SlideShare a Scribd company logo
1 of 12
Download to read offline
Effect of boundary layer thickness on secondary structures in a short
inlet curved duct
Jeremy Gartner 1
, Michael Amitay 2,⇑
Rensselaer Polytechnic Institute, Troy, NY 12180, USA
a r t i c l e i n f o
Article history:
Received 4 February 2014
Received in revised form 15 October 2014
Accepted 17 October 2014
Available online 13 November 2014
Keywords:
S-shape inlet
Secondary flows
Three dimensional separation
Secondary flow structures
a b s t r a c t
The flow pattern in short inlet ducts with aggressive curvature has been shown to lead, in some cases, to
an asymmetric flow field at the aerodynamic interface plane. In the present work, a two-dimensional
honeycomb mesh was added upstream of the curved duct to create a pressure drop across it, and there-
fore to an increased velocity deficit in the boundary layer. This velocity deficit led to a stronger stream-
wise separation, overcoming the instability that can result in an asymmetric flow field at the
aerodynamic interface plane. Experiments were conducted at Mach numbers of M = 0.2, 0.44 and 0.58
in an expanding aggressive duct with rectangle to a square cross section with area ratio of 1.27. Steady
and unsteady pressure measurements, together with Particle Image Velocimetry (PIV), were used to
explore the effect of the honeycomb on the symmetry of the flow field. The effect of inserting a honey-
comb was tested by increasing its height from 0 to 2.2 times the boundary layer thickness of the baseline
flow upstream of the curve. Using the honeycomb, flow symmetry was achieved for the specific geomet-
rical configuration tested with a negligible decrease of the pressure recovery.
Ó 2014 Elsevier Inc. All rights reserved.
1. Introduction
Heightened interest in short and curved inlet ducts for aircraft
has led us to further explore the flow field existing in such devices
as well as to a better knowledge of the issues associated with their
design and methods of mitigating such issues. Several factors
related to engine and aircraft performance and operation drive
the use of short inlet duct designs, such as the overall airframe
length reduction enabled by a shorter duct and reduction of frontal
planform by burying the engine into the airframe. These factors led
to a reduction in weight and fuel consumption, and allowed for
innovative external and integrated aerodynamics, such as Blended
Wing Bodies (Dagget et al., 2003). Another factor that must be
taken into consideration is the stability margin for operation of a
jet engine following the duct, where uneven pressure distribution
and secondary flow structures can lead to engine stall at the fan/
compressor stages (surge stall) (Scribben et al., 2006; Mattingly
et al., 2002).
A considerable body of work is available in the literature con-
cerning the analysis of the flow field in short inlet ducts (Bansod
and Bradshaw, 1972; Launder and Ying, 1972; Enayet et al.,
1982; Wellborn et al., 1992, 1993; Whitelaw and Yu, 1993a,b; Ng
et al., 2008, 2006). These previous research efforts have shed light
on the main features of the flow field existing in aggressively
curved ducts, where the rapid curvature in the duct results in pres-
sure gradients in the direction normal to the turn, leading to the
onset of secondary flow structures in the form of two counter
rotating vortices. Other structures were also noticed to co-exist
with these counter-rotating vortices, such as cross stream flow at
the internal surfaces, which invade the local boundary layer lead-
ing to further flow detachment (Wellborn et al., 1992, 1993; Ng
et al., 2008, 2006; Chen, 2012) disrupting the flow and creating
recirculation zones in the duct. The symmetric counter-rotating
vortices can be described by inviscid flow equations, caused solely
by the turning of the flow. These pressure driven counter-rotating
vortices convect the low momentum fluid of the boundary layer
towards the center of the duct impacting flow uniformity and pres-
sure recovery at the face of the engine located downstream, at the
aerodynamics interface plane.
Implementation of passive and active flow control techniques in
short inlet ducts has been an active field of research (Scribben
et al., 2006; Ng et al., 2008, 2006; Chen, 2012; Vaccaro, 2011;
Debronsky, 2012; Amitay et al., 2002; Gissen et al., 2011; Jirasek,
http://dx.doi.org/10.1016/j.ijheatfluidflow.2014.10.016
0142-727X/Ó 2014 Elsevier Inc. All rights reserved.
⇑ Corresponding author.
E-mail address: amitam@rpi.edu (M. Amitay).
1
Department of Mechanical, Aerospace & Nuclear Engineering, 110 8th Street,
Troy, NY, USA.
2
James L. Decker’45 Endowed Chair of Aerospace Engineering, and Director of the
Center for Flow Physics and Control (CeFPaC), 110 8th Street, Troy, NY, USA.
International Journal of Heat and Fluid Flow 50 (2014) 467–478
Contents lists available at ScienceDirect
International Journal of Heat and Fluid Flow
journal homepage: www.elsevier.com/locate/ijhff
2006; Reichert and Wendt, 1994). The predominant forms of actu-
ation are vortex generators, steady and unsteady jet blowing tan-
gent to the surface, synthetic jet actuators, and many more.
Recent work has studied multiple actuation devices (Vaccaro,
2011; Gissen et al., 2011) including combination of flow control
techniques. Although most of the previous work was focused on
circular cross section ducts (Wellborn et al., 1992, 1993;
Whitelaw and Yu, 1993a,b; Gissen et al., 2011; Jirasek, 2006),
emphasis was also given to rectangular cross section ducts
(Launder and Ying, 1972; Ng et al., 2008, 2006; Chen, 2012;
Vaccaro, 2011; Debronsky, 2012; Amitay et al., 2002). All of the
work performed with flow control had the objective of improving
the pressure recovery and pressure distribution at the exit of the
duct.
Recent experiments (Vaccaro, 2011; Debronsky, 2012) and
numerical simulations (Chen, 2012) have shown that, under some
geometrical conditions, the flow can become asymmetric, where
one of the counter-rotating vortical structure supersedes the other.
In the case of the rectangular ducts, the secondary flow structures
were shown to move towards one of the corners of the duct (Chen,
2012; Vaccaro, 2011; Debronsky, 2012).
As noted by Chen (2012), the secondary flow phenomenon
(i.e., a turbulent flow with mean streamwise vorticity) is attrib-
uted to two mechanisms: (i) the skew induced, inviscid mecha-
nism, which is caused by any bend in the flow path of ducts
with any cross sectional shape as shown by Miller (1991) (and
Fig. 1 below), and (ii) a stress-induced mechanism occurring in
any non-circular ducts, straight or not, due to anisotropy of the
Reynolds stresses. A more in-depth description of secondary flow
can be found in Perkins (1970) and Bradshaw (1987). Also note
that further complexity in the flow structures is due to swirl
development in the second bend of the s-duct. This reverse in
the curvature is accredited with the crossover of the transverse
velocity component near the side walls, an essentially inviscid
process. Another feature of short inlet ducts is the adverse pres-
sure gradient caused by the opposite curvature of the second
bend. Therefore, the secondary flow generated by the first bend
is attenuated, being reversed depending on the aggressiveness
of the turn (i.e., the aspect ratio L/D, the offset and area ratio
between the inlet and the exit sections).
Fig. 2a (taken from Wellborn et al., 1993) shows a three-dimen-
sional perspective of the owl face separation topology with a coun-
ter-rotating pair of vortices orientated with upwelling along the
centerline. The skeleton drawing of Fig. 2b shows the schematic
of a symmetric but unstable owl face separation of the first kind.
Perry and Hornung (1984) suggested that this unstable symmetric
distribution could exist due to slight variations in the flow field.
However, they stated that it was a special condition and that any
asymmetry in the flow would cause the streakline pattern to shift
to one side as shown in Fig. 2c.
Based on previous experiments (Ng et al., 2008, 2006; Vaccaro,
2011) and numerical analysis (Chen, 2012), it was shown that a
critical length to diameter ratio exists in a rectangular cross section
compact inlet duct, controlling the asymmetry of the secondary
structures. For ducts longer than this critical length the flow pat-
terns are asymptotically stable. With the reduction of duct length
below this critical value, the streamwise pressure gradient
increases and interacts with the transverse invasion. Simulta-
neously, the forward moving main flow confronts the backflow
to the streamwise separation. The symmetric pattern becomes
unstable due to the saddle–saddle connections existing in the
topology of the flow, leading to a flow bifurcation (Tobak and
Peak, 1982).
This asymmetric configuration is the starting point for the cur-
rent work, which also derives from the observation (Enayet et al.,
1982; Chen, 2012) that the secondary flow structures forming on
the inside of duct bends have their strength dependent on inlet
flow conditions, specifically the momentum thickness of the
Nomenclature
AIP aerodynamic interface plane
D width of the duct (mm)
h honeycomb height (mm)
d boundary layer thickness (mm)
h/d relative height of the honeycomb
h momentum thickness (mm)
L length of the duct (mm)
Minlet Mach number at the inlet
PR pressure recovery
Pinlet static pressure at the inlet (atm)
c specific heat ratio of air
Cp pressure coefficient
P0 total pressure (atm)
P static pressure (atm)
Fig. 1. Development of secondary flows in a pipe bend showing the (a) presence of an adverse pressure gradient, and (b) direction and orientation of the secondary flow (from
Miller, 1991).
468 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478
Fig. 2. Owl face separation of the first kind topology showing the (a) three-dimensional perspective of the separation, (b) symmetric skeleton schematic of streaklines, and (c)
skeleton schematic of asymmetric streakline pattern (from Wellborn et al., 1993).
Fig. 3. The experimental facility.
Fig. 4. Exploded view of the inlet duct assembly.
∗
Fig. 5. Inlet duct with the Instrumentation Can for AIP measurements.
Fig. 6. Pressure sensors distribution at the AIP.
J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478 469
incoming boundary layer. Therefore, the objective of the current
work is to show that proper manipulation of the incoming bound-
ary layer can lead to modification of the flow and pressure fields
without incurring any significant pressure loss. The main gain
expected from the manipulation of the incoming boundary layer
is on the cross flow symmetry of the pressure distribution at the
AIP.
Honeycomb Mesh
Quarter of a dollar
Fig. 7. Compact inlet with the honeycomb inserted.
Fig. 8. Internal view with the honeycomb inserted.
Fig. 9. Inlet total pressure field measured at M = 0.44; (a) across the duct height,
and (b) across the duct span.
Fig. 10. Cross-stream distributions of the streamwise velocity profiles upstream of
the turn at M = 0.44.
Fig. 11. Effect of the honeycomb height on the shape of the velocity profile at x/
D = À0.58.
Fig. 12. Effect of the honeycomb height on the shape of the velocity profile at x/
D = À0.14.
Table 1
Momentum displacement at x/D = À0.58.
h/d0 h (mm)
0 0.55
0.74 0.95
1.47 1.04
2.2 1.16
470 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478
2. Experimental setup
The high subsonic facility at the Center for Flow Physics and
Control (CeFPaC) at Rensselaer Polytechnic Institute was utilized
in the current experiments. The high subsonic facility is a blow
down, open return wind tunnel. The inlet duct geometry allowed
for Mach numbers up to 0.6 although most data presented was
conducted at a Mach number of 0.44 (mass flow rate up to
1.76 kg/s). Fig. 3 shows the flow path of the facility and labels each
component. The air begins in the blower and transitions through
the blower diffuser into the settling chamber. Next, it enters con-
traction followed by the inlet duct (i.e., the test section), and finally
exits the facility through a diffuser.
The blower used is a Cincinnati Fan model HP-12G29 run by a
100 HP motor that is controlled by a variable frequency drive.
The blower can produce a volumetric flow rate up to 170 m3
/
min. The air exits the blower and enters the diffuser section that
transitions the circular cross-section of the blower to the square
cross-section of the settling chamber. The air is slowed as it enters
the settling chamber where the fluid is conditioned through a set
of screens and honeycomb. In addition, a thermocouple and a static
pressure ring were monitored for all experiments and were used in
the calculation of the Mach number. The air then enters the con-
traction section with a contraction ratio of 142:1 and a conven-
tional 5th order polynomial curvature.
The air then enters the inlet duct, which has a constant cross-
section with a length of 0.3048 m for boundary layer growth as
well as to measure the inlet Mach number. Another static pressure
ring, as well as a thermocouple, was instrumented in this constant
cross-section section. Utilizing a one-dimensional isentropic flow
assumption, the inlet Mach number was found from the static
pressure ring in the inlet as well as the static pressure ring of the
settling chamber. This assumption was validated in the previous
work done by Vaccaro (2011) utilizing a total pressure Kiel probe
in the constant cross-section region of the inlet duct. The total
pressure of the inlet nearly matched the pressure measured in
the static pressure ring of the settling chamber, where the air
velocity is small enough to be assumed to be zero allowing the
pressure and temperature measured in the settling chamber to
be taken as the total quantities. From the total quantities measured
in the settling chamber and the static pressure measured in the
inlet, the Mach number from the isentropic flow assumption is:
Minlet ¼
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P0
Pinlet
 cÀ1
c
À 1
 #
Ã
2
c À 1
 
