You are on page 1of 24

AIAA Propulsion and Energy Forum 1 10.2514/6.

2019-4422
19-22 August 2019, Indianapolis, IN
AIAA Propulsion and Energy 2019 Forum

Effect of Side Gust on the Performance of


a Supersonic Inlet with Bleed System
Hussein K. Halwas and Suresh Aggarwal§
Department of Mechanical and Industrial Engineering
University of Illinois at Chicago
842 W Taylor St, Chicago, IL 60607
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

We report a computational investigation on the effects of side gust on the performance of a


supersonic inlet with a bleed system at Mach 1.8. This is an extension to our recent study that
analyzed a similar inlet without bleed. The supersonic section is designed using the Taylor–Maccoll
method, and the rest of the inlet using a methodology from literature. The bleed is positioned at the
throat inner wall, and its shape is a combination of a ram scoop and flush slot. ICEM CFD software
is used to generate a 3–D structured mesh. ANSYS Fluent 18.2 code is used to perform 3–D
simulations. The effective grid tests are performed using local grid refinements and varying the
total grid points from 3 to 10 million. The effects of side gust on the shock system and flow are
analyzed by examining the Mach number and pressure distributions. The inlet performance is
characterized in term of total pressure loss, flow distortion, and mass flow ratio parameters for gust
speed of 56 m/s and angles of 0, 30 and 600. Results indicate that the inlet performance is
significantly affected by the side gust, although the bleed system mostly removes the low
momentum flow.

I. Nomenclature
2
A = area, 𝑚
M = Mach number
𝑚̇ = mass flow rate, 𝑘𝑔⁄𝑠
𝑝 = static pressure, Pa, 𝑘𝑔⁄𝑚. 𝑠 2
𝑃𝑜 = total pressure, Pa, 𝑘𝑔⁄𝑚. 𝑠 2
𝑇 = temperature, 𝐾
V = velocity magnitude, 𝑚⁄𝑠
 = side gust angle, deg
H = Altitude, m
Subscripts
𝑎𝑣𝑔 = average
𝑖 = inlet
max = maximum
min = minimum
y = condition immediately behind the normal shock wave
∞ = free stream condition
c = capture
b = bleed
out = outlet

 Ph.D. student, MIE Department, University of Illinois at Chicago, Illinois, USA. Email: hhalwa2@uic.edu.
§ Professor, MIE Department, University of Illinois at Chicago, Illinois, USA. Email: ska@uic.edu.

1
Copyright © 2019 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
2

II. Introduction

S UPERSONIC inlet is a critical component of air-breathing propulsion systems, as it converts kinetic energy of
supersonic airflow into pressure energy and delivers it to engine. The delivered airflow should be in correct
quantity at about Mach 0.4~0.6, and with minimum flow distortions and total pressure (Po) loss over a wide range of
flight envelops [1, 2, 3]. During the deceleration or compression process, a series of oblique/conical shock waves
(shock chain) occurs in supersonic part of the inlet, and is terminated by a normal shock, followed by subsonic
compression in a diffuser. Various supersonic inlets are classified based on the location of the shock chain [4, 5]. In
an external compression inlet, the shock chain occurs externally on a compression surface ahead of the cowl lip,
followed by a normal shock near or at the lip [6, 7], while in internal compression inlet, the shock chain occurs
internally after the cowl lip, followed by a normal shock at the throat [8]. Finally, in a mixed compression inlet, the
shock chain occurs both externally upstream of the cowl lip and internally after the cowl lip, followed by a normal
shock at the throat [9]. The propulsion system is significantly affected by the inlet performance. Total pressure
recovery is one of the important parameters used to assess the supersonic inlet performance, and consequently affects
the engine thrust and thus the fuel consumption. For each 1% total pressure loss, there is at least 1-1.5% loss in engine
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

