turbulence Energy and Combustion [TEC] Research Group, School of Mechanical Engineering, The University of Adelaide, SA 5005, Australia
Received 19 June 2006; accepted 25 June 2007
Abstract
The paper reports an investigation of the inuence of geometric prole of a long slot nozzle on the statistical properties of a plane jet
discharging into a large space. The nozzle-exit prole was varied by changing orice-plates with dierent exit radii (r) over the range of
0 < r/h < 3.60, where h is the slot-height. The present measurements were made at a slot-height based Reynolds number (Re
h
) of
1.80 10
4
and a slot aspect ratio (span/height) of 72. The results obtained show that both the initial ow and the downstream ow
are dependent upon the ratio r/h. A top-hat mean exit velocity prole is closely approximated when r/h approaches 3.60. The decay
and spread rates of the jets mean velocity decrease asymptotically as r/h is increased, with the dierences becoming small as r/h
approaches 3.60. A decrease in r/h results in a higher formation rate of the primary vortices in the near-eld. The far-eld values of
the centerline turbulence intensity are higher for smaller r/h, and display asymptotic-like convergence as r/h approaches 3.60. Overall,
the eect of r/h on the mean and turbulence elds decreases as r/h increases.
Crown Copyright 2007 Published by Elsevier Inc. All rights reserved.
Keywords: Plane jet; Turbulence structure; Nozzle-exit geometry; Turbulence statistics
1. Introduction
Plane jets have received signicant attention after the
work of Schlichting [1], e.g. [24]. One reason is because
their two-dimensional nature oers advantages in numeri-
cal modeling, e.g. the validation of turbulence models [5].
However they also have application in heat and mass trans-
fer in air curtains [6] and ventilation and air conditioning
units [7]. In a laboratory experiment, a plane jet is pro-
duced by a rectangular slot of dimensions w h, where
w )h and two parallel plates, known as sidewalls,
attached to the slots short sides. The conguration ensures
mean jet propagation in streamwise (x) direction, spread in
the lateral (y) direction and no entrainment in the spanwise
(z) direction due to the presence of sidewalls parallel to the
xy plane. Such a conguration has been found to result in
statistical two-dimensionality over a reasonably large
downstream distance, although this depends upon the noz-
zle aspect ratio, AR w/h [8].
In most cases, smoothly contoured plane nozzles have
been used to produce a top-hat mean velocity prole
[3,4] and a laminar ow state at the nozzle exit, while some
have adopted a sharp-edged orice-plate [2,9], which pro-
duces a saddle-backed velocity prole. A conventional
sharp-edged orice-plate has an upstream-facing 45 bev-
eled edge at the nozzle exit. However, there are very few
studies using a plane jet issuing from a sharp-edged ori-
ce-plate [10], perhaps due to its initial and near-eld ow
structure being far more complex (e.g. the existence of a
vena contracta) than that from a smoothly contracting
plane nozzle. Nevertheless, due to the simplicity of its
design and manufacture, investigations on most non-planar
0894-1777/$ - see front matter Crown Copyright 2007 Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.expthermusci.2007.06.004
*
Corresponding author. Tel.: +61 8 8303 5460; fax: +61 8 8303 4367.
E-mail address: r.deo1@uq.edu.au (R.C. Deo).
1
Present address: Center for Remote Sensing and Spatial Sciences,
School of Geographical Sciences and Planning, The University of
Queensland, Brisbane 4072, Australia.
2
Present address: Department of Energy and Resource Engineering,
College of Engineering, Peking University, Beijing 100871, China.
www.elsevier.com/locate/etfs
Available online at www.sciencedirect.com
Experimental Thermal and Fluid Science 32 (2007) 545559
and non-circular jets have employed sharp-edged orice-
plates [1113].
Experimental evidence reveals that nozzles of dierent
geometry produce signicantly dierent downstream ows
[14] and the choice of each conguration depends on the
application. In addition, although the impacts of initial
conditions on downstream ow are becoming well-known
[15,16], seldom, if ever, has any study investigated plane
nozzles of dierent geometry in identical ow facilities.
To address this need, the present study aims to report the
statistical behavior of a plane jet by exploring the impacts
of varying nozzle contraction prole, as characterized by
the parameter r/h, where r is the nozzles inner-wall con-
traction radii.
Our choice of varying the nozzle-exit geometric prole is
motivatedby a number of roundand plane jet investigations,
which provide substantiating evidence that a jets down-
stream behavior is signicantly governed by upstream (exit)
conditions. While comparing the scalar mixing elds from
three types of (axisymmetric) nozzle geometries, Mi et al.
[13] found the highest mean scalar decay and highest fre-
quency of the primary vortex formation froma sharp-edged
orice-plate, followed by a smoothly contoured nozzle and
the lowest for a pipe-jet. Likewise, in another comparison
of the same three ow types, Mi and Nathan [17] concluded
that the highest velocity decay rate occurs in the ow from a
sharp-edged orice-plate, following by a smoothly con-
toured nozzle and the lowest for a pipe-jet. Other investiga-
tors, e.g. Antonia and Zhao [16], Hussain and Zedan [18]
studied smoothly contracting axisymmetric nozzles and
pipe-jets, and arrived at similar conclusions. Further infor-
mation is available fromMi et al. [19] who measured jet ows
from nine nozzle congurations, comprising of a smoothly
contoured circular, an elliptical, a triangular, a square, a
rectangular, a cross-shaped and a star-shaped exit. Their
results revealed that, relative to the circular jet, the centerline
mean velocity of non-circular jets decay more rapidly, imply-
ing anincreasedentrainment rate of the ambient uid. Hence
their investigation found a signicant inuence of nozzle
geometric prole on jets fromdierently congured nozzles.
Nomenclature
AR nozzle-exit aspect ratio, AR = w/h
d
d
boundary layer displacement thickness, d
d

_
h=2
0
1 U=U
0;c
dy
f
*
normalized vortex shedding frequency (f
*
=
fh/U
0,b
)
F
u
centerline atness factor, F
u
= hu
4
i/(hu
2
i)
2
F
max
u
maximum value of centerline atness factor
h slot-height of a plane nozzle
H shape factor of the initial velocity prole, H =
d
d
/h
m
K
u
decay rate of centerline mean velocity
K
y
jet-spreading (widening) rate
L vertical dimension of wind tunnel
M
x
axial component of jet momentum
M
0
y
lateral component of jet momentum
h
m
boundary layer momentum thickness, h
m