v
u
u
t ð1Þ
Following the constant area section of the inlet is the diffusing
S-shaped section. After the air goes through the inlet duct, it exits
the wind tunnel through a diffuser. This diffuser angle is 3° to
reduce flow speed as it exits into the open room.
The inlet duct has a length-to-diameter ratio of 1.6, where the
initial rectangular cross-section area of 90 mm tall by 114.3 mm
wide transitions to a square cross-section 114.3 mm by
114.3 mm resulting in an area ratio of Aexit/Ainlet of 1.27. The design
Fig. 13. Streamwise pressure coefficient distribution for the three different honeycomb actuator heights at M = 0.2.
J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478 471
was modular to allow for easy removal and exchange of parts as
can be seen in Fig. 4, which shows an exploded view of the inlet
duct. For example, two sets of windows were fabricated: one set
made of aluminum for most of the experiments and the other
out of optical grade acrylic solely for the use of PIV. This reduced
the chance of scratches occurring on the windows that would have
a negative impact on the PIV quality and potentially causing glares
and/or inaccuracies in the collected data.
The inlet duct was instrumented with 126 static pressure taps
along the lower surface of the duct. The spanwise density of static
taps increased towards the side-wall with an array on the opposite
side for checking symmetry (see Fig. 4). The static pressure taps
were sampled by the means of four pressure scanners (Scanivalve
DSA3217, 16 channels each, ±5 psid full scale and accuracy of
±0.05% of full scale, or ±0.0025% psi). Because only 64 ports could
be sampled at a time, two runs were required to sample all of
the static pressure taps.
The static pressures measured from these taps were used to cal-
culate the pressure coefficient CP at each location. The definition of
CP used for this calculation is:
CP ¼
2
cM2
inlet
P
Pinlet
À 1
 