thrust [9, 10]. Thus, maximizing the total pressure recovery at the engine face is an important consideration. In
supersonic inlets, the adverse effects (adverse pressure gradient) of shock waves, boundary layers and their interactions
with shock waves are the main sources of total pressure loss [11]. The adverse effect of normal shock is much higher
than that of oblique shock of same free stream Mach number, and, therefore, to minimize the adverse effects of a
strong normal shock, a multi-stage supersonic compression surface is utilized [6]. The interaction of normal shock
with the boundary layer leads to rapid boundary layer thickening, and possibly flow separation, and consequently high
total pressure loss and flow distortion [12, 13]. To alleviate these adverse effects of boundary layer and its interaction
with shock waves, boundary layer management techniques are employed. Both the active and passive techniques are
currently being used in supersonic aircraft inlets [14]. The former type works on partially removing the low
momentum portion of the boundary layer by using a bleed system via slots [15, 16], porous/perforated surfaces [17]
or scoops [18]. The passive techniques, such as, micro-array actuators, passive bleed system, streamwise slots,
recirculating cavities, vortex generators, contour bumps, and grooves, have also been investigated [11, 19, 20, 21, 22].
In the present study, we used an active bleed system in order to remove the adverse effects of shock boundary layer
interactions, as discussed later in this paper.
Gust is a very important weather phenomenon for both subsonic and supersonic flights. Hurricanes,
thunderstorms, and tornadoes are some of the weather phenomena that produce gusts. Gust occurs at any level in the
atmosphere, including troposphere (6-10km) and above, where most airplanes fly. Gusts have played a significant
role in many catastrophic airplane accidents during the last century [23]. While the effects of gust on airplanes are
important from both structural and aerodynamic (performance) considerations, most previous studies have focused on
the structural aspects. Moreover, there have been numerous experimental and computational studies dealing with the
analysis and design of supersonic inlets, but relatively little work concerning the effects of gust on inlet aerodynamics
and performance [24].
We reported recently a computation investigation on the effects of side gust on the performance of a double-cone
external compression supersonic inlet [25]. Detailed 3-D simulations were performed using the ANSYS Fluent 18.2
code, and the effects of gust on flow field and oblique and normal shocks were analyzed. The inlet performance was
found to be significantly affected by the presence of gust. A major effect of gust was that the shock structures and the
external and internal flows become asymmetric with respect to the longitudinal axis. This results in a stronger shock-
boundary layer interaction causing significant boundary layer separation in the diffuser section, and, consequently,
increase in total pressure loss and flow distortion. The present study extends our previous investigation to include a
bleed system, and thus examines the effects of side gust and bleed on the performance of supersonic inlet. Similar to
previous study, the gust is characterized by its speed and direction. Thus, a gust speed of 56 m/s (126 mph) and gust
angles of 30o and 60o are considered in the present study. This gust speed corresponds to one-half of the strongest
gust speed of 408 km/h, recorded officially in Australia on April 10, 1996 [26].
The use of active bleed systems in supersonic inlets has been extensively investigated. A number of studies have
focused on determining an effective location for the bleed system. Various locations investigated include the
supersonic compression surface [27, 28], throat inner walls [16, 18, 28, 29, 30], throat inner and outer walls [17, 31,
32], and inner and outer walls in the subsonic diffuser [33]. There have also been studies dealing with an optimum
shape of the bleed system [18, 28, 34, 35]. In the present study, the bleed system is located on the throat inner wall,
and its shape is a combination of a ram scoop and a flush slot. More details of the bleed system design are provided
in the next section.

2
3

As in the previous work [25], 3-D simulations are performed using the ANSYS Fluent 18.2 commercial code.
The geometric model considered is an axisymmetric double-cone external compression supersonic inlet with bleed
system. The cone angles are optimized on the basis of total pressure recovery by using the Taylor–Maccoll method
[36, 37]. The second cone angle is further optimized by using the method of characteristics. The transit or blending
zone, which includes the cowl, inner surface, and the subsonic zone are designed by following the methodology
reported in Refs. [13, 38]. The effects of gust and bleed system on the inlet performance are analyzed in term of total
pressure recovery, effective mass flow rate and flow distortion parameters.
This paper is organized as follows. In the next section, we describe the design of the inlet geometry. Section 4
discusses the computational method, including the governing equations, boundary conditions, grid generation, solution
methodology, performance parameters, grid-dependency, and optimum grid. Results and discussion are presented in
Section 5, while the conclusions are provided in Section 6.