_
h=2
0
1 U=U
0;c
U=U
0;c
dy
r radius of contraction (plate thickness) of con-
toured nozzles
Re
h
Reynolds number based on slot-height (h),
Re
h
U
0,b
h/t
Re
h
m
Reynolds number based on momentum thick-
ness, Re
h
m
U
0;b
h
m
=m
S
u
ww centerline skewness factor, S
u
= hu
3
i/(hu
2
i)
3/2
S
min
u
minimum value of centerline skewness factor
U exit mean velocity
u uctuation component of jet velocity
u
0
root-mean-square (rms) of the velocity uctua-
tion, u
0
= hu
2
i
1/2
u
0
c;1
asymptotic value of centerline turbulence inten-
sity
u
0
p
peak uctuation component of the jet velocity
within jets shear-layers
u
0
n;c
normalized centerline turbulence intensity,
u
0
n;c
u
0
c
=U
c
, where the subscript c stands
for centerline value
u
0
out
U
out
ratio of non-uniformities in streamwise compo-
nents of velocities
U
c
centerline local mean velocity
U
0,b
exit area-averaged bulk mean velocity
U
m-a
momentum-averaged exit mean velocity (or the
characteristic normalization scale suitable for
nozzles of dierent exit velocity proles)
U
0,c
mean exit centerline velocity
U
m,c
maximum centerline velocity
U
n,c
normalized centerline mean velocity, U
n,c
=
U
c
/U
0,b
v
0
out
U
out
ratio of non-uniformities in lateral components
of velocities
w slot-span of the plane nozzle
x
01
virtual origin of the normalized mean centerline
velocity
x
02
virtual origin of the normalized velocity half-
width
x
p
length of the jets potential core
x, y, z streamwise (x), lateral (y) and spanwise (z) coor-
dinate system
y
0.5
velocity half-width, calculated at the y-location
at which Ux
1
2
U
c
x
m kinematic viscosity of the air (%1.5
10
5
m
2
s
1
at 20 C ambient temperature)
546 R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559
From measurements in a plane jet, Hussain and Clark
[20] found that an initially laminar boundary layer results
in a higher rate of change of the mass ux (spreading rate)
and the attainment of asymptotic state closer to the exit
plane. In a similar investigation using Schlieren photogra-
phy and spectral analysis, Chambers et al. [21] found more
organized and symmetric large-scale shear-layer structures
to be dominant in the initially laminar jet than in the ini-
tially turbulent one. While a higher centerline decay rate
was noted for the laminar case, the initially turbulent case
produced more three-dimensional and asymmetric struc-
tures, indicating that the nature of the initial boundary layer
controls the development of large-scale structures. Goldsch-
midt and Bradshaw [22] studied the eect of exit centerline
turbulence intensity on the ow eld of a plane jet. Impor-
tantly, they found a larger jet-spreading angle for jets of
higher exit turbulence intensity. This stands in contrast to
the spreading rate of a pipe-jet, which has higher exit turbu-
lence intensity than a smooth contraction nozzle, but a
lower spreading rate. Likewise, Hussain [23] noted some
dependence of primary vortex formation on nozzle geome-
try from ow visualization of an initially laminar plane jet.
Similarly, Russ and Strykowski [24] found that as the
boundary layer thickness at the exit plane increases, vortex
shedding and pairing occurs further downstream, leading to
a reduction in both the potential core lengths and the trans-
port of jet momentum. Likewise Eaton and Johston [25]
provided an evidence of the inuence of initial boundary
layer thickness on downstream development of free shear
ows. A more recent study conducted by Ali and Foss
[26] found that the geometric design of plane nozzles pro-
duces an inuence on the discharge properties of submerged
plane jets. Their study revealed that for plane jets having a
Reynolds numbers greater than 1500, the shape of the
downstream portion of the nozzle produces an inuence
on the entrainment rate of the plane jet by altering its pres-
sure eld at the exit plane, although they did not attempt to
quantify the extent of this inuence by using dierent noz-
zle-exit geometric proles.
From the above discussion, it is evident that the nozzle
shape inuences the initial boundary layer characteristics
to an appreciable extent, so causing signicant dierences
in the evolving ow properties of both plane and round
jets. However, no systematic study appears to have been
undertaken of the inuence of nozzle contraction prole
on a plane jet, such as by changing the parameter r/h. This
information is of fundamental and practical relevance for
the development and validation of turbulence models and
for the optimization of nozzle designs. In addition, it is use-
ful to determine what, if any, radius of contraction is nec-
essary to closely approximate a top-hat exit velocity
prole, since radially contoured nozzles are easier to
design, construct and congure than are conventional
smoothly contoured ones. With this view, we assess the
inuence of r/h on the mean ow and turbulent statistics
of a plane jet up to a downstream distance of 85h using ve
nozzles of dierent r but identical h.
2. Experiment details
The plane nozzle facility, shown schematically in Fig. 1,
consists of an open circuit wind tunnel, ow straightening
elements including a honeycomb and screens, and a
smooth contraction exit of dimensions 720 mm 340 mm.
The honeycomb is composed of drinking straws, with its
cells aligned with, and stacked perpendicular to the main
stream. This assists in reducing velocity uctuations in
the transverse direction, while producing minimum eect
on streamwise velocity because of the small pressure drop.
Likewise, the screens help reduce the velocity defect in the
turbulent boundary layer. The present screens have an
open area ratio of approximately 60% and the smooth con-
traction is based on a third order polynomial curve.
Two at plates were mounted to the end of the wind
tunnel contraction (Fig. 1b), with radially contracting
long-sides of four dierent exit radii (r) and two parallel
plates, as sidewalls, attached to the slots short sides to cre-
ate a plane nozzle. The inner-wall of all at plates was radi-
ally contoured, where the radius of contraction (r) equals
the plate thickness (see Fig. 1a). The slot-height of the noz-
zle was xed at h = 10 mm and the span (separation
between the sidewalls) was kept at w = 720 mm, producing
a large aspect ratio plane nozzle of w/h = 72.
The inner-wall radius of the nozzle plate (r) was varied
from r = 4.5 to 36 mm by a factor of 2 for each case, as
shown in Fig. 1b. This resulted in the four radially con-
toured nozzles of r/h = 0.45, 0.90, 1.80 and 3.60. During
measurements, the nozzle contraction faced upstream for
these four cases. In addition to these four nozzles, a fth
conguration of r/h % 0 was achieved by reversing the ori-
entation of the orice-plates of r/h = 0.45, so that it opens
out downstream. The symbol % is used in recognition
that this conguration is somewhat dierent from the con-
ventional sharp-edged (45 beveled) congurations more
commonly employed [2,9,11]. Since it is not possible to
achieve any conguration with an identically zero radius,
and, given the sensitivity of a ow to inlet conditions, it
is probable that subtle dierences may exist for all types
of sharp-edged orice-plates. Further, the presently cho-
sen conguration does have a sharp right-angle opposite
the contraction side of the plate, and does give consistent
(though probably not unique) trends in the data. For these
reasons, we have also chosen to use dashed lines (- - -) to
connect data points (shown later in results) between the
congurations r/h = 0.45 and r/h % 0.
The plane jet facility, located in a low noise, uid
mechanics laboratory of dimensions 18 m (long) 7 m
(wide) 2.5 m (high), was mounted horizontally, with the
plane nozzle located at the mid point between the oor
and ceiling. Throughout the present investigation, great
care was taken to ensure that the experimental facility
remained isolated from any external disturbances. The dis-
tance from the jet exit to the front wall of the laboratory
was approximately 1400h and between the jet and ceiling/
oor was approximately 125h, allowing the unheated jet
R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559 547
to discharge freely into still air. Based on the approach of
Hussain et al. [27], the eects of room connement is esti-
mated to produce less than 0.5% momentum loss for all
plane jets at a downstream distance of 85h. Hence the pres-
ent jets closely approximate plane jets in an innite envi-
ronment. For all cases of the present investigation, the