ð2Þ
The coordinate systems is defined such the origin is at the mid-
dle of the cross-section plane at the beginning of the curvature. x is
the streamwise direction, y is the cross-stream direction (i.e., nor-
mal to the surface) and z is along the span. Another axis was
defined as y⁄
in order to quantify the boundary layers, where y⁄
is aligned with y, but is shifted such that it is defined to be zero
on the floor (see Fig. 5).
In addition to the static surface pressure sensors, an Instrumen-
tation Can was mounted at the exit of the inlet duct to acquire the
total pressure field at the aerodynamic interface plane (AIP). Fig. 5
shows the Instrumentation Can mounted to the inlet duct with
eight rakes each containing five sensors (a total of 40 sensors). Each
sensor location on the rake has a high frequency pressure trans-
ducer (Kulite model XCQ-062, ±5 psid full scale and repeatability
of ±0.5% of full scale, or ±0.025% psi) as well as a steady total pres-
sure tap that can be sampled with the four Scanivalves. The num-
bering scheme of the Kulite sensors at the AIP is shown in Fig. 6.
The perspective of Fig. 6 is the same orientation as Fig. 5 where
the observer is looking downstream toward the AIP.
The 40 Kulites at the AIP were utilized to measure the steady
and unsteady pressure. The steady pressure was used to plot con-
tours maps of the pressure recovery, which is the total pressure at
the AIP normalized by the total pressure at the inlet. The average
pressure recovery, PRavg, and the lower half pressure recovery,
PRlowerhalf, are defined respectively as
PRavg ¼
ð
P40
1 P0;iAIPÞ=40
P0inlet
ð3Þ
PRlowerhalf ¼
ð
P35
21P0;iAIPÞ=15
P0inlet
ð4Þ
Fig. 14. Streamwise pressure coefficient distribution for the three different honeycomb actuator heights at M = 0.44.
472 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478
where the Kulites are numbered as represented in Fig. 6.
Fig. 7 shows the compact inlet with the honeycomb inserted
inside it. In order to provide some proportional perspective of
the experiment’s scaling, a quarter of US dollar is situated next
to the honeycomb mesh. An additional internal view of the tunnel
is presented in Fig. 8, where the flow direction is out to the page
with the same quarter displayed in Fig. 7.
In this work, Particle Image Velocimetry (PIV) measurements
were conducted in order to collect quantitative flow measurement
of the velocity field at different streamwise and spanwise planes
within the duct. All PIV experiments conducted in this work uti-
lized a commercial LaVision System of software, and TSI hardware.
The hardware included two 120 mJ Nd:YAG lasers and a single
1000 Â 1016 pixel resolution TSI CCD camera. Additional hardware
was a Martin Magnum 850 fog machine utilized for seeding the
flow with O(1 lm) smoke particles. An array of different optics
included a cylindrical lens utilized to create the laser light sheet,
focal lens to focus the sheet within the measurement field, and
camera optics for focusing the camera at different focal lengths
and desired fields of interest. Also, the camera and the laser were
mounted on computer-controlled traverses to provide a precise
location of the measurement planes.
The velocity components were computed by the DaVis 7.1 soft-
ware using the cross-correlation technique of pairs of successive
images with 50% overlap between the interrogation windows.
The successive images were processed using a multi-pass method
in which the initial and final passes were 32 Â 32 pixels and
16 Â 16 pixels, respectively. Lastly, the averaged velocity field in
each plane was averaged over 500 instantaneous velocity fields.
The maximum velocity (approximately 150 m/s) corresponds to
an average displacement of about 4 pixels with an approximate
maximum error of ±0.1 pixel, which corresponds to an error of
±2.5% of the inflow velocity.
3. Results
The results section is divided into three sections: (i) upstream of
the curved duct, (ii) along the curved duct, and (iii) at the AIP. Both
PIV and pressure measurements were conducted and are presented
in this section.
3.1. Upstream of the curved duct
Prior to the analysis of the effect of the boundary layer thickness
on the flow field downstream, we must ascertain that the incoming
flow is uniform and symmetric. Thus, the pressure distributions
and the boundary layer velocity profiles were measured upstream
of the curved section, i.e. prior to the inlet plane of the duct (x/
D = 0).
Two techniques were utilized to quantify the incoming flow
field. The first method utilized a total pressure Kiel probe, which
was traversed across the height and span upstream of the inlet
plane at x/D = À1.56. The traversing of the probe was computer
controlled and stepped in 0.5 mm increments close to the wall,
and 2.5 mm intervals towards the center of the duct. The height
and span were normalized by the inlet width (D = 114.3 mm),
Fig. 15. Streamwise pressure coefficient distribution for the three different honeycomb actuator heights at M = 0.58.
J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478 473
and the pressure measured by the Kiel probe was normalized by
the total pressure in the settling chamber of the flow facility.
Fig. 9 shows the total pressure at Mach 0.44, which is nearly uni-
form across the height as well as the span of the duct, except close
to the duct walls, where a boundary layer exists.
The second method utilized was the PIV, which captured the
velocity flow field before the turn (x/D = À0.58) at three spanwise
locations (z/D = 0, ±0.25) and enabled the extraction of the bound-
ary layer profiles. As presented in Fig. 10, the flow field upstream of
the turn is two dimensional.
As was mentioned above, the main goal of the present work is
to change the weighted contribution of the flow separation to
affect the secondary flow structures. Thus, a honeycomb was
inserted upstream of the turn at various heights (into the flow)
to create a pressure drop across it which generated a velocity def-
icit. As the honeycomb was inserted deeper into the flow (i.e.,
increase h/d0) the velocity deficit of the boundary layer increased,
which made it more susceptive to separation. Fig. 11 shows the
velocity profiles at a streamwise location of x/D = À0.58. The
velocity was normalized by the freestream velocity, U1, and the
height was normalized by the boundary layer thickness
(d0 = 0.99U1) of the baseline. The value of d0 is defined at x/
D = À0.58 for M = 0.44 and is kept the same throughout the paper
for all the different Mach numbers. Similar trends were seen for
M = 0.2 and 0.58 but not shown here for brevity.
The velocity profiles were also measured at a streamwise loca-
tion of x/D = À0.14 and are shown in Fig. 12. The same trend seen
upstream is also present at this streamwise location, except that
the flow accelerates due to the proximity to the turn.
In order to quantify the results presented in Fig. 11 and to show
the momentum deficit, which increases due to the insertion of the
honeycomb, the momentum thickness, h, is presented in Table 1
for each velocity profile.
3.2. Along the curved surface
Next, the distributions of the pressure coefficient, Cp, along the
curved surface are presented at the three Mach numbers of 0.2,
0.44 and 0.58 (Figs. 13–15, respectively), at three spanwise loca-
tions of z/D = 0, ±0.25 for four different honeycomb heights. These
plots are divided into three regions:
 Region 1, where the flow accelerates around the turn and three
dimensionalities develop for the baseline case. However, when
the honeycomb is introduced into the flow it results in quasi
two-dimensional flow acceleration, where the asymmetry
decreases as h/d0 increases.
 Region 2 represents the region where the baseline flow is sep-
arated as indicated by the constant pressure. The presence of
the honeycomb affects the separation region by making it
more two-dimensional and extends it farther downstream, as
is expected since the velocity deficit of the incoming boundary
layer was increased (as discussed above).
 Region 3, where without the honeycomb, the flow on left side of
the duct (z/D = À0.25) reattaches first. By increasing the honey-
comb height, the difference in the pressure coefficient between
z/D = À0.25 to the other two spanwise locations (z/D = 0 and
0.25) decreases. This means that the reattachment point is far-
Fig. 16. Pressure recovery at the AIP for the baseline and the three different honeycomb heights at M = 0.2.
474 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478
Fig. 17. Pressure recovery at the AIP for the baseline and the three different honeycomb heights at M = 0.44.
Fig. 18. Pressure recovery at the AIP for the baseline and the three different honeycomb heights at M = 0.58.
J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478 475
ther downstream with the honeycomb than without it. There-
fore, elevating the honeycomb increases the extent of the sepa-
ration region.
3.3. At the AIP
Next, the pressure recovery at the AIP was measured and is
discussed in this section. Figs. 16–18 present the time-averaged
pressure recovery at three Mach numbers and four honeycomb
heights. The red region represents a pressure recovery of 1,
meaning that there are no pressure losses, and the blue region
represents pressure losses. As can be seen, without the honey-
comb, the asymmetry, which is a result of the superposition of
the separation and the secondary flow structures, is clearly pres-
ent for all three Mach numbers. Here, the lowest pressure recov-
ery is on the bottom right side. However, increasing the
honeycomb height decreases the asymmetry, with minimal effect
of the overall pressure recovery.
Table 2 presents the pressure recovery values averaged over the
entire AIP plane, PRave, and over only the lower half of the AIP
plane, PRlowerhalf, for the different honeycomb heights. An important
result from this study is the low impact of the honeycomb mesh on
the pressure recovery while improving the symmetry of the pres-
sure distribution. By looking at the values of Table 2, the largest
impact on the pressure recovery is at M = 0.58 between the base-
line case (without honeycomb) and the case with the honeycomb
inserted at h/d0 = 2.2, where the difference is 0.68%.
In addition to the time-averaged pressure recovery, the effect of
the honeycomb on the fluctuating total pressure was also explored.
Fig. 19 presents the standard deviation of the pressure recovery for
the same cases showed in Fig. 18. For the baseline cases (i.e., with-
out the honeycomb present) at all three Mach numbers, there is a
large concentration of pressure fluctuations in the lower portion of
the AIP, which is the region where the PR is the lowest. When the
honeycomb is inserted into the duct (where h/d0 = 2.2), the fluctu-
ating pressure fields show symmetric distribution of the pressure
fluctuations field.
In addition to a global evaluation of the fluctuating pressure
field, more detailed examination was conducted where the power
spectra at three spanwise locations (sensors 23, 28, 33) at the AIP
were calculated and are presented in Fig. 20. For all three Mach
numbers, without the honeycomb in the flow, the power spectra
at the three spanwise locations are different, whereas the spectral
content on the left side and the middle of the AIP (sensors 23 and
28, the red and green lines, respectively) contains higher energy
than that on the right side (sensor 33, blue line). When the hon-
eycomb is inserted into the flow there is a negligible change on
the energy content in the left side of the AIP, meanwhile the
energy on the right side increases considerably. This modification
in the amplitude of the power spectra further suggests that the
introduction of the honeycomb induce a more symmetrical flow
field (both the time average and the fluctuating). Also, the mag-
nitude of the peak, which represents the dominant shedding fre-
quency of the separated flow, increases with the honeycomb
inserted into the flow, corroborating the results shown previ-
ously. The increase in amplitude, especially at the shedding
energy, is another indication that the separation becomes more
dominant and by doing so helps push the saddle–saddle point
farther downstream in the duct. This saddle–saddle point, as
mentioned previously, causes the onset of the instability respon-
sible for the asymmetry encountered in the flow without the
honeycomb actuator.
Table 2
Pressure recovery at the AIP.
h/d0 M = 0.2 M = 0.4 M = 0.58
PRave PRlower half PRave PRlower half PRave PRlower half
0 0.98904 0.98185 0.94934 0.91631 0.92183 0.87502
0.74 0.98916 0.98217 0.94905 0.91563 0.92115 0.87326
1.47 0.98879 0.98126 0.94690 0.91068 0.91731 0.86505
2.2 0.98857 0.98070 0.94517 0.90656 0.91560 0.86134
Fig. 19. Standard deviation of the pressure recovery at the AIP at M = 0.2, 0.44 and 0.58. Upper row: baseline flow (h/d0 = 0), and lower row: the most symmetrical
configuration achieved in the current experiments (h/d0 = 2.2).
476 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478
4. Conclusion
The present work showed that, under some conditions, an
asymmetric flow field could develop inside a short inlet duct. In
order to correct this asymmetry, a two-dimensional honeycomb
mesh was inserted upstream of the curved duct to increase the
velocity deficit of the incoming boundary layer. It was shown
experimentally, using steady and unsteady pressure measure-
ments and PIV, that the manipulations of the velocity deficit led
to a more promptly flow separation, which, as a result, modified
the secondary flow structures and corrected the flow asymmetry
at the AIP. The experiments were conducted at Mach numbers of
0.2, 0.44 and 0.58, and the effect of inserting a honeycomb was
tested by increasing its penetration into the flow from 0 to 2.2
times the boundary layer thickness of the baseline flow upstream
of the curve. Using the honeycomb, flow symmetry was achieved
for the specific geometrical configuration tested with a negligible
decrease of the pressure recovery at the AIP. Moreover, the span-
wise uniformity of the flow increased as the height of the honey-
comb was increased.
Acknowledgements
This work was funded by the Northrop Grumman Corporation
(monitored by Ms. Florine Cannelle). The authors would also like
to thank the help of Dr. Israel Salvador and Mr. Brian Debronsky.
References
Amitay, M., Pitt, D., Glezer, A., 2002. Separation control in duct flows. J. Aircraft 39
(4), 616–620.
Bansod, P., Bradshaw, P., 1972. The flow in S-shaped ducts. Aeronautical Quarterly
23, 131–140.
Bradshaw, P., 1987. Turbulent secondary flows. Annu. Rev. Fluid Mech. 19, 53–
74.
Chen, Y., 2012. Subsonic Flows through S-Ducts with Flow Control. Ph.D. Thesis.
Department of Mechanical, Aerospace  Nuclear Engineering, Rensselaer
Polytechnic Institute, Troy, NY.
Dagget, D.L., Kawai, R., Friedman, D., 2003. Blended Wing Body Systems Studies:
Boundary Layer Ingestion Inlets with Active Flow Control. Technical Report
NASA/CR-2003-212670, Hampton, VA.
Debronsky, B., 2012. Experimental Investigation of Flow Control Techniques in a
Compact Inlet Duct. M.S. Dissertation. Department of Mechanical, Aerospace 
Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, NY.
Enayet, M.M., Gibson, M.M., Taylor, A.P., Yianneskis, M., 1982. Laser-Doppler
measurements of laminar and turbulent flow in a pipe bend. Int. J. Heat Fluid
Flow 3 (4), 213–219.
Gissen, A.N., Vukasinovic, B., McMillan, M.L., Glezer, A., 2011. Dynamics of Hybrid
Flow Control in a Boundary-Layer-Ingesting Offset Diffuser. AIAA Paper 2011-
3096, 41st AIAA Fluid Dynamics Conference and Exhibit, Honolulu, HI.
Jirasek, A., 2006. Design of vortex generator flow control in inlets. J. Aircraft 43 (6),
1886–1892.
Launder, B.E., Ying, W.M., 1972. Secondary flows in ducts of square cross-section. J.
Fluid Mech. 54 (Part 2), 289–295.
Mattingly, J.D., Heiser, W.H., Preatt, D.T., 2002. Aircraft Engine Design. AIAA
Educational Series.
Miller, D.S., 1991. Internal Flow Systems, 1990. BHRA, Cranfied, Bedford, UK.
Ng, Y.T., Luo, S.C., Lim, T.T., Ho, Q.W., 2006. On swirl development in a square cross-
sectioned, S-shaped duct. Exp. Fluids 41 (6), 975–989.
Ng, Y.T., Luo, S.C., Lim, T.T., Ho, Q.W., 2008. On the relation between centrifugal force
and radial pressure gradient in flow inside curved and S-shaped ducts. Phys.
Fluids 20 (5), 055109–055109-15.
Perkins, H.J., 1970. The formation of streamwise vorticity in turbulent flow. J. Fluid
Mech. 44 (4), 721–740.
Perry, A.E., Hornung, H., 1984. Some aspects of three-dimensional separation. II-
Vortex skeletons. Zeitshrift fuer Flugwissenschaften und Weltraumforschung 8,
155–160.
Reichert, B.A., Wendt, B.J., 1994. Improving Diffusing S-Duct Performance by
Secondary Flow Control. NASA Technical Memorandum 106492, also AIAA 32nd
Aerospace Sciences Meeting and Exhibit, Reno, NV.
Scribben, A.C., Ng, W., Burdisso, R., 2006. Effectiveness of a serpentine inlet duct
flow control technique at design and off-design simulated flight conditions. J.
Turbomach. 128, 332–339.
Tobak, M., Peak, D.J., 1982. Topology of three-dimensional separated flows. Annu.
Rev. Fluid Mech. 14, 61–85.
Vaccaro, J.C., 2011. Flow Control in a Compact Inlet. Ph.D. Thesis. Department of
Mechanical, Aerospace  Nuclear Engineering, Rensselaer Polytechnic Institute,
Troy, NY.
Wellborn, S., Reichert, B., Okiishi, T., 1992. An experimental investigation of the flow
in a diffusing S-duct. NASA Technical Memorandum 105809.
Fig. 20. Power spectra on the bottom part of the AIP at three spanwise locations and for the three Mach numbers. Upper row: h/d0 = 2.2, and lower row: h/d0 = 2.2.
J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478 477
Wellborn, S., Okiishi, T., Reichert, B., 1993. A study of the compressible flow through
a diffusing S-duct. NASA Technical Memorandum 106411.
Whitelaw, J.H., Yu, S.C.M., 1993a. Velocity measurements in an S-shaped diffusing
duct. Exp. Fluids 15 (4–5), 364–367.
Whitelaw, J.H., Yu, S.C.M., 1993b. Turbulent flow characteristics in an S-shaped
diffusing duct. Flow Meas. Instrum. 4 (3), 171–179.
478 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478