III. Design of Inlet Geometry


The supersonic inlet under study is a bi-conical axisymmetric external compression inlet with a bleed system. It
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

comprises of supersonic, transit, and subsonic parts. The supersonic part consists of two cones. For the optimum total
pressure recovery and low drag at design condition, the shock train, which includes two oblique shocks and one normal
shock, should impinge at the cowl lip [9, 13, 16, 39]. Thus, the geometric angles of these cones are designed and
optimized accordingly by using the Taylor–Maccoll method [36, 37]. The method is based on the assumption of
inviscid and conical flow, and involves the use of Euler equations to describe the flow. The inviscid flow assumption
implies that the computation of total pressure recovery considers only the shock losses. The conical flow implies
constant flow properties along conical compression surfaces starting from the cone vertex [40]. Consequently, the
Euler equations are reduced to ordinary differential equations, which are solved numerically using a fourth order
Range-Kutta method. The computations were performed by using a code written in QBasic language. The field Mach
number of the first cone is considered to be an average of Mach numbers in the region behind the first oblique shock.
The total pressure loss across the second oblique shock is then calculated using the assumption that flow undergoes a
deflection equal to the difference between the second and the first cone half angles. Similar to the procedure used for
the first cone, the average Mach number of the second cone can be determined, and then the additional total pressure
loss at the normal shock can be computed.
Figure 1 shows the total pressure recovery as a function of two semi-cone angles at the design Mach number of
M∞ = 1.8. Using the data shown in this figure, the design of the supersonic section, including the two cone angles
and cowl lip angle, were optimized. As indicated in the figure, the highest total pressure recovery achieved is 0.984,
corresponding to the first and second semi-cone angles as 𝛿c1 = 200 and 𝛿c2 = 280 , measured from the horizontal
centerline, respectively. However, the second cone angle was further optimized based on considerations that the
intersection point of the two oblique shocks should impinge very close and upstream of the cowl lip, which is also the
location of normal shock at the design condition, but not at the cowl lip. This results in a more stable shock system
over the range of operations [14, 38, 41], compared to having the interaction section point exactly at the cowl lip.
Another consideration is that while the first shock is conical and flow field behind it is conical and irrotational, the
second shock is curved and the flow field behind it is rotational. A procedure proposed by Kennedy [42], who used
the conical flow theory in the irrotational region and the method of characteristic in the rotational region, was
employed to construct the curved shock and find the shocks intersection point. Consequently, the optimized second
cone angle, obtained through an iterative procedure such that the shocks intersection point is located slightly upstream
of the cowl lip, was 𝛿c2 = 260 measured from the horizontal centerline. This resulted in a relatively small spillage of
~0.5% of captured mass flow rate. Finally, the internal cowl lip surface was designed to have the initial incline parallel
to the local flow direction.
In the transit section the supersonic flow converts to subsonic flow via a normal shock, followed by subsonic
compression in the diffuser. As discussed in an earlier study [25], this section starts at the cowl lip and forms the
annular duct of the inlet along with the subsonic diffuser. As noted above, the interior cowl surface angle is aligned
with the local flow direction to make the exterior shock weaker [38, 43]. The exterior cowl surface angle should be
greater than the interior angle by 3-50 for structural considerations [9, 38, 39]. The subsonic diffuser is the last section
of the inlet. The design data for this part has been taken from a previous study [38]. It connects the transit section of
the inlet to the compressor face, which has a radius of 53.34cm, and a co-annular shape with a hub to tip ratio of 0.2.
The hub has an elliptical cross section with an axial length (major axis) of 15.24cm and a radial length (minor axis)
of 10.67cm. The centerbody planar profile assumes a smooth curve intersecting a cylindrical hub at a distance two
times the axial length of the rotor. The cowl planar profile is designed based on the area distribution through the
subsonic diffuser. The surfaces of the subsonic duct are then created by extruding the planar profiles about the axis