(area-averaged) jet discharge bulk mean velocity was kept
xed at U
0,b
% 27 ms
1
, resulting in a Reynolds number
based on slot-height (h) and kinematic viscosity (m) of
Re
h
% 1.80 10
4
.
The velocity measurements were performed over the
ow region 0 6 x/h 6 85 using a single hot-wire anemome-
ter, under isothermal conditions of ambient temperature
20.0 0.1 C. To avoid aerodynamic interference of the
prongs on the hot wire, the present probe was carefully
aligned horizontally and with the prongs parallel to the
Fig. 1. Schematic view of experimental setup showing (a) the wind tunnel and nozzle attachment; (b) nozzles of contraction proles denoted by (i) r/h % 0,
(ii) r/h = 0.45, (iii) r/h = 0.90, (iv) r/h = 1.80, (v) r/h = 3.60 and (c) other experimental apparatus. Note that sidewalls have been omitted for clarity, and
diagrams drawn are not to scale.
548 R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559
plane jet. Hence its output corresponds closely to the
streamwise component of the ow velocity. The single
hot-wire anemometer, if used with caution, encounters
reduced errors when compared with dual or triple wires,
in which the adjacent probe can probably inuence the
measured velocity [28]. However, a single wire is not able
to distinguish between streamwise and normal contribu-
tions to the cooling velocity thus the issue of directional
ambiguity remains to a certain degree. The hot-wire (tung-
sten) sensor was 5 lm in diameter and 0.8 mm in length,
aligned to be parallel to the long nozzle sides. The overheat
ratio of the wire was 1.5 and square wave test revealed a
maximum frequency response of 15 kHz. Hot-wire calibra-
tions were conducted using a standard Pitot-static tube,
placed side by side with the hot-wire probe, at the jets exit
(x/h % 0), where the turbulence intensity was approxi-
mately 0.5%, before and after measurements of each case
of r/h. The low initial turbulence intensity, as documented
by Stainback and Nagabushana [29], is a necessity for an
accurate calibration procedure. Both calibration functions
were tested for discrepancies, and if velocity drift exceeded
0.5%, the experiment was repeated. No further corrections
to the velocity measurements were applied. Thus it is
expected that measurements away from the centerline
(towards the outer region of the jet) are signicantly in
error, because of high velocity uctuations relative to the
mean value. Nevertheless, the central aim here is to com-
pare the measurement of one case of r/h with another, with
most data taken on the jet centerline. While converting
data points from voltages to velocities using a fourth order
polynomial curve similar to the one proposed by George
et al. [30], the average accuracy of each calibration function
was found to be 0.2%.
The signals obtained were low-pass ltered with an iden-
tical cut-o frequency of f
c
= 9.20 kHz to eliminate high
frequency noise at all the measured locations. The voltage
signals were oset to within 03 V (as a precautionary mea-
sure to avoid signal clipping [31]) and amplied appropri-
ately through the circuits, and then digitized on a
personal computer at f
s
= 18.4 kHz via a 16 channel,
12-bit PC-30F A/D converter (Fig. 1c) of signal input
range 05 V. The sampling duration was approximately
22 s, during which 400,000 instantaneous data points were
gathered. Using the inaccuracies in calibration and
observed scatter in present measurements, the uncertainties
are estimated to have a mean error of 4% at the outer
edge of the jet and 0.8% on the centerline. The errors in
the centerline mean velocity, root-mean-square (rms) veloc-
ity, skewness S
u
hu
3
i/(hu
2
i)
3/2
and atness F
u
hu
4
i/
(hu
2
i)
2
were found to be approximately 0.8, 1.8%,
2% and 1.5%, respectively. The errors in the momentum
integral quantities and jet virtual origins are estimated to
be 3%.
Although a smoothly contoured wind tunnel is used to
generate reasonably uniform ows, and to avoid signicant
ow separation upstream from the exit plane, any abrupt
change in the nozzle shape within the vicinity of the nozzle
plates, as in the present case, produces some inevitable ow
separations. This is due to unsteadiness in separations on
the curved walls, leading to uctuations in the y-compo-
nent momentum at the exit, thus altering the apping of
the jet in the xy plane. This eect, if substantial, can result
in an increase in spreading rate. One would expect this phe-
nomenon to depend to some extent on nozzle shape and in
particular on the magnitude of r. For the present measure-
ments, it is expected that the turbulence in the main con-
traction is very low; hence these separations are expected
to be not much unsteady, and thus insignicant. That this
is indeed the case is demonstrated in Appendix 1.
3. Characterization of jet exit ow
The exit ows of each plane nozzle were characterized by
measuring the velocity proles at x/h % 0.2 for r/h = 0.45
3.60 and at x/h = 1 for r/h % 0 along the lateral (y) direction
over the range 0.60 6 n 6 0.60, where n = y/h. These are
presented in Fig. 2ae. Herein, U
0,c
is the exit centerline
mean velocity. Clearly, the exit velocity proles depend on
r/h and specically undergo a substantial transition from
being saddle-backed for r/h % 0 (Fig. 2a) to closely approx-
imate a top-hat prole for r/h = 1.80 and 3.60 (Fig. 2d
and e). Although the present case for r/h % 0 cannot be clas-
sied as a conventional sharp-edged orice-plate, the sad-
dle-backed velocity proles indicate that the reversing of
the nozzle plates of smallest contraction radius (Fig. 1b
(i), r/h % 0) produce an exit ow close to that of a sharp-
edged orice-plate. It is also interesting to note that the exit
velocity proles for the cases of r/h 6 0.90 have the highest
velocity located towards the edge of the jet, resulting in the
observed saddle-back. The cases r/h 6 0.90 are found to
generate vena contracta immediately downstream from
the exit plane. Interestingly, the smaller the value of r/h,
the greater is the velocity decit on the centerline relative
to the maximum value close to edge of the jet. This trend
suggests that the nozzle conguration will approach a
sharp-edged orice-plate with a suciently small radius of
contraction. The presence of vena contractas for nozzles
of r/h 6 0.90 is also broadly consistent with a number of
previous investigations such as Quinn [11], Mi and Nathan
[12], Mi et al. [13] and Tsuchiya [32], all of which found vena
contractas in jets issuing from sharp-edged orice-plates.
The exit ow appears to be uniform within the region
jnj 6 0.45 for the cases of r/h = 1.80 and 3.60.
There is also a consistent trend in the initial turbulence
intensity proles, u
0
n
hu
2
i
1=2
=U
0;c
, with changes in r/h
(Fig. 2, Table 1). As r/h is increased from 0.45 to 3.60,
the turbulence intensity in the shear-layer decreases from
about 17% to 4%, as does that in the middle of the jet, from
u
0
n
% 2:0% for r/h = 0.45 to u
0
n
% 1:5% for r/h = 3.60. This
observation stands in contrast to round orice-plates [33],
which typically produce weaker velocity uctuations and
thus lower rms values. Such a dierence underlines the fun-
damental inuence of nozzle-exit geometry (i.e. planar ver-
sus round) on the exit velocity eld of these two jets.
R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559 549
An estimate of the initial boundary layer characteristics
is undertaken, since jet development has previously been
found to be dependent on exit conditions [18,20]. Using
Fig. 2, the exit conditions are characterized by estimating
0
0.25
0.50
0.75
1.00
1.25
-0.6 -0.3 0 0.3 0.6
0
0.05
0.10
0.15
0.20
0.25
0
0.25
0.50
0.75
1.00
1.25
-0.6 -0.3 0 0.3 0.6
0
0.05
0.10
0.15
0.20
0.25
u
'
/
U
o
,
c
u
'
/
U
o
,
c
u
'
/
U
o
,
c
u
'
/
U
o
,
c
u
'
/
U
o
,
c
-0.6 -0.3 0 0. 3 0. 6
0
0.05
0.10
0.15
0.20
0.25
-0.6 -0.3 0 0.3 0.6
0
0.05
0.10
0.15
0.20
0.25
U
/
U
o
,
c
0
0.25
0.50
0.75
1.00
1.25
U
/
U
o
,
c
0
0.25
0.50
0.75
1.00
1.25
U
/
U
o
,
c
0
0.25
0.50
0.75
1.00
1.25
U
/
U
o
,
c
U
/
U
o
,
c
= y/h
= y/h
= y/h = y/h
-0.6 -0.3 0 0.3 0.6
0
0.05
0.10
0.15
0.20
= y/h
Fig. 2. Lateral proles of the mean velocity, U/U
0,c
(denoted by the symbol - -h- -) and turbulence intensity, u
0
/U
0,c
(denoted by the symbol s) for
(i) r/h % 0 at the x/h = 1, (ii) r/h = 0.45, (iii) r/h = 0.90, (iv) r/h = 1.80 and (v) r/h = 3.60 at the x/h = 0.2.
550 R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559
the boundary layer displacement (d
d
) and momentum
(h
m
) thickness using momentum integral equations: d
d