More Related Content

What's hot

A Computational Analysis of Flow Development through a Constant Area C- Duct
A Computational Analysis of Flow Development through a Constant Area C- DuctA Computational Analysis of Flow Development through a Constant Area C- Duct
A Computational Analysis of Flow Development through a Constant Area C- DuctIJERA Editor
 
unsteady flow simulation along staggered cylinder arrangement
unsteady flow simulation along staggered cylinder arrangementunsteady flow simulation along staggered cylinder arrangement
unsteady flow simulation along staggered cylinder arrangementRamakant Pandey
 
Drag Optimization of Bluff Bodies using CFD for Aerodynamic Applications
Drag Optimization of Bluff Bodies using CFD for Aerodynamic ApplicationsDrag Optimization of Bluff Bodies using CFD for Aerodynamic Applications
Drag Optimization of Bluff Bodies using CFD for Aerodynamic Applicationsijceronline
 
Large eddy simulation of the flow over a circular cylinder at high reynolds n...
Large eddy simulation of the flow over a circular cylinder at high reynolds n...Large eddy simulation of the flow over a circular cylinder at high reynolds n...
Large eddy simulation of the flow over a circular cylinder at high reynolds n...Jesús Martínez
 
Bdp shear aci
Bdp shear aciBdp shear aci
Bdp shear aciOscar A
 
A study on dynamic analysis of a dam intake tower and foundation
A study on dynamic analysis of a dam intake tower and foundationA study on dynamic analysis of a dam intake tower and foundation
A study on dynamic analysis of a dam intake tower and foundationeSAT Journals
 
Control of Suddenly Expanded Flow at Low Supersonic Mach Numbers
Control of Suddenly Expanded Flow at Low Supersonic Mach NumbersControl of Suddenly Expanded Flow at Low Supersonic Mach Numbers
Control of Suddenly Expanded Flow at Low Supersonic Mach NumbersIJERA Editor
 
Open channel flow velocity profiles for different reynolds numbers and roughn...
Open channel flow velocity profiles for different reynolds numbers and roughn...Open channel flow velocity profiles for different reynolds numbers and roughn...
Open channel flow velocity profiles for different reynolds numbers and roughn...eSAT Journals
 
Fluid and structural modeling of cavitating propeller flows
Fluid and structural modeling of cavitating propeller flowsFluid and structural modeling of cavitating propeller flows
Fluid and structural modeling of cavitating propeller flowsPhuong Dx
 
Open channel flow velocity profiles for different
Open channel flow velocity profiles for differentOpen channel flow velocity profiles for different
Open channel flow velocity profiles for differenteSAT Publishing House
 
Prefabricated vertical drains
Prefabricated vertical drainsPrefabricated vertical drains
Prefabricated vertical drainsMadhuReddy163523
 
Analysis of Stress Distribution in a Curved Structure Using Photoelastic and ...
Analysis of Stress Distribution in a Curved Structure Using Photoelastic and ...Analysis of Stress Distribution in a Curved Structure Using Photoelastic and ...
Analysis of Stress Distribution in a Curved Structure Using Photoelastic and ...IOSR Journals
 
Estimation of Damping Derivative of a Delta Wing with Half Sine Wave Curved L...
Estimation of Damping Derivative of a Delta Wing with Half Sine Wave Curved L...Estimation of Damping Derivative of a Delta Wing with Half Sine Wave Curved L...
Estimation of Damping Derivative of a Delta Wing with Half Sine Wave Curved L...IOSR Journals
 

What's hot (20)

GT2011-45427
GT2011-45427GT2011-45427
GT2011-45427
 
A Computational Analysis of Flow Development through a Constant Area C- Duct
A Computational Analysis of Flow Development through a Constant Area C- DuctA Computational Analysis of Flow Development through a Constant Area C- Duct
A Computational Analysis of Flow Development through a Constant Area C- Duct
 
unsteady flow simulation along staggered cylinder arrangement
unsteady flow simulation along staggered cylinder arrangementunsteady flow simulation along staggered cylinder arrangement
unsteady flow simulation along staggered cylinder arrangement
 
Drag Optimization of Bluff Bodies using CFD for Aerodynamic Applications
Drag Optimization of Bluff Bodies using CFD for Aerodynamic ApplicationsDrag Optimization of Bluff Bodies using CFD for Aerodynamic Applications
Drag Optimization of Bluff Bodies using CFD for Aerodynamic Applications
 
D42051929
D42051929D42051929
D42051929
 
Large eddy simulation of the flow over a circular cylinder at high reynolds n...
Large eddy simulation of the flow over a circular cylinder at high reynolds n...Large eddy simulation of the flow over a circular cylinder at high reynolds n...
Large eddy simulation of the flow over a circular cylinder at high reynolds n...
 
Bdp shear aci
Bdp shear aciBdp shear aci
Bdp shear aci
 
Lg3420362048
Lg3420362048Lg3420362048
Lg3420362048
 
A study on dynamic analysis of a dam intake tower and foundation
A study on dynamic analysis of a dam intake tower and foundationA study on dynamic analysis of a dam intake tower and foundation
A study on dynamic analysis of a dam intake tower and foundation
 
Control of Suddenly Expanded Flow at Low Supersonic Mach Numbers
Control of Suddenly Expanded Flow at Low Supersonic Mach NumbersControl of Suddenly Expanded Flow at Low Supersonic Mach Numbers
Control of Suddenly Expanded Flow at Low Supersonic Mach Numbers
 
I05714349
I05714349I05714349
I05714349
 
In situ stress
In situ stressIn situ stress
In situ stress
 
Open channel flow velocity profiles for different reynolds numbers and roughn...
Open channel flow velocity profiles for different reynolds numbers and roughn...Open channel flow velocity profiles for different reynolds numbers and roughn...
Open channel flow velocity profiles for different reynolds numbers and roughn...
 
Fluid and structural modeling of cavitating propeller flows
Fluid and structural modeling of cavitating propeller flowsFluid and structural modeling of cavitating propeller flows
Fluid and structural modeling of cavitating propeller flows
 
Open channel flow velocity profiles for different
Open channel flow velocity profiles for differentOpen channel flow velocity profiles for different
Open channel flow velocity profiles for different
 
Ijetcas14 473
Ijetcas14 473Ijetcas14 473
Ijetcas14 473
 
Prefabricated vertical drains
Prefabricated vertical drainsPrefabricated vertical drains
Prefabricated vertical drains
 
ijhff2
ijhff2ijhff2
ijhff2
 
Analysis of Stress Distribution in a Curved Structure Using Photoelastic and ...
Analysis of Stress Distribution in a Curved Structure Using Photoelastic and ...Analysis of Stress Distribution in a Curved Structure Using Photoelastic and ...
Analysis of Stress Distribution in a Curved Structure Using Photoelastic and ...
 
Estimation of Damping Derivative of a Delta Wing with Half Sine Wave Curved L...
Estimation of Damping Derivative of a Delta Wing with Half Sine Wave Curved L...Estimation of Damping Derivative of a Delta Wing with Half Sine Wave Curved L...
Estimation of Damping Derivative of a Delta Wing with Half Sine Wave Curved L...
 

Viewers also liked

Claim Form HB Full Set
Claim Form HB Full SetClaim Form HB Full Set
Claim Form HB Full SetDiyana Arus
 
Personal Statement Form PAMB
Personal Statement Form PAMBPersonal Statement Form PAMB
Personal Statement Form PAMBDiyana Arus
 
AES CAPABILITIES Brochure
AES CAPABILITIES BrochureAES CAPABILITIES Brochure
AES CAPABILITIES BrochureJason Wood
 
MSA Corporate Overview Linked In
MSA Corporate Overview Linked InMSA Corporate Overview Linked In
MSA Corporate Overview Linked InJuan Quesada, QSSP
 
Advanced Fluid Mechanics
Advanced Fluid MechanicsAdvanced Fluid Mechanics
Advanced Fluid Mechanicsguest36d460
 
Asthma Questioinnaire
Asthma QuestioinnaireAsthma Questioinnaire
Asthma QuestioinnaireDiyana Arus
 
Personal Medical Attendant's Report
Personal Medical Attendant's ReportPersonal Medical Attendant's Report
Personal Medical Attendant's ReportDiyana Arus
 
Fluid mechanics - Motion of Fluid Particles and Stream
Fluid mechanics - Motion of Fluid Particles and StreamFluid mechanics - Motion of Fluid Particles and Stream
Fluid mechanics - Motion of Fluid Particles and StreamViraj Patel
 
Flow Control in a Diffuser at Transonic Conditions
Flow Control in a Diffuser at Transonic ConditionsFlow Control in a Diffuser at Transonic Conditions
Flow Control in a Diffuser at Transonic ConditionsJeremy Gartner
 
Separation of boundary layer
Separation of boundary layerSeparation of boundary layer
Separation of boundary layerYash Patel
 
PCEIA English Version
PCEIA English VersionPCEIA English Version
PCEIA English VersionDiyana Arus
 
PCEIA BM Version for Insurance Exam
PCEIA BM Version for Insurance ExamPCEIA BM Version for Insurance Exam
PCEIA BM Version for Insurance ExamDiyana Arus
 
A STUDY ON VISCOUS FLOW (With A Special Focus On Boundary Layer And Its Effects)
A STUDY ON VISCOUS FLOW (With A Special Focus On Boundary Layer And Its Effects)A STUDY ON VISCOUS FLOW (With A Special Focus On Boundary Layer And Its Effects)
A STUDY ON VISCOUS FLOW (With A Special Focus On Boundary Layer And Its Effects)Rajibul Alam
 
Boundary layer concept for external flow
Boundary layer concept for external flowBoundary layer concept for external flow
Boundary layer concept for external flowManobalaa R
 
Boundary layer equation
Boundary layer equationBoundary layer equation
Boundary layer equationJustin Myint
 
Boundary Layer Displacement Thickness & Momentum Thickness
Boundary Layer Displacement Thickness & Momentum ThicknessBoundary Layer Displacement Thickness & Momentum Thickness
Boundary Layer Displacement Thickness & Momentum ThicknessHaroon Rashid
 

Viewers also liked (19)

Claim Form HB Full Set
Claim Form HB Full SetClaim Form HB Full Set
Claim Form HB Full Set
 
Personal Statement Form PAMB
Personal Statement Form PAMBPersonal Statement Form PAMB
Personal Statement Form PAMB
 
AES CAPABILITIES Brochure
AES CAPABILITIES BrochureAES CAPABILITIES Brochure
AES CAPABILITIES Brochure
 
MSA Corporate Overview Linked In
MSA Corporate Overview Linked InMSA Corporate Overview Linked In
MSA Corporate Overview Linked In
 
Advanced Fluid Mechanics
Advanced Fluid MechanicsAdvanced Fluid Mechanics
Advanced Fluid Mechanics
 
Asthma Questioinnaire
Asthma QuestioinnaireAsthma Questioinnaire
Asthma Questioinnaire
 
Personal Medical Attendant's Report
Personal Medical Attendant's ReportPersonal Medical Attendant's Report
Personal Medical Attendant's Report
 
Fluid mechanics - Motion of Fluid Particles and Stream
Fluid mechanics - Motion of Fluid Particles and StreamFluid mechanics - Motion of Fluid Particles and Stream
Fluid mechanics - Motion of Fluid Particles and Stream
 
Flow Control in a Diffuser at Transonic Conditions
Flow Control in a Diffuser at Transonic ConditionsFlow Control in a Diffuser at Transonic Conditions
Flow Control in a Diffuser at Transonic Conditions
 
Separation of boundary layer
Separation of boundary layerSeparation of boundary layer
Separation of boundary layer
 
PCEIA English Version
PCEIA English VersionPCEIA English Version
PCEIA English Version
 
Java EE Pattern: The Boundary Layer
Java EE Pattern: The Boundary LayerJava EE Pattern: The Boundary Layer
Java EE Pattern: The Boundary Layer
 
PCEIA BM Version for Insurance Exam
PCEIA BM Version for Insurance ExamPCEIA BM Version for Insurance Exam
PCEIA BM Version for Insurance Exam
 
A STUDY ON VISCOUS FLOW (With A Special Focus On Boundary Layer And Its Effects)
A STUDY ON VISCOUS FLOW (With A Special Focus On Boundary Layer And Its Effects)A STUDY ON VISCOUS FLOW (With A Special Focus On Boundary Layer And Its Effects)
A STUDY ON VISCOUS FLOW (With A Special Focus On Boundary Layer And Its Effects)
 
Bl concept
Bl conceptBl concept
Bl concept
 
Lecture 7
Lecture 7Lecture 7
Lecture 7
 
Boundary layer concept for external flow
Boundary layer concept for external flowBoundary layer concept for external flow
Boundary layer concept for external flow
 
Boundary layer equation
Boundary layer equationBoundary layer equation
Boundary layer equation
 
Boundary Layer Displacement Thickness & Momentum Thickness
Boundary Layer Displacement Thickness & Momentum ThicknessBoundary Layer Displacement Thickness & Momentum Thickness
Boundary Layer Displacement Thickness & Momentum Thickness
 

Similar to Effect of boundary layer thickness on secondary structures in a short inlet curved duct

Flow Investigation inside A Curved Square Duct
Flow Investigation inside A Curved Square DuctFlow Investigation inside A Curved Square Duct
Flow Investigation inside A Curved Square DuctIJERA Editor
 
Numerical and experimental investigation of co shedding vortex generated by t...
Numerical and experimental investigation of co shedding vortex generated by t...Numerical and experimental investigation of co shedding vortex generated by t...
Numerical and experimental investigation of co shedding vortex generated by t...Alexander Decker
 
CFD simulation of Lid driven cavity flow
CFD simulation of Lid driven cavity flowCFD simulation of Lid driven cavity flow
CFD simulation of Lid driven cavity flowIJSRD
 
Calculation of Fluid Dynamic for Wind Flow around Reinforced Concrete Walls
Calculation of Fluid Dynamic for Wind Flow around Reinforced Concrete WallsCalculation of Fluid Dynamic for Wind Flow around Reinforced Concrete Walls
Calculation of Fluid Dynamic for Wind Flow around Reinforced Concrete WallsIJERA Editor
 
IRJET- Study of Jet Impingement Heat Transfer
IRJET-  	  Study of Jet Impingement Heat TransferIRJET-  	  Study of Jet Impingement Heat Transfer
IRJET- Study of Jet Impingement Heat TransferIRJET Journal
 
Turbulent Flow in Curved Square Duct: Prediction of Fluid flow and Heat trans...
Turbulent Flow in Curved Square Duct: Prediction of Fluid flow and Heat trans...Turbulent Flow in Curved Square Duct: Prediction of Fluid flow and Heat trans...
Turbulent Flow in Curved Square Duct: Prediction of Fluid flow and Heat trans...IRJET Journal
 
Atmospheric turbulent layer simulation for cfd unsteady inlet conditions
Atmospheric turbulent layer simulation for cfd unsteady inlet conditionsAtmospheric turbulent layer simulation for cfd unsteady inlet conditions
Atmospheric turbulent layer simulation for cfd unsteady inlet conditionsStephane Meteodyn
 
Poor blade curvature a contributor in the loss of performance in the compre...
Poor blade curvature   a contributor in the loss of performance in the compre...Poor blade curvature   a contributor in the loss of performance in the compre...
Poor blade curvature a contributor in the loss of performance in the compre...Alexander Decker
 
Advanced fluid dynamics
Advanced fluid dynamicsAdvanced fluid dynamics
Advanced fluid dynamicsjairamanathan
 

Similar to Effect of boundary layer thickness on secondary structures in a short inlet curved duct (20)

In3614821488
In3614821488In3614821488
In3614821488
 
Flow Investigation inside A Curved Square Duct
Flow Investigation inside A Curved Square DuctFlow Investigation inside A Curved Square Duct
Flow Investigation inside A Curved Square Duct
 
Numerical and experimental investigation of co shedding vortex generated by t...
Numerical and experimental investigation of co shedding vortex generated by t...Numerical and experimental investigation of co shedding vortex generated by t...
Numerical and experimental investigation of co shedding vortex generated by t...
 
CFD simulation of Lid driven cavity flow
CFD simulation of Lid driven cavity flowCFD simulation of Lid driven cavity flow
CFD simulation of Lid driven cavity flow
 
Calculation of Fluid Dynamic for Wind Flow around Reinforced Concrete Walls
Calculation of Fluid Dynamic for Wind Flow around Reinforced Concrete WallsCalculation of Fluid Dynamic for Wind Flow around Reinforced Concrete Walls
Calculation of Fluid Dynamic for Wind Flow around Reinforced Concrete Walls
 
M01312106112
M01312106112M01312106112
M01312106112
 
Dr33711716
Dr33711716Dr33711716
Dr33711716
 
Dr33711716
Dr33711716Dr33711716
Dr33711716
 
IRJET- Study of Jet Impingement Heat Transfer
IRJET-  	  Study of Jet Impingement Heat TransferIRJET-  	  Study of Jet Impingement Heat Transfer
IRJET- Study of Jet Impingement Heat Transfer
 
20320140503004
2032014050300420320140503004
20320140503004
 
Vertical canal fall
Vertical canal fallVertical canal fall
Vertical canal fall
 
F0442935
F0442935F0442935
F0442935
 
Turbulent Flow in Curved Square Duct: Prediction of Fluid flow and Heat trans...
Turbulent Flow in Curved Square Duct: Prediction of Fluid flow and Heat trans...Turbulent Flow in Curved Square Duct: Prediction of Fluid flow and Heat trans...
Turbulent Flow in Curved Square Duct: Prediction of Fluid flow and Heat trans...
 
Co24593599
Co24593599Co24593599
Co24593599
 
Atmospheric turbulent layer simulation for cfd unsteady inlet conditions
Atmospheric turbulent layer simulation for cfd unsteady inlet conditionsAtmospheric turbulent layer simulation for cfd unsteady inlet conditions
Atmospheric turbulent layer simulation for cfd unsteady inlet conditions
 
Poor blade curvature a contributor in the loss of performance in the compre...
Poor blade curvature   a contributor in the loss of performance in the compre...Poor blade curvature   a contributor in the loss of performance in the compre...
Poor blade curvature a contributor in the loss of performance in the compre...
 