_
yh=2
y0
1 U=U
0;c
dy and h
m

_
yh=2
y0
1 U=U
0;c
U=
U
0;c
dy (Table 1). To compensate for the comparatively
low measurement resolution and paucity of data points
through the boundary layer, a best-t spline curve was used
to perform the numerical integration of two momentum
equations, on both sides of each velocity prole (i.e. from
y = 0 to h/2 and y = h/2 to 0), yielding two independent
values of d
d
and h
m
. Both were then averaged to further
reduce errors. Note also that the case r/h % 0 has an
entirely dierent geometry, which required that the mea-
surements be conducted further downstream (x/h = 1).
This prevented a reliable estimate of its boundary layer
thickness.
Table 1 reveals that, as r/h is increased from 0.45 to 3.60,
the boundary layer thickness, d
d
increases from 0.054h to
0.151h, while h
m
increases from 0.030h to 0.061h and so
does the Reynolds number based on h
m
. The decrease in tur-
bulence intensity at x/h = 0.2 as r is increased (Fig. 2be), is
partly attributable to the apping of initial mixing layer,
which produces rms-uctuations as the mean velocity pro-
le oscillates in the y-direction further downstream from
the measurement point, although this eect is quite small.
It is expected that these variations in the exit ow, which
control the shear-layer development, have a signicant
inuence on downstream behavior of present plane jets.
This is borne out in the data presented in the next section.
The maximum cumulative error in momentum integral
quantities is 6%. This, combined with the consistent trends
evident in Table 1, gives condence in the results. However,
it must be noted that the measurements were performed at
x/h = 0.2, so that the values will dier somewhat from the
actual exit values. The corresponding shape factors
(H = d
d
/h
m
), which are often used to determine the atness
(uniformity) of the mean velocity proles [34], possess val-
ues between 1.80 and 2.48, compared with a value of 2.60
for a true Blasius exit velocity prole. Thus the present
plane nozzles of r/h = 1.80 and 3.60 may be characterized
as having an initially laminar boundary layer, since the
shape factors closely resemble those of a Blasius velocity
prole [35].
4. Statistical properties of the downstream ow
Fig. 3 shows the near-eld evolution of mean centerline
velocity, U
c
, normalized by bulk mean exit velocity, U
0,b
.
As expected, there is a consistent dependence of U
c
/U
0,b
on r/h, in particular, the case of r/h % 0 being discernibly
dierent from that for other cases. A hump in U
c
/U
0,b
(at
x/h % 2 for the cases 0 6 r/h 6 0.90) becomes obvious,
although its magnitude tends to decrease as r/h is
increased. This hump is yet another feature which unam-
biguously supports the existence of vena contractas for
nozzles of r/h 6 0.90 and is consistent with Fig. 2ac.
Fig. 3 also shows that, for x/h > 6, the decay rate of the
mean centerline velocity depends on r/h, with the jet issuing
from r/h % 0 decaying at the highest rate. This trend is con-
sistent with previous ndings of round orice-jets [13].
An assessment of the dependence of the near-eld hump
of U
c
/U
0,b
on r/h (at x/h % 2) is shown in Fig. 4. That is, as
r/h is increased from 0 to 3.60, the ratios of the maximum
centerline velocity U
m,c
to the exit bulk mean velocity U
0,b
are found to decrease asymptotically from approximately
1.30 to 1.00. Interestingly, the nozzle prole with r/h % 0
produces U
m,c
/U
0,b
% 1.30. This value is lower than
U
m,c
/U
0,b
% 1.55 found by Quinn [11] from a sharp-edged
orice-plate, although the trends are similar (Fig. 3). Taken
together, Figs. 24 provide sucient evidence to show that
the present case of r/h % 0 produces an exit and near-eld
ow structure that is qualitatively similar to other sharp-
edged orice-plate ows, with the subtle dierences attrib-
utable to the dierences in the nozzle designs.
To investigate the inuence of r/h on the primary vortex
shedding in the near-eld, we have analyzed the centerline
Table 1
Initial boundary layer characteristics obtained at x = 0.2h
r/h d
d
h
m
H = d
d
/h
m
Re
hm
U
0;b
h
m
=m u
0
c
=U
c
(%) u
0
p
=U
c
(%)
0
0.45 0.054h 0.030h 1.80 551 2.0 16.8
0.90 0.068h 0.035h 1.95 642 1.9 8.1
1.80 0.127h 0.052h 2.44 955 1.9 5.5
3.60 0.151h 0.061h 2.48 1120 1.5 3.9
0.4
0.6
0.8
1.0
1.2
1.4
1.6
0 5 10 15 20
r/h = 0
0.45
0.90
1.80
3.60
Quinn [10]
orifice-plate (rectangular)
x/h
U
c
/
U
o
,
b
Quinn [11], orifice-plate
(rectangular) nozzle
r/h 0
0.45
0.90
1.80
3.60
Fig. 3. Near-eld evolution of the normalized centerline mean velocity,
U
c
/U
0,b
for various cases of investigation.
R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559 551
velocity spectra U
u
(f
*
) within the potential core region mea-
sured at x/h = 3, a location typically known to produce
KelvinHelmholtz vortices in plane jets. Here, the normal-
ized vortex shedding frequency, f
*
fh/U
0,b
and the inte-
gral
_
U
u
f

df

1.
The spectra (Fig. 5) reveal that there are clear dierences
in the underlying ow structure of all ve jets. Each jet
exhibits a broad peak in U
u
(f
*
), revealing the periodic pas-
sage of primary vortices in the near-nozzle region. The pro-
cesses of vortex formation and growth in the near region of
plane jets are well established. For instance, it is well-
known that 2-D roller-like counter-rotating vortices domi-
nate the shear-layers which bound the potential core [36].
Thus the present spectra clearly conrm the regular occur-
rence of primary vortices from all the tested nozzles. Simi-
larly, ow visualizations of a round jet from a sharp-edged
orice-plate by Mi et al. [13] revealed well-dened coherent
vortices along their potential core region. In their smoke
visualization experiments, Tsuchiya et al. [32] noted axially
symmetric vortices within 04 nozzle widths downstream.
The mechanism leading to vortex formation immediately
downstream from the nozzle exit is a known feature [37]
as is the roll-up of the unstable laminar shear-layers to pro-
duce the primary vortices. During their streamwise propa-
gation, the vortices convect the irrotational ambient uid
into the jet. Early observations of plane jets by Brown
[38] and Beavers and Wilson [39] found that the symmetri-
cal vortices occur on alternate sides of a plane jet. Their
successive growth into larger and larger vortices through
coalescence with adjacent vortices [40] causes them to even-
tually breakdown as they propagate downstream. The pro-
cess of coalescence typically depends on the exit conditions,
as can be seen from the work of Sato [41], who found that
an externally driven noise at a frequency close to that of the
natural vortex shedding frequency, causes vortices to grow
and coalesce closer to the nozzle exit. The dependence of
vortex dynamics on initial conditions is also well-known,
e.g. [14]. Collating from the past and present work, it
becomes apparent that as the exit conditions are varied
by changing r/h from 3.60 to 0, the normalized vortex shed-
ding frequency, St
h
increases from approximately 0.24 to
0.39. Recall from Fig. 2 and Table 1 that the boundary
layer gets thicker as r/h increases. Therefore, a thinner
boundary layer, with more concentrated vorticity, results
in a higher formation rate of the primary vortices for the
case r/h % 0 [13]. As evidenced, dierent nozzle-exit-geo-
metric proles probably generate structurally dierent vor-
tices, which convect downstream at dierent rates. Also
note that the exit and near-eld centerline turbulence inten-
sity is higher for smaller values of r/h (Figs. 2 and 12 shown
later). This indicates that a higher formation rate of the pri-
mary vortices (found for smaller r/h), is associated with lar-
ger velocity uctuations. However, a higher frequency is
typically associated with a smaller scale of the vortex
motion, indicating that the higher turbulence intensity
results, at least in part, from a greater variability in the
instantaneous location of the vortex cores.
Table 2 assembles the f
*
data from previous investiga-
tions of round and plane nozzles. Also listed in Table 2
are the key initial conditions. The present measurement
of f
*
= 0.39 using r/h % 0 is in good agreement with the val-
ues f
*
= 0.43 and 0.40 measured by Beavers and Wilson
[39] and Tsuchiya et al. [32] using plane and rectangular
nozzles, respectively. This close comparison provides fur-
ther support that the geometry of r/h % 0 produces a ow
structure similar to those from other sharp-edged orice-
plates. However, all the f
*
values are signicantly higher
than a value of 0.23 measured by Sato [41] for a channel
(analogous to a pipe). This dierence reects the key role
that a nozzles geometry plays in vortex formation. Impor-
tantly, the values of 0.27 and 0.24 measured by Namar and
Otugen [37] using smoothly contracting nozzles and the
present value for r/h = 1.80 and 3.60 are in good agreement
too, conrming that they closely approximate other types
of smoothly contoured plane nozzles.
Next we assess the inuence of nozzle-exit contraction
proles on the far-eld ow of the present plane jets.
0.8
1.0
1.2
1.4
1.6
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
present data
Quinn [11]
r/h
U
m
,
c
/
U
o
,
b
Fig. 4. Dependence of the ratio of mean centerline velocity maximum,
U
m,c
and exit bulk mean velocity, U
0,b
on r/h obtained at x/h = 3.
0.00001
0.0001
0.001
0.01
0.1
0.5 1.0 1.5
(1.80, 0.24)
(r/h, f*) (3.60, 0.24)
(0.90, 0.26)
(0.45, 0.28)
(r/h, f*) (0, 0.39)
f = f h / U
o,b