Advanced fluid dynamics
Advanced fluid dynamicsAdvanced fluid dynamics
Advanced fluid dynamics
 
fmiFinal
fmiFinalfmiFinal
fmiFinal
 
I012545259
I012545259I012545259
I012545259
 
Lk2519761985
Lk2519761985Lk2519761985
Lk2519761985
 

Effect of boundary layer thickness on secondary structures in a short inlet curved duct

  • 1. Effect of boundary layer thickness on secondary structures in a short inlet curved duct Jeremy Gartner 1 , Michael Amitay 2,⇑ Rensselaer Polytechnic Institute, Troy, NY 12180, USA a r t i c l e i n f o Article history: Received 4 February 2014 Received in revised form 15 October 2014 Accepted 17 October 2014 Available online 13 November 2014 Keywords: S-shape inlet Secondary flows Three dimensional separation Secondary flow structures a b s t r a c t The flow pattern in short inlet ducts with aggressive curvature has been shown to lead, in some cases, to an asymmetric flow field at the aerodynamic interface plane. In the present work, a two-dimensional honeycomb mesh was added upstream of the curved duct to create a pressure drop across it, and there- fore to an increased velocity deficit in the boundary layer. This velocity deficit led to a stronger stream- wise separation, overcoming the instability that can result in an asymmetric flow field at the aerodynamic interface plane. Experiments were conducted at Mach numbers of M = 0.2, 0.44 and 0.58 in an expanding aggressive duct with rectangle to a square cross section with area ratio of 1.27. Steady and unsteady pressure measurements, together with Particle Image Velocimetry (PIV), were used to explore the effect of the honeycomb on the symmetry of the flow field. The effect of inserting a honey- comb was tested by increasing its height from 0 to 2.2 times the boundary layer thickness of the baseline flow upstream of the curve. Using the honeycomb, flow symmetry was achieved for the specific geomet- rical configuration tested with a negligible decrease of the pressure recovery. Ó 2014 Elsevier Inc. All rights reserved. 1. Introduction Heightened interest in short and curved inlet ducts for aircraft has led us to further explore the flow field existing in such devices as well as to a better knowledge of the issues associated with their design and methods of mitigating such issues. Several factors related to engine and aircraft performance and operation drive the use of short inlet duct designs, such as the overall airframe length reduction enabled by a shorter duct and reduction of frontal planform by burying the engine into the airframe. These factors led to a reduction in weight and fuel consumption, and allowed for innovative external and integrated aerodynamics, such as Blended Wing Bodies (Dagget et al., 2003). Another factor that must be taken into consideration is the stability margin for operation of a jet engine following the duct, where uneven pressure distribution and secondary flow structures can lead to engine stall at the fan/ compressor stages (surge stall) (Scribben et al., 2006; Mattingly et al., 2002). A considerable body of work is available in the literature con- cerning the analysis of the flow field in short inlet ducts (Bansod and Bradshaw, 1972; Launder and Ying, 1972; Enayet et al., 1982; Wellborn et al., 1992, 1993; Whitelaw and Yu, 1993a,b; Ng et al., 2008, 2006). These previous research efforts have shed light on the main features of the flow field existing in aggressively curved ducts, where the rapid curvature in the duct results in pres- sure gradients in the direction normal to the turn, leading to the onset of secondary flow structures in the form of two counter rotating vortices. Other structures were also noticed to co-exist with these counter-rotating vortices, such as cross stream flow at the internal surfaces, which invade the local boundary layer lead- ing to further flow detachment (Wellborn et al., 1992, 1993; Ng et al., 2008, 2006; Chen, 2012) disrupting the flow and creating recirculation zones in the duct. The symmetric counter-rotating vortices can be described by inviscid flow equations, caused solely by the turning of the flow. These pressure driven counter-rotating vortices convect the low momentum fluid of the boundary layer towards the center of the duct impacting flow uniformity and pres- sure recovery at the face of the engine located downstream, at the aerodynamics interface plane. Implementation of passive and active flow control techniques in short inlet ducts has been an active field of research (Scribben et al., 2006; Ng et al., 2008, 2006; Chen, 2012; Vaccaro, 2011; Debronsky, 2012; Amitay et al., 2002; Gissen et al., 2011; Jirasek, http://dx.doi.org/10.1016/j.ijheatfluidflow.2014.10.016 0142-727X/Ó 2014 Elsevier Inc. All rights reserved. ⇑ Corresponding author. E-mail address: amitam@rpi.edu (M. Amitay). 1 Department of Mechanical, Aerospace & Nuclear Engineering, 110 8th Street, Troy, NY, USA. 2 James L. Decker’45 Endowed Chair of Aerospace Engineering, and Director of the Center for Flow Physics and Control (CeFPaC), 110 8th Street, Troy, NY, USA. International Journal of Heat and Fluid Flow 50 (2014) 467–478 Contents lists available at ScienceDirect International Journal of Heat and Fluid Flow journal homepage: www.elsevier.com/locate/ijhff
  • 2. 2006; Reichert and Wendt, 1994). The predominant forms of actu- ation are vortex generators, steady and unsteady jet blowing tan- gent to the surface, synthetic jet actuators, and many more. Recent work has studied multiple actuation devices (Vaccaro, 2011; Gissen et al., 2011) including combination of flow control techniques. Although most of the previous work was focused on circular cross section ducts (Wellborn et al., 1992, 1993; Whitelaw and Yu, 1993a,b; Gissen et al., 2011; Jirasek, 2006), emphasis was also given to rectangular cross section ducts (Launder and Ying, 1972; Ng et al., 2008, 2006; Chen, 2012; Vaccaro, 2011; Debronsky, 2012; Amitay et al., 2002). All of the work performed with flow control had the objective of improving the pressure recovery and pressure distribution at the exit of the duct. Recent experiments (Vaccaro, 2011; Debronsky, 2012) and numerical simulations (Chen, 2012) have shown that, under some geometrical conditions, the flow can become asymmetric, where one of the counter-rotating vortical structure supersedes the other. In the case of the rectangular ducts, the secondary flow structures were shown to move towards one of the corners of the duct (Chen, 2012; Vaccaro, 2011; Debronsky, 2012). As noted by Chen (2012), the secondary flow phenomenon (i.e., a turbulent flow with mean streamwise vorticity) is attrib- uted to two mechanisms: (i) the skew induced, inviscid mecha- nism, which is caused by any bend in the flow path of ducts with any cross sectional shape as shown by Miller (1991) (and Fig. 1 below), and (ii) a stress-induced mechanism occurring in any non-circular ducts, straight or not, due to anisotropy of the Reynolds stresses. A more in-depth description of secondary flow can be found in Perkins (1970) and Bradshaw (1987). Also note that further complexity in the flow structures is due to swirl development in the second bend of the s-duct. This reverse in the curvature is accredited with the crossover of the transverse velocity component near the side walls, an essentially inviscid process. Another feature of short inlet ducts is the adverse pres- sure gradient caused by the opposite curvature of the second bend. Therefore, the secondary flow generated by the first bend is attenuated, being reversed depending on the aggressiveness of the turn (i.e., the aspect ratio L/D, the offset and area ratio between the inlet and the exit sections). Fig. 2a (taken from Wellborn et al., 1993) shows a three-dimen- sional perspective of the owl face separation topology with a coun- ter-rotating pair of vortices orientated with upwelling along the centerline. The skeleton drawing of Fig. 2b shows the schematic of a symmetric but unstable owl face separation of the first kind. Perry and Hornung (1984) suggested that this unstable symmetric distribution could exist due to slight variations in the flow field. However, they stated that it was a special condition and that any asymmetry in the flow would cause the streakline pattern to shift to one side as shown in Fig. 2c. Based on previous experiments (Ng et al., 2008, 2006; Vaccaro, 2011) and numerical analysis (Chen, 2012), it was shown that a critical length to diameter ratio exists in a rectangular cross section compact inlet duct, controlling the asymmetry of the secondary structures. For ducts longer than this critical length the flow pat- terns are asymptotically stable. With the reduction of duct length below this critical value, the streamwise pressure gradient increases and interacts with the transverse invasion. Simulta- neously, the forward moving main flow confronts the backflow to the streamwise separation. The symmetric pattern becomes unstable due to the saddle–saddle connections existing in the topology of the flow, leading to a flow bifurcation (Tobak and Peak, 1982). This asymmetric configuration is the starting point for the cur- rent work, which also derives from the observation (Enayet et al., 1982; Chen, 2012) that the secondary flow structures forming on the inside of duct bends have their strength dependent on inlet flow conditions, specifically the momentum thickness of the Nomenclature AIP aerodynamic interface plane D width of the duct (mm) h honeycomb height (mm) d boundary layer thickness (mm) h/d relative height of the honeycomb h momentum thickness (mm) L length of the duct (mm) Minlet Mach number at the inlet PR pressure recovery Pinlet static pressure at the inlet (atm) c specific heat ratio of air Cp pressure coefficient P0 total pressure (atm) P static pressure (atm) Fig. 1. Development of secondary flows in a pipe bend showing the (a) presence of an adverse pressure gradient, and (b) direction and orientation of the secondary flow (from Miller, 1991). 468 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478
  • 3. Fig. 2. Owl face separation of the first kind topology showing the (a) three-dimensional perspective of the separation, (b) symmetric skeleton schematic of streaklines, and (c) skeleton schematic of asymmetric streakline pattern (from Wellborn et al., 1993). Fig. 3. The experimental facility. Fig. 4. Exploded view of the inlet duct assembly. ∗ Fig. 5. Inlet duct with the Instrumentation Can for AIP measurements. Fig. 6. Pressure sensors distribution at the AIP. J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478 469
  • 4. incoming boundary layer. Therefore, the objective of the current work is to show that proper manipulation of the incoming bound- ary layer can lead to modification of the flow and pressure fields without incurring any significant pressure loss. The main gain expected from the manipulation of the incoming boundary layer is on the cross flow symmetry of the pressure distribution at the AIP. Honeycomb Mesh Quarter of a dollar Fig. 7. Compact inlet with the honeycomb inserted. Fig. 8. Internal view with the honeycomb inserted. Fig. 9. Inlet total pressure field measured at M = 0.44; (a) across the duct height, and (b) across the duct span. Fig. 10. Cross-stream distributions of the streamwise velocity profiles upstream of the turn at M = 0.44. Fig. 11. Effect of the honeycomb height on the shape of the velocity profile at x/ D = À0.58. Fig. 12. Effect of the honeycomb height on the shape of the velocity profile at x/ D = À0.14. Table 1 Momentum displacement at x/D = À0.58. h/d0 h (mm) 0 0.55 0.74 0.95 1.47 1.04 2.2 1.16 470 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478
  • 5. 2. Experimental setup The high subsonic facility at the Center for Flow Physics and Control (CeFPaC) at Rensselaer Polytechnic Institute was utilized in the current experiments. The high subsonic facility is a blow down, open return wind tunnel. The inlet duct geometry allowed for Mach numbers up to 0.6 although most data presented was conducted at a Mach number of 0.44 (mass flow rate up to 1.76 kg/s). Fig. 3 shows the flow path of the facility and labels each component. The air begins in the blower and transitions through the blower diffuser into the settling chamber. Next, it enters con- traction followed by the inlet duct (i.e., the test section), and finally exits the facility through a diffuser. The blower used is a Cincinnati Fan model HP-12G29 run by a 100 HP motor that is controlled by a variable frequency drive. The blower can produce a volumetric flow rate up to 170 m3 / min. The air exits the blower and enters the diffuser section that transitions the circular cross-section of the blower to the square cross-section of the settling chamber. The air is slowed as it enters the settling chamber where the fluid is conditioned through a set of screens and honeycomb. In addition, a thermocouple and a static pressure ring were monitored for all experiments and were used in the calculation of the Mach number. The air then enters the con- traction section with a contraction ratio of 142:1 and a conven- tional 5th order polynomial curvature. The air then enters the inlet duct, which has a constant cross- section with a length of 0.3048 m for boundary layer growth as well as to measure the inlet Mach number. Another static pressure ring, as well as a thermocouple, was instrumented in this constant cross-section section. Utilizing a one-dimensional isentropic flow assumption, the inlet Mach number was found from the static pressure ring in the inlet as well as the static pressure ring of the settling chamber. This assumption was validated in the previous work done by Vaccaro (2011) utilizing a total pressure Kiel probe in the constant cross-section region of the inlet duct. The total pressure of the inlet nearly matched the pressure measured in the static pressure ring of the settling chamber, where the air velocity is small enough to be assumed to be zero allowing the pressure and temperature measured in the settling chamber to be taken as the total quantities. From the total quantities measured in the settling chamber and the static pressure measured in the inlet, the Mach number from the isentropic flow assumption is: Minlet ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi P0 Pinlet cÀ1 c À 1 # Ã 2 c À 1 v u u t ð1Þ Following the constant area section of the inlet is the diffusing S-shaped section. After the air goes through the inlet duct, it exits the wind tunnel through a diffuser. This diffuser angle is 3° to reduce flow speed as it exits into the open room. The inlet duct has a length-to-diameter ratio of 1.6, where the initial rectangular cross-section area of 90 mm tall by 114.3 mm wide transitions to a square cross-section 114.3 mm by 114.3 mm resulting in an area ratio of Aexit/Ainlet of 1.27. The design Fig. 13. Streamwise pressure coefficient distribution for the three different honeycomb actuator heights at M = 0.2. J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478 471
  • 6. was modular to allow for easy removal and exchange of parts as can be seen in Fig. 4, which shows an exploded view of the inlet duct. For example, two sets of windows were fabricated: one set made of aluminum for most of the experiments and the other out of optical grade acrylic solely for the use of PIV. This reduced the chance of scratches occurring on the windows that would have a negative impact on the PIV quality and potentially causing glares and/or inaccuracies in the collected data. The inlet duct was instrumented with 126 static pressure taps along the lower surface of the duct. The spanwise density of static taps increased towards the side-wall with an array on the opposite side for checking symmetry (see Fig. 4). The static pressure taps were sampled by the means of four pressure scanners (Scanivalve DSA3217, 16 channels each, ±5 psid full scale and accuracy of ±0.05% of full scale, or ±0.0025% psi). Because only 64 ports could be sampled at a time, two runs were required to sample all of the static pressure taps. The static pressures measured from these taps were used to cal- culate the pressure coefficient CP at each location. The definition of CP used for this calculation is: CP ¼ 2 cM2 inlet P Pinlet À 1 ð2Þ The coordinate systems is defined such the origin is at the mid- dle of the cross-section plane at the beginning of the curvature. x is the streamwise direction, y is the cross-stream direction (i.e., nor- mal to the surface) and z is along the span. Another axis was defined as y⁄ in order to quantify the boundary layers, where y⁄ is aligned with y, but is shifted such that it is defined to be zero on the floor (see Fig. 5). In addition to the static surface pressure sensors, an Instrumen- tation Can was mounted at the exit of the inlet duct to acquire the total pressure field at the aerodynamic interface plane (AIP). Fig. 5 shows the Instrumentation Can mounted to the inlet duct with eight rakes each containing five sensors (a total of 40 sensors). Each sensor location on the rake has a high frequency pressure trans- ducer (Kulite model XCQ-062, ±5 psid full scale and repeatability of ±0.5% of full scale, or ±0.025% psi) as well as a steady total pres- sure tap that can be sampled with the four Scanivalves. The num- bering scheme of the Kulite sensors at the AIP is shown in Fig. 6. The perspective of Fig. 6 is the same orientation as Fig. 5 where the observer is looking downstream toward the AIP. The 40 Kulites at the AIP were utilized to measure the steady and unsteady pressure. The steady pressure was used to plot con- tours maps of the pressure recovery, which is the total pressure at the AIP normalized by the total pressure at the inlet. The average pressure recovery, PRavg, and the lower half pressure recovery, PRlowerhalf, are defined respectively as PRavg ¼ ð P40 1 P0;iAIPÞ=40 P0inlet ð3Þ PRlowerhalf ¼ ð P35 21P0;iAIPÞ=15 P0inlet ð4Þ Fig. 14. Streamwise pressure coefficient distribution for the three different honeycomb actuator heights at M = 0.44. 472 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478
  • 7. where the Kulites are numbered as represented in Fig. 6. Fig. 7 shows the compact inlet with the honeycomb inserted inside it. In order to provide some proportional perspective of the experiment’s scaling, a quarter of US dollar is situated next to the honeycomb mesh. An additional internal view of the tunnel is presented in Fig. 8, where the flow direction is out to the page with the same quarter displayed in Fig. 7. In this work, Particle Image Velocimetry (PIV) measurements were conducted in order to collect quantitative flow measurement of the velocity field at different streamwise and spanwise planes within the duct. All PIV experiments conducted in this work uti- lized a commercial LaVision System of software, and TSI hardware. The hardware included two 120 mJ Nd:YAG lasers and a single 1000 Â 1016 pixel resolution TSI CCD camera. Additional hardware was a Martin Magnum 850 fog machine utilized for seeding the flow with O(1 lm) smoke particles. An array of different optics included a cylindrical lens utilized to create the laser light sheet, focal lens to focus the sheet within the measurement field, and camera optics for focusing the camera at different focal lengths and desired fields of interest. Also, the camera and the laser were mounted on computer-controlled traverses to provide a precise location of the measurement planes. The velocity components were computed by the DaVis 7.1 soft- ware using the cross-correlation technique of pairs of successive images with 50% overlap between the interrogation windows. The successive images were processed using a multi-pass method in which the initial and final passes were 32 Â 32 pixels and 16 Â 16 pixels, respectively. Lastly, the averaged velocity field in each plane was averaged over 500 instantaneous velocity fields. The maximum velocity (approximately 150 m/s) corresponds to an average displacement of about 4 pixels with an approximate maximum error of ±0.1 pixel, which corresponds to an error of ±2.5% of the inflow velocity. 3. Results The results section is divided into three sections: (i) upstream of the curved duct, (ii) along the curved duct, and (iii) at the AIP. Both PIV and pressure measurements were conducted and are presented in this section. 3.1. Upstream of the curved duct Prior to the analysis of the effect of the boundary layer thickness on the flow field downstream, we must ascertain that the incoming flow is uniform and symmetric. Thus, the pressure distributions and the boundary layer velocity profiles were measured upstream of the curved section, i.e. prior to the inlet plane of the duct (x/ D = 0). Two techniques were utilized to quantify the incoming flow field. The first method utilized a total pressure Kiel probe, which was traversed across the height and span upstream of the inlet plane at x/D = À1.56. The traversing of the probe was computer controlled and stepped in 0.5 mm increments close to the wall, and 2.5 mm intervals towards the center of the duct. The height and span were normalized by the inlet width (D = 114.3 mm), Fig. 15. Streamwise pressure coefficient distribution for the three different honeycomb actuator heights at M = 0.58. J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478 473