u
(
f
)
*
*
Fig. 5. Power spectra, U
u
(f
*
) of the centerline velocity uctuations
measured at x/h = 3.
552 R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559
Fig. 6 presents the far-eld mean centerline velocity, U
c
,
normalized by the exit bulk mean velocity U
0,b
.
3
It is
revealed that in the self-similar region, U
c
$ x
1/2
, leading
to the well-known relationship of the form
U
0;b
U
c
_ _
2
K
u
x
h

x
01
h
_ _
1
where K
u
represents the velocity decay rate and x
01
is its
virtual origin. As with the near-eld case, the decay rates
of the far-eld mean centerline velocity reveal a consistent
dependence on r/h, with the nozzle of r/h % 0 exhibiting the
highest far-eld velocity decay. The velocity decay rates,
shown explicitly in Fig. 7, exhibit an asymptotic-like con-
vergence toward a single curve as r/h approaches 3.60.
While the dierences between the cases of r/h = 1.80 and
3.60 are within experimental uncertainty, the trend is con-
sistent, both internally and with other data presented later.
Fig. 8 presents the lateral distributions of the normalized
mean velocity at selected downstream locations for r/h % 0,
r/h = 0.45, 0.90 and 3.60. The mean velocity proles
become approximately self-similar at x/h = 20 for
r/h % 0, which is signicantly further downstream than
the equivalent x/h = 5 for r/h = 3.60. The largest distance
is required for the. case r/h % 0. That is, the downstream
distance required for the lateral proles of the mean veloc-
ity to achieve self-similarity decreases as r/h is increased.
All the self-similar proles conform closely to a Gaussian
relation, U
n
= exp[ln2(y
n
)
2
]. Likewise, the streamwise
variations (Fig. 9) of the normalized velocity half-widths,
y
0.5
/h, conform to the far-eld relationship
y
0:5
h
K
y
x
h

x
02
h
_ _
2
0.15
0.17
0.19
0.21
0.23
0.25
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
r/h
K
u
Fig. 7. The decay rates of the centerline mean velocity for r/h = 03.60.
Table 2
The normalized vortex shedding frequency, f
*
for previous jets of round, rectangular and plane congurations
Investigation Geometry Nozzle prole Re AR f
*
Beavers and Wilson [39] Plane Orice-plate 5003000 0.43
Tsuchiya et al. [32] Rectangular Orice-plate 3500 5 0.40
Sato
a
[41] Plane Channel 15008000 1067 0.23
Namar and Otugen [37] Rectangular Contoured 10007000 56 0.27
Present, r/h % 0 Planar Orice-like 18,000 72 0.39
Present, r/h = 1.80, 3.60 Planar Radially contoured 18,000 72 0.24
Beavers and Wilson [39] Round Orice-plate 5003000 0.63
Johansen [42] Round Orice-plate 2001000 0.60
Mi et al. [13] Round Orice-plate 16,000 0.70
Ko and Davis [43] Round Contoured 0.20
Crow and Champagne [44] Round Contoured 10,50030,900 0.30
Mi et al. [13] Round Contoured 16,000 0.40
a
Note: Sato [41] used a contoured planar nozzle with an upstream channel of length between 300 and 1100 mm.
0
5
10
15
20
25
0 10 20 30 40 50 60 70 80 90
r/h 0
0.45
0.90
1.80
3.60
x/h
(
U
o
,
b
/
U
c
)
2
Fig. 6. The normalized proles of centerline mean velocity for dierent
values of r/h.
3
We have also checked the normalization by the momentum-averaged
exit velocity (U
m-a
), not shown, and found the relative results between the
dierent cases being similar to those normalized by U
0,b
. Note that the
characteristic velocity
U
m-a

1
h
_
h=2
h=2
U
2
dy
_ _
1=2
since the exit momentum is M
0
qwhU
2
m-a
qw
_
h=2
h=2
U
2
dy. Obviously,
the normalization by U
m-a
takes into account the dierent mean momen-
tum ux rates with dierent exit mean velocity proles.
R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559 553
where K
y
is the spreading rate and x
02
is the virtual origin
of the half-width. Clearly, the dierent values of r/h pro-
duce dierent values of y
0.5
/h, conrming that the jet-
spreading angles dier for each nozzle geometry. Fig. 10
shows that the spreading rate decreases, approximately
asymptotically, as r/h is increased from 0 to 3.60, with
the highest spreading rate for r/h % 0. This trend, in turn,
coincides with the measured trends in the decay of U
c
(Fig. 7), and thus shows internal consistency of the present
data.
The magnitude of the virtual origins, x
01
and x
02
(each
with uncertainty of approximately 8%) is found to increase
asymptotically with r/h (Fig. 11), although with greater
scatter as expected. That is, the nozzle of r/h % 0 has the
smallest of these virtual origins, consistent with the pres-
ence of a vena contracta (Figs. 3 and 4) for this case. A
dependence of virtual origins on initial conditions was also
revealed by Gouldin et al. [5]. Importantly, Flora and
Goldschmidt [45] noted that their virtual origin moved
upstream with a relatively modest increase in exit turbu-
lence intensity from 1.06% to 1.28%. This trend again, is
consistent with the present results, where an increase in
the initial turbulence intensity from approximately 1.7%
to 2.3% is associated with a translation of the virtual ori-
gins from x
01
/h % 3.9 to x
01
/h % 0.4 and x
02
/h % 4.7 to
x
02
/h % 2.0. There is only a small dierence between the
0.03
0.05
0.07
0.09
0.11
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
r/h
K
y
Fig. 10. The jet-spreading rates for r/h = 03.60.
0
0.2
0.4
0.6
0.8
1.0
0 0.5 1.0 1.5 2.0 2.5
U
n
= exp[-ln 2 (y
n
)
2
]
y
n
= y/y
0.5
U
n
=