  • 8. and the pressure measured by the Kiel probe was normalized by the total pressure in the settling chamber of the flow facility. Fig. 9 shows the total pressure at Mach 0.44, which is nearly uni- form across the height as well as the span of the duct, except close to the duct walls, where a boundary layer exists. The second method utilized was the PIV, which captured the velocity flow field before the turn (x/D = À0.58) at three spanwise locations (z/D = 0, ±0.25) and enabled the extraction of the bound- ary layer profiles. As presented in Fig. 10, the flow field upstream of the turn is two dimensional. As was mentioned above, the main goal of the present work is to change the weighted contribution of the flow separation to affect the secondary flow structures. Thus, a honeycomb was inserted upstream of the turn at various heights (into the flow) to create a pressure drop across it which generated a velocity def- icit. As the honeycomb was inserted deeper into the flow (i.e., increase h/d0) the velocity deficit of the boundary layer increased, which made it more susceptive to separation. Fig. 11 shows the velocity profiles at a streamwise location of x/D = À0.58. The velocity was normalized by the freestream velocity, U1, and the height was normalized by the boundary layer thickness (d0 = 0.99U1) of the baseline. The value of d0 is defined at x/ D = À0.58 for M = 0.44 and is kept the same throughout the paper for all the different Mach numbers. Similar trends were seen for M = 0.2 and 0.58 but not shown here for brevity. The velocity profiles were also measured at a streamwise loca- tion of x/D = À0.14 and are shown in Fig. 12. The same trend seen upstream is also present at this streamwise location, except that the flow accelerates due to the proximity to the turn. In order to quantify the results presented in Fig. 11 and to show the momentum deficit, which increases due to the insertion of the honeycomb, the momentum thickness, h, is presented in Table 1 for each velocity profile. 3.2. Along the curved surface Next, the distributions of the pressure coefficient, Cp, along the curved surface are presented at the three Mach numbers of 0.2, 0.44 and 0.58 (Figs. 13–15, respectively), at three spanwise loca- tions of z/D = 0, ±0.25 for four different honeycomb heights. These plots are divided into three regions: Region 1, where the flow accelerates around the turn and three dimensionalities develop for the baseline case. However, when the honeycomb is introduced into the flow it results in quasi two-dimensional flow acceleration, where the asymmetry decreases as h/d0 increases. Region 2 represents the region where the baseline flow is sep- arated as indicated by the constant pressure. The presence of the honeycomb affects the separation region by making it more two-dimensional and extends it farther downstream, as is expected since the velocity deficit of the incoming boundary layer was increased (as discussed above). Region 3, where without the honeycomb, the flow on left side of the duct (z/D = À0.25) reattaches first. By increasing the honey- comb height, the difference in the pressure coefficient between z/D = À0.25 to the other two spanwise locations (z/D = 0 and 0.25) decreases. This means that the reattachment point is far- Fig. 16. Pressure recovery at the AIP for the baseline and the three different honeycomb heights at M = 0.2. 474 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478
  • 9. Fig. 17. Pressure recovery at the AIP for the baseline and the three different honeycomb heights at M = 0.44. Fig. 18. Pressure recovery at the AIP for the baseline and the three different honeycomb heights at M = 0.58. J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478 475
  • 10. ther downstream with the honeycomb than without it. There- fore, elevating the honeycomb increases the extent of the sepa- ration region. 3.3. At the AIP Next, the pressure recovery at the AIP was measured and is discussed in this section. Figs. 16–18 present the time-averaged pressure recovery at three Mach numbers and four honeycomb heights. The red region represents a pressure recovery of 1, meaning that there are no pressure losses, and the blue region represents pressure losses. As can be seen, without the honey- comb, the asymmetry, which is a result of the superposition of the separation and the secondary flow structures, is clearly pres- ent for all three Mach numbers. Here, the lowest pressure recov- ery is on the bottom right side. However, increasing the honeycomb height decreases the asymmetry, with minimal effect of the overall pressure recovery. Table 2 presents the pressure recovery values averaged over the entire AIP plane, PRave, and over only the lower half of the AIP plane, PRlowerhalf, for the different honeycomb heights. An important result from this study is the low impact of the honeycomb mesh on the pressure recovery while improving the symmetry of the pres- sure distribution. By looking at the values of Table 2, the largest impact on the pressure recovery is at M = 0.58 between the base- line case (without honeycomb) and the case with the honeycomb inserted at h/d0 = 2.2, where the difference is 0.68%. In addition to the time-averaged pressure recovery, the effect of the honeycomb on the fluctuating total pressure was also explored. Fig. 19 presents the standard deviation of the pressure recovery for the same cases showed in Fig. 18. For the baseline cases (i.e., with- out the honeycomb present) at all three Mach numbers, there is a large concentration of pressure fluctuations in the lower portion of the AIP, which is the region where the PR is the lowest. When the honeycomb is inserted into the duct (where h/d0 = 2.2), the fluctu- ating pressure fields show symmetric distribution of the pressure fluctuations field. In addition to a global evaluation of the fluctuating pressure field, more detailed examination was conducted where the power spectra at three spanwise locations (sensors 23, 28, 33) at the AIP were calculated and are presented in Fig. 20. For all three Mach numbers, without the honeycomb in the flow, the power spectra at the three spanwise locations are different, whereas the spectral content on the left side and the middle of the AIP (sensors 23 and 28, the red and green lines, respectively) contains higher energy than that on the right side (sensor 33, blue line). When the hon- eycomb is inserted into the flow there is a negligible change on the energy content in the left side of the AIP, meanwhile the energy on the right side increases considerably. This modification in the amplitude of the power spectra further suggests that the introduction of the honeycomb induce a more symmetrical flow field (both the time average and the fluctuating). Also, the mag- nitude of the peak, which represents the dominant shedding fre- quency of the separated flow, increases with the honeycomb inserted into the flow, corroborating the results shown previ- ously. The increase in amplitude, especially at the shedding energy, is another indication that the separation becomes more dominant and by doing so helps push the saddle–saddle point farther downstream in the duct. This saddle–saddle point, as mentioned previously, causes the onset of the instability respon- sible for the asymmetry encountered in the flow without the honeycomb actuator. Table 2 Pressure recovery at the AIP. h/d0 M = 0.2 M = 0.4 M = 0.58 PRave PRlower half PRave PRlower half PRave PRlower half 0 0.98904 0.98185 0.94934 0.91631 0.92183 0.87502 0.74 0.98916 0.98217 0.94905 0.91563 0.92115 0.87326 1.47 0.98879 0.98126 0.94690 0.91068 0.91731 0.86505 2.2 0.98857 0.98070 0.94517 0.90656 0.91560 0.86134 Fig. 19. Standard deviation of the pressure recovery at the AIP at M = 0.2, 0.44 and 0.58. Upper row: baseline flow (h/d0 = 0), and lower row: the most symmetrical configuration achieved in the current experiments (h/d0 = 2.2). 476 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478
  • 11. 4. Conclusion The present work showed that, under some conditions, an asymmetric flow field could develop inside a short inlet duct. In order to correct this asymmetry, a two-dimensional honeycomb mesh was inserted upstream of the curved duct to increase the velocity deficit of the incoming boundary layer. It was shown experimentally, using steady and unsteady pressure measure- ments and PIV, that the manipulations of the velocity deficit led to a more promptly flow separation, which, as a result, modified the secondary flow structures and corrected the flow asymmetry at the AIP. The experiments were conducted at Mach numbers of 0.2, 0.44 and 0.58, and the effect of inserting a honeycomb was tested by increasing its penetration into the flow from 0 to 2.2 times the boundary layer thickness of the baseline flow upstream of the curve. Using the honeycomb, flow symmetry was achieved for the specific geometrical configuration tested with a negligible decrease of the pressure recovery at the AIP. Moreover, the span- wise uniformity of the flow increased as the height of the honey- comb was increased. Acknowledgements This work was funded by the Northrop Grumman Corporation (monitored by Ms. Florine Cannelle). The authors would also like to thank the help of Dr. Israel Salvador and Mr. Brian Debronsky. References Amitay, M., Pitt, D., Glezer, A., 2002. Separation control in duct flows. J. Aircraft 39 (4), 616–620. Bansod, P., Bradshaw, P., 1972. The flow in S-shaped ducts. Aeronautical Quarterly 23, 131–140. Bradshaw, P., 1987. Turbulent secondary flows. Annu. Rev. Fluid Mech. 19, 53– 74. Chen, Y., 2012. Subsonic Flows through S-Ducts with Flow Control. Ph.D. Thesis. Department of Mechanical, Aerospace Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, NY. Dagget, D.L., Kawai, R., Friedman, D., 2003. Blended Wing Body Systems Studies: Boundary Layer Ingestion Inlets with Active Flow Control. Technical Report NASA/CR-2003-212670, Hampton, VA. Debronsky, B., 2012. Experimental Investigation of Flow Control Techniques in a Compact Inlet Duct. M.S. Dissertation. Department of Mechanical, Aerospace Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, NY. Enayet, M.M., Gibson, M.M., Taylor, A.P., Yianneskis, M., 1982. Laser-Doppler measurements of laminar and turbulent flow in a pipe bend. Int. J. Heat Fluid Flow 3 (4), 213–219. Gissen, A.N., Vukasinovic, B., McMillan, M.L., Glezer, A., 2011. Dynamics of Hybrid Flow Control in a Boundary-Layer-Ingesting Offset Diffuser. AIAA Paper 2011- 3096, 41st AIAA Fluid Dynamics Conference and Exhibit, Honolulu, HI. Jirasek, A., 2006. Design of vortex generator flow control in inlets. J. Aircraft 43 (6), 1886–1892. Launder, B.E., Ying, W.M., 1972. Secondary flows in ducts of square cross-section. J. Fluid Mech. 54 (Part 2), 289–295. Mattingly, J.D., Heiser, W.H., Preatt, D.T., 2002. Aircraft Engine Design. AIAA Educational Series. Miller, D.S., 1991. Internal Flow Systems, 1990. BHRA, Cranfied, Bedford, UK. Ng, Y.T., Luo, S.C., Lim, T.T., Ho, Q.W., 2006. On swirl development in a square cross- sectioned, S-shaped duct. Exp. Fluids 41 (6), 975–989. Ng, Y.T., Luo, S.C., Lim, T.T., Ho, Q.W., 2008. On the relation between centrifugal force and radial pressure gradient in flow inside curved and S-shaped ducts. Phys. Fluids 20 (5), 055109–055109-15. Perkins, H.J., 1970. The formation of streamwise vorticity in turbulent flow. J. Fluid Mech. 44 (4), 721–740. Perry, A.E., Hornung, H., 1984. Some aspects of three-dimensional separation. II- Vortex skeletons. Zeitshrift fuer Flugwissenschaften und Weltraumforschung 8, 155–160. Reichert, B.A., Wendt, B.J., 1994. Improving Diffusing S-Duct Performance by Secondary Flow Control. NASA Technical Memorandum 106492, also AIAA 32nd Aerospace Sciences Meeting and Exhibit, Reno, NV. Scribben, A.C., Ng, W., Burdisso, R., 2006. Effectiveness of a serpentine inlet duct flow control technique at design and off-design simulated flight conditions. J. Turbomach. 128, 332–339. Tobak, M., Peak, D.J., 1982. Topology of three-dimensional separated flows. Annu. Rev. Fluid Mech. 14, 61–85. Vaccaro, J.C., 2011. Flow Control in a Compact Inlet. Ph.D. Thesis. Department of Mechanical, Aerospace Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, NY. Wellborn, S., Reichert, B., Okiishi, T., 1992. An experimental investigation of the flow in a diffusing S-duct. NASA Technical Memorandum 105809. Fig. 20. Power spectra on the bottom part of the AIP at three spanwise locations and for the three Mach numbers. Upper row: h/d0 = 2.2, and lower row: h/d0 = 2.2. J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478 477
  • 12. Wellborn, S., Okiishi, T., Reichert, B., 1993. A study of the compressible flow through a diffusing S-duct. NASA Technical Memorandum 106411. Whitelaw, J.H., Yu, S.C.M., 1993a. Velocity measurements in an S-shaped diffusing duct. Exp. Fluids 15 (4–5), 364–367. Whitelaw, J.H., Yu, S.C.M., 1993b. Turbulent flow characteristics in an S-shaped diffusing duct. Flow Meas. Instrum. 4 (3), 171–179. 478 J. Gartner, M. Amitay / International Journal of Heat and Fluid Flow 50 (2014) 467–478