U

/

U
c
U
n
=

U

/

U
c
0
0.2
0.4
0.6
0.8
1.0
0 0.5 1.0 1.5 2.0 2.5
x/h = 3
5
10
20
40
80
y
n
= y/y
0.5
y
n
= y/y
0.5
y
n
= y/y
0.5
0
0.2
0.4
0.6
0.8
1.0
0 0.5 1.0 1.5 2.0 2.5
0
0.2
0.4
0.6
0.8
1.0
0 0.5 1.0 1.5 2.0 2.5
x/h = 3
5
10
20
40
80
Fig. 8. Lateral proles of the mean velocity, U/U
c
for (a) r/h % 0;
(b) r/h = 0.45; (c) r/h = 0.90 and (d) r/h = 3.60.
0
2
4
6
8
10
0 10 20 30 40 50 60 70 80 90
r/h = 0
0.45
0.90
3.60
x/h
y
0
.
5
/
h
r/h 0
0.45
0.90
1.80
3.60
Fig. 9. Streamwise evolutions of mean velocity half-width for dierent
values of r/h.
0
2
4
6
0 1 2 3 4
-5
-3
-1
1
3
5
r/h
x
0
1
/
h
x
0
2
/
h
Fig. 11. The jets virtual origins for r/h = 03.60.
554 R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559
values of x
01
for the cases of r/h = 1.80 and r/h = 3.60, con-
rming that the ows from these congurations are very
similar.
Next we consider the dependence of the turbulent veloc-
ity eld on r/h. Fig. 12 presents the streamwise evolutions
of locally normalized turbulence intensity, u
0
n;c
u
0
c
=U
c
,
for dierent values of r/h, where u
0
c
hu
2
i
1=2
. As with the
mean velocity eld, the turbulent eld exhibits a consistent
dependence on r/h. This dependence is, as expected, great-
est in near-eld but does not vanish in the far-eld. In gen-
eral, the centerline turbulence intensity decreases as r/h is
increased. The initial rapid increase of u
0
n;c
is a distinct fea-
ture of all plane jets, reecting the streamwise growth of the
shear-layer instability [46] due to the large-scale structures,
perhaps similar to those evidenced from the plane jet ow
visualizations of Gordeyev and Thomas [47] and Shlien
and Hussain [48]. It is these large-scale structures which
are responsible for large-scale engulfment of the ambient
uid, higher velocity uctuations and higher decay of mean
velocity, and thus high turbulence intensity. It is also
deduced from previous work that the far-eld ow is inu-
enced by propagation of these structures [37], and the dom-
inance of large-scale structures diminish as they convect
downstream due to the generation of a broader range of
smaller eddies.
The dierent shape in the evolution of u
0
n;c
for jets of dif-
ferent r/h implies dierences in the underlying large-scale
structures of these jets. A distinct hump in u
0
n;c
is found
at x/h % 13 for r/h % 0. A hump in turbulence intensity is
probably associated with stronger intermittent incursions
of low-velocity, predominantly ambient, uid at this loca-
tion, causing higher velocity uctuations relative to the
mean values.
In the self-similar far-eld (x/h > 30), the centerline tur-
bulence intensity clearly depends on r/h. This dependence
is highlighted in Fig. 13, which plots u
0
c;1
against r/h. Despite
some scatter, a consistent trend emerges. As r/h is increased
from 0 to 3.60, u
0
c;1
decreases in an asymptotic-like manner
with r/h. This systematic dependence indicates that the state
of fully-developed turbulent ow is dependent on the noz-
zle-exit geometric prole. Further evidence of the depen-
dence of the turbulent velocity eld on nozzle-exit
contraction prole is given by the lateral distributions of tur-
bulence intensity, u
0
n
hu
2
i
1=2
=U
c
as shown in Fig. 14ad.
There are clear dierences resulting from changes in the val-
ues of r/h. Consistent with the trends in the lateral proles of
0.1
0.2
0.3
0 10 20 30 40 50 60 70 80 90
r/h = 0
0.45
0.90
1.80
3.60
x/h
u
'
n
,
c
=

u
'
c
/

U
c
r/h 0
0.45
0.90
1.80
3.60
Fig. 12. Streamwise evolutions of locally normalized turbulence intensity,
u
0
c
=U
c
, for dierent values of r/h.
0.22
0.24
0.26
0.28
0.30
0 1 2 3 4
u
'
c
,

r/h
Fig. 13. Variations of far-eld asymptotic turbulence intensity, u
0
c;1
with
r/h.
0
0.1
0.2
0.3
0.4
0
0.1
0.2
0.3
0.4
0 0.5 1.0 1.5 2.0 2.5
x/h = 3
5
10
20
40
80
0
0.1
0.2
0.3
0.4
0 0.5 1.0 1.5 2.0 2.5
y
n
= y /y
0.5
y
n
= y /y
0.5
0 0.5 1.0 1.5 2.0 2.5 0 0.5 1.0 1.5 2.0 2.5
y
n
= y /y
0.5
y
n
= y /y
0.5
u
'
n
=

u
'
/
U
c
0
0.1
0.2
0.3
0.4
u
'
n
=

u
'
/
U
c
x/h = 3
5
10
20
40
80
Fig. 14. Lateral proles of the turbulence intensity, u
0
/U
c
for (a) r/h % 0;
(b) r/h = 0.45; (c) r/h = 0.90 and (d) r/h = 3.60.
R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559 555
the mean velocity, the axial distance at which turbulence
intensity proles become self-similar increases with r/h.
For instance, when r/h % 0, the axial distance required to
attain self-similarity of u
0
n
is at x/h = 40, whereas for
r/h = 3.60, this distance is reduced to x/h = 10. This indi-
cates that the development of the large-scale structures in
the outer shear-layers also depends on the nozzle-exit geo-
metric prole.
The dependence of the ow statistics on r/h is further
examined using the higher order moments of velocity uc-
tuations. Figs. 15 and 16 present the centerline evolutions
of the skewness S
u
and atness (kurtosis) F
u
factors of all
ve jets. These were each determined from a large sample
of approximately 400,000 data points of the instantaneous
velocity, so the convergence of the calculations is good.
Further, an appropriate voltage oset was applied to the
analogue-to-digital range ensured that no clipping
occurred [31]. The proles are here vertically oset by
unity, and the ordinate is drawn on a logarithmic scale
for clarity. Both factors evolve from nearly Gaussian val-
ues (S
u
, F
u
) = (0, 3) at the origin, to highly non-Gaussian
values, around 4 < x/h < 6, consistent with previous work.
For example, Browne et al. [36] found that their passive
temperature uctuations at the exit of a plane nozzle were
Gaussian, while those within the potential core region
(between 3 and 5h) were highly non-Gaussian. A departure
from Gaussian values is typically interpreted to result from
the presence of coherent, non-random motions due to the
growth of the large-scale roller-like structures in the
shear-layers. The near-eld trends of S
u
and F
u
, which
are governed by r/h, reect a dependence of the underlying
large-scale shear-layer structures on source (exit) condi-
tions. Also importantly, the absence of potential cores
for cases of small r/h (e.g. r/h = 0, 0.45 and 0.90) implies
the more rapid development of large-scale structures
through its shear-layer, increased uid entrainment and
quite possibly, more coherent large-scale structures.
To inspect the variations of the minima S
min
u
and max-
ima F
max
u
of both factors due to changes in r/h, we have
plotted their relative magnitudes in Fig. 17. Despite some
scatter, a clear asymptotic-like dependence of both values
on r/h is evident, with the cases r/h = 1.80 and 3.60 possess-
ing similar values. The present nozzles of small r/h pro-
duces larger values of skewness and kurtosis, indicating
that the near-eld ow encounters higher instabilities, per-
haps due to greater incursion of low-velocity ambient uid,
than with nozzles of larger r/h. In the interaction and fully-
developed regions (i.e. x/h > 20), both factors approach,
but do not reach, truly Gaussian values. The departure
of the moments of higher order statistics from their respec-
tive Gaussian values are consistent with Browne et al. [36],
whose passive scalar measurements were non-Gaussian in
the self-similar eld.
5. Conclusions
In summary, the statistical properties of the present jets
were found to depend systematically on the nozzle-exit con-
traction proles measured over the range 0 < r/h < 3.60.
The results reveal consistent dierences throughout the
1 10
-2
0
-1
1
0
0
0
0
Gaussian
x/h
S
u
=

<
u
3
>

/

(
<
u
2
>
)
3
/
2
100
Fig. 15. Streamwise evolutions of the skewness, S
u
for dierent values of
r/h. Note that each prole is shifted vertically by unity relative to its
neighbour for clarity, and symbols are identical to Fig. 12.
1 10 100
3
2
6
3
3
3
4
5
3
Gaussian
x/h
F
u
=

<
u
2
>

/

(
<
u
2
>
)
4
Fig. 16. Streamwise evolutions of the atness, F
u
for dierent values of r/h.
Note that each prole is shifted vertically by unity relative to its neighbour
for clarity, and symbols are identical to Fig. 12.
-8
-6
-4
-2
0
2
0 1 2 3 4
3
4
5
6
7
r/h
S
u m
i
n
F
u
m
a
x
Fig. 17. The dependence of the near-eld minima in skewness, S
min
u
, and
maxima in atness, F
max
u
, on r/h.
556 R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559
ow, extending from the exit velocity proles, through the
near-eld, and into the far-eld. These dierences are
deduced to result from dierences in the underlying ow
structure that propagates downstream from the dierent
nozzle-exit geometric proles. (As shown in the Appendix,
any eect of ow separations generated at the junction
between the wind tunnel contraction and the nozzle plates
is found to be very small, and would also be consistently
incorporated in all experiments.)
The exit velocity proles were found to depend system-
atically upon r/h, with a gradual transition from being sad-
dle-backed for the case r/h % 0 to closely approximate a
top-hat for the cases r/h = 1.80 and 3.60. For the cases
r/h 6 0.90, the mean exit velocity proles exhibit a sad-
dle-back, a shape which characterizes sharp-edged ori-
ce-plates. Importantly, these congurations exhibit vena
contractas, indicating upstream ow separations. The
extent of the departure of these velocity proles from a
top-hat, as characterized by the ratio of the maxima in
U
c
, to the exit bulk mean velocity, U
0,b
, decreased in an
asymptotic-like manner with r/h.
Likewise, the thickness of the initial boundary layer was
found to increase monotonically with an increase in r/h,
while its peak turbulence intensity decreased. The power
spectra of the centerline velocity uctuations revealed that
the near-eld vortex shedding frequency decreases mono-
tonically, from a value of f
*
= 0.39 for r/h % 0 to f
*
= 0.24
for r/h = 1.80 and 3.60. In the self-similar far-eld, the rates
of centerline velocity decay and jet spread were found to
decrease in an asymptotic-like manner with an increase in
r/h, so that the dierences between r/h = 1.80 and 3.60 were
small. The streamwise turbulence intensity revealed a dis-
tinct hump for the case r/h = 0 near to x/h = 13, while no
signicant hump was found for the radially contoured noz-
zles (r/h = 1.80 and 3.60). The far-eld values of turbulence
intensity also decreased in an asymptotic-like manner as r/h
was increased to 3.60. However, one would expect that a
further increase in r/h towards innity would cause the jet
properties to depart from those of a top-hat exit ow, to
converge towards a fully-developed channel ow.
The collective ndings from present work, together with
the proposed hypothesis of George [15], experimental work
of George and Davidson [49] and recent measurements of
Deo [50] conrm that the downstream development of
any plane jet is dependent upon its exit boundary (i.e. noz-
zle-exit proles) and upstream conditions. In other words,
even in the fully-developed state, a plane jet does not for-
get its origin. Therefore, the classical theory, which argues
that all jets should become asymptotically independent of
source conditions and that the jet properties will depend
only on the rate at which momentum is added and the dis-
tance from its source, is not valid for a plane jet.
Acknowledgements
The experimental work for this paper was undertaken by
R.D. at The University of Adelaide (UA), with support by
an international postgraduate funding, UA Achievers
Scholarship and an ARCLinkage Grant in partnership with
FCT-Combustion. Thanks to Dr. Peter Lanspeary for his
contribution in evaluating the role of the upstream contrac-
tion on the ow. We are grateful to Prof. W.K. George for
some valuable discussions. Finally, we would also like to
thank the reviewers for their insightful comments, which
have strengthened the paper.
Appendix 1.
The eect on the jet ow of the inevitable corner separa-
tion immediately upstream from the nozzle plate is assessed
here both by dimensional reasoning and by measurement.
We note that the shedding frequency of the corner eddies
can be expected to be lower than that of the natural shed-
ding of the jet eddies by a factor of the order (L/h)
2
, i.e.
by three orders of magnitude (here, L is the height of the
nozzle plate, Fig. 2a). This is because the natural shedding
frequency of an oscillation scales directly with the charac-
teristic length dimension, and inversely with the character-
istic velocity at their respective planes (i.e. upstream from
the contraction and at the nozzle exit).
On this basis, even though there will be small dierences
in the local Strouhal number of the two oscillations, it is
clearly impossible for the two types of oscillations to cou-
ple directly. Rather, any possible inuence would be
through the generation of a low frequency oscillation of
the entire jet. The most probable mode can be expected
to be a low frequency apping motion of the emerging
jet, which would arise were the oscillations on either side
of the contraction to be out of phase. Nevertheless, a sym-
metrical oscillation is also possible. The extent of such a
low-frequency oscillation is determined by the ratio of
the momentum of the lateral component of the oscillation,
M
0
y
, to that of the axial component of the entire jet, M
x
.
We rstly characterize the magnitude of velocity uctu-
ations in the corner eddies, u
0
and v
0
as being the same
order as the local mean velocity U. A contraction reduces
such non-uniformities in the velocity eld by the following
ratios [51]:
u
0
out
U
out

h
L
_ _
2
u
0
in
U
in
A1
and
v
0
out
U
out

h
L
_ _
1=2
v
0
in
U
in
A2
where L/h = 34 is the area contraction ratio. Hence the
streamwise components of non-uniformities are reduced
by u
0
/U % 0.0009, making it negligible, and the spanwise
component by v
0
/U % 0.17.
To obtain the momentum ux of the lateral uctuations,
we note that the extent of such uctuations can be charac-
terized by the thickness of the boundary layer. Its thickness
can be estimated from Schlictings at plate solution [1],
R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559 557
since the ow speed through the contraction is suciently
low for the boundary layers to be laminar. On this basis
U
U
0
3tanh
2
1

2
p
y
x

U
0
x
m
_
1:146
_ _
2 A3
which yields a boundary layer thickness at the edge of the
slot d
0.99
% 0.22 mm. This accords well with the measure-
ments obtained at x/h = 0.2 (Fig. 2). Assuming constant
density, the momentum ratio of the lateral uctuations in
the boundary layer relative to the axial momentum in the
jet then reduces to
M
0
y
M
x

2d
h
_ _

h
L
_
% 3:3 10
3
A4
Hence the eect of the corner oscillations on the emerging
jet ow can be expected to be negligible. Note that L/h is
the area contraction ratio of the plane nozzle.
The validity of this dimensional reasoning was veried
by measurement of the frequency spectra, since any large-
scale oscillation of the jet can be expected to be identiable
by measurement. The record length of our data samples
(22 s) is sucient to capture some 10100 of such oscilla-
tions. To assess this, the data at x/h = 0.2 were rst low-
pass ltered at 100 Hz, 30 Hz and 1 Hz, and then analyzed.
No evidence of any low frequency oscillation was found.
This conrms that the eect of the corner eddies on the
emerging ow is negligible, and any ow separation within
the corners of the anterior portion of wind tunnel has insig-
nicant eect on the emerging ow.
References
[1] H. Schlichting, Laminare strahlausbreitung, Z. Angew. Math. Mech.
13 (1933) 260263.
[2] G. Heskestad, Hot-wire measurements in a plane turbulent jet, Trans.
ASME J. Appl. Mech. 32 (1965) 721734.
[3] L.J.S. Bradbury, The structure of a self-preserving turbulent planar
jet, J. Fluid Mech. 23 (1965) 3164.
[4] E. Gutmark, I. Wygnanski, The planar turbulent jet, J. Fluid Mech.
73 (3) (1976) 465495.
[5] F.C. Gouldin, R.W. Schefer, S.C. Johnson, W. Kollmann, Non-
reacting turbulent mixing ows, Prog. Energy Combust. Sci. 12 (1986)
257303.
[6] M. Stephane, S. Camille, P. Michel, Parametric analysis of the
impinging plane air jet on a variable scaled-down model, in: Proc.
ASME/JSME Fluids Engineering Division Summer Meeting, vol. F-
227, Boston, Massachusetts, 2000.
[7] B. Moshfegh, M. Sandberg, S. Amiri, Spreading of turbulent warm/
cold plane air jet in a well-insulated room, 2004, Supported by:
University of Gavle and K K-Foundation. <http://www.hig.se/tinst/
forskning/em/spreading-of-turbulent.htm>.
[8] R.C. Deo, G.J. Nathan, J. Mi, The inuence of nozzle aspect ratio on
plane jets, Expl. Therm. Fluid Sci., 2006. http://dx.doi.org/10.1016/
j.expthermusci.2006.08.009.
[9] B.G. Van der Hegge Zijnen, Measurements of the distribution of heat
and matter in a plane turbulent jet of air, Appl. Sci. Res. A7 (1958)
277292.
[10] R.A.M. Wilson, P. V Danckwerts, Studies in turbulent mixing II. A
hot air jet, Chem. Eng. Sci. 19 (1964) 885895.
[11] W.R. Quinn, Development of a large-aspect ratio rectangular
turbulent free jet, AIAA J. 32 (3) (1994) 547554.
[12] J. Mi, G.J. Nathan, Eect of small vortex generators on scalar mixing
in the developing region of a turbulent jet, Int. J. Heat Mass Transfer
42 (1999) 39193926.
[13] J. Mi, G.J. Nathan, D.S. Nobes, Mixing characteristics of axisym-
metric free jets from a contoured nozzle, an orice plate and a pipe,
J. Fluids Eng. 123 (2001) 878883.
[14] E.J. Gutmark, F.F. Grinstein, Flow control with non-circular jets,
Ann. Rev. Fluid Mech. 31 (1999) 239272.
[15] W.K. George, The self-preservation of turbulent ows and its relation
to initial conditions, in: Recent Advances in Turbulence, Hemisphere,
New York, 1989, pp. 3973.
[16] R.A. Antonia, Q. Zhao, Eects of initial conditions on a circular jet,
Exp. Fluids 31 (2001) 319323.
[17] J. Mi, G.J. Nathan, Mean velocity decay of axisymmetric turbulent
jets with dierent initial velocity proles, in: Proc. 4th Int. Conf. on
Fluid Mechanics, Dalian, China, 2004.
[18] A.K.M.F. Hussain, M.F. Zedan, Eect of the initial conditions of the
axisymmetric free shear layer: eect of initial momentum thickness,
Phys. Fluids 21 (1978) 11001112.
[19] J. Mi, G.J. Nathan, R.E. Luxton, Centerline mixing characteristics of
jets from nine dierently shaped nozzles, Exp. Fluids 28 (2000) 9394.
[20] A.K.M.F. Hussain, A.R. Clark, Upstream Inuence on the near eld
of a planar turbulent jet, Phys. Fluids 20 (9) (1977).
[21] A.J. Chambers, R.A. Antonia, L.W.B. Browne, Eect of symmetry
and asymmetry of turbulent structures on the interaction region of a
plane jet, Exp. Fluids 3 (1985) 343348.
[22] V.W. Goldschmidt, P. Bradshaw, Eect of nozzle exit turbulence on
the spreading (or widening) rate of plane free jets, in: Joint
Engineering, Fluid Engineering and Applied Mechanics Conference,
ASME, 17, Boulder, Colarado, June 2224, 1981.
[23] A.K.M.F. Hussain, Coherent structures reality and myth, Phys.
Fluids 26 (1983) 28162850.
[24] S. Russ, P.J. Strykowski, Turbulent structure and entrainment in
heated jets: the eect of initial conditions, Phys. Fluids A 5 (12) (1993)
32163225.
[25] J.K. Eaton, J.P. Johnston, A review of research on subsonic turbulent
ow reattachment, AIAA J. 19 (9) (1981) 10931100.
[26] S.K. Ali, J.S. Foss, The discharge coecient of a planar submerged
slit-jet, ASME J. Fluids Eng. 125 (2003) 613619.
[27] H.J. Hussain, S.P. Capp, W.K. George, Velocity measurements in a
high Reynolds number momentum conserving axisymmetric turbu-
lent jet, J. Fluid Mech. 258 (1994) 3175.
[28] F.E. Jorgensen, How to measure turbulence using hot wire anemom-
eters a practical guide, Dantec Dynamics, Pub. No. 9040U6151,
Skovlunde Denmark, 2000.
[29] P.C. Stainback, K.A. Nagabushana, Review of hot-wire anemometry
techniques and the range of their applicability for various ows,
ASME Electron. J. Fluids Eng. 167 (1993) 93.
[30] W.K. George, P.D. Beuther, A. Shabbir, Polynomial calibrations for
hot wires in thermally varying ows, Exp. Therm. Fluid Sci. 2 (1989)
230235.
[31] J. Tan-Atichat, W.K. George, S. Woodward, Use of computer for
data acquisition and processing, in: A. Fuhs (Ed.), Handbook of
Fluids and Fluids Engineering, vol. 15, Wiley, NY, 1996, pp. 1098
1116.
[32] Y. Tsuchiya, C. Horikoshi, T. Sato, M. Takahashi, A study of the
spread of rectangular jets: (the shear layer near the jet exit and
visualization by the dye methods), JSME Int. J. 32 Series II (1) (1989)
1117.
[33] J. Mi, P.A.M. Kalt, G.J. Nathan, PIV measurements of a turbulent
jet from round sharp-edged plate, Exp. Fluids 42 (4) (2007) 625637.
[34] S.B. Pope, Turbulent Flows, Cambridge University Press, London,
2000, p. 303.
[35] H. Schlichting, Boundary Layer Theory, McGraw-Hill, 1968 (Chap-
ter 7).
[36] L.W.B. Browne, R.A. Antonia, S. Rajagopalan, A.J. Chambers, The
interaction region of a turbulent plane jet, J. Fluid Mech. 149 (1984)
355373.
558 R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559
[37] I. Namar, M.V. O

tu gen, Velocity measurements in a planar turbulent


air jet at moderate Reynolds numbers, Exp. Fluids 6 (1988) 387399.
[38] G.B. Brown, On vortex motion in gaseous jets and origin of their
sensitivity in sound, Proc. Phys. Soc. London 47 (1935) 703732.
[39] G.S. Beavers, T.A. Wilson, Vortex growth in jets, J. Fluid Mech. 44
(1) (1970) 97112.
[40] D.O. Rockwell, W.O. Niccolls, Natural breakdown of planar jets,
Trans. ASME J. Basic Eng. 94 (1972) 720730.
[41] H. Sato, The stability and transition of a two-dimensional jet, J. Fluid
Mech. 7 (1960) 5380.
[42] F.C. Johansen, Flow through pipe orices at low Reynolds numbers,
Proc. Roy. Soc. 126 (1929) 231245.
[43] N.W.M. Ko, P.O.A.L. Davies, The near eld within the potential
core of subsonic cold jets, J. Fluid Mech. 50 (1971) 4978.
[44] S.C. Crow, F.H. Champagne, Orderly structure in jet turbulence, J.
Fluid Mech. 48 (3) (1971) 547591.
[45] J.J. Flora, V.W. Goldschmidt, Virtual origins of a free plane
turbulent Jet, AIAA J. 7 (12) (1969) 23442446.
[46] R.A. Antonia, W.B. Browne, S. Rajagopalan, A.J. Chambers, On
organized motion of a turbulent planar jet, J. Fluid Mech. 134 (1983)
4966.
[47] S.V. Gordeyev, F.O. Thomas, Visualization of the topology of the
large-scale structure in the planar turbulent jet, in: Proc. 9th
International Symposium on Flow Visualization, 13, 2000.
[48] D.J. Shlien, A.K.M.F. Hussain, Visualization of large-scale motions
of a plane jet, ow visualization III, in: Proc. Third International
Symposium, Ann Arbor, MI, September 69, Washington DC 498-
502, 1985.
[49] W.K. George, L. Davidson, Role of initial conditions in establishing
asymptotic behavior, AIAA J. 42 (3) (2004) 438446.
[50] R.C. Deo, Experimental investigations of the inuence of Reynolds
number and boundary conditions on a plane air jet, Ph.D. Thesis,
School of Mechanical Engineering, The University of Adelaide, South
Australia 5005, 2005. Available at Australian Digital Thesis Program:
<http://thesis.library.adelaide.edu.au/public/adt-SUA20051025.054550/
index.html>.
[51] P.V. Lanspeary, Private communication, regarding the eects of a
contraction on ow separation, School of Mechanical Engineering,
The University of Adelaide, South Australia 5005, 2007.
R.C. Deo et al. / Experimental Thermal and Fluid Science 32 (2007) 545559 559

S-ar putea să vă placă și