3
4

of symmetry. Finally, the bleed slot is located at the compression surface and at the transit section. The bleed entrance
is a combination of ram scoop and flush slot. The ram scoop leads to improved inlet performance near the design
Mach number, while the flush slot gives improvement at smaller Mach numbers [18]. Moreover, the ram scoop
entrance has small or no barrier shock that could increase drag [44]. Figures 2 and 3 depict the geometry of the
supersonic inlet design based on the procedure described above, while Table 1 lists the geometric data corresponding
to the design conditions.

IV. Numerical Method

A. Governing Equations
Details of the governing equations, turbulence model, and computation procedure have been provided in an earlier
study [25]. As discussed, the compressible form of the Navier-Stokes equations (mass, momentum and energy), along
with the equation of state and the Menter’s two-equation SST turbulence model [45], are employed. The turbulence
model uses a combination of k-ω model in the near wall region and k-ε model in the outer region and connects them
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

by blending functions. It represents a significant improvement over the standard two-equation turbulence model,
especially for the prediction of adverse pressure gradient flows. The compact form of 3-D Navier-Stokes equations
along with the SST turbulence model has been discussed in the previous paper [25].

B. Boundary Conditions
At the entrance section of the supersonic inlet, far field boundary conditions are applied. Far field pressure and
temperature are calculated as:
101325
𝑝∞ = 𝐻 (1)
10 ⁄19200

𝑇∞ = 288.16 − 0.0065𝐻 (2)

Thus 𝑝∞ = 19558.77𝑝𝑎 and 𝑇∞ = 199.01𝐾 correspond to the atmospheric pressure and temperature,
respectively, at an altitude of 13716m. The inlet back pressure (static pressure) is specified as 4.69 times the far field
pressure in order to have the normal shock at the cowl lip. No slip flow and adiabatic thermal conditions are specified
at all the inlet walls.

C. Grid Generation and Solution Methodology


ANSYS Fluent 18.2 code is used to perform simulations. The Fluent solver employs a finite-volume approach to
discretize the governing equations. Thus, the computational domain is divided into discrete control volumes using a
general curvilinear grid. After designing the inlet geometry, the AutoCAD 2019 was used to draw the scale model,
which was then exported in a format readable by grid generation software ICEM CFD, which is available in the
ANSYS Fluent package. A three-dimensional structured multi-block mesh was generated for the flow domain using
the ICEM CFD software. The mesh was clustered near the walls to capture the turbulent boundary layer so that the
first grid point is away from the wall at distance of 𝑦 + ≈ 1 as required by the turbulence model. The mesh is also
clustered in the shock regions near the first and second cone zones, and also in the transit zone to the beginning of the
subsonic zone. The computational domain consists of 12 blocks, with 9 million nodes. Simulations performed to
establish grid independence are discussed in the next section. Figure 4 shows the three-dimensional computational
domain and its vertical grid plane.
A desktop workstation with 4 CPUs was used to run the code using a parallel solver. It is important to note that
the Fluent code is very sensitive to the initial solution used to obtain the converged solution. To obtain a converged
solution for a new case, the flow needs to be initialized close to a stagnant state. Thus, the assumed initial solution
corresponds to uniform flow field with a relatively small velocity, and pressure and temperature corresponding to the
ambient conditions at the assumed altitude. The energy equation is not solved during the first few iterations. The
under-relaxation factors of 0.3 and 0.4 are used for solving the discretized momentum and pressure equations,
respectively. As solution gets stable after few iterations, the energy equation is turned on. An under-relaxation factor
of one is used for the energy equation, while the under-relaxation factor for the pressure equation is changed to 0.7.
For a new simulation case, a converged solution requires about 60,000 iterations based on values of residuals being
10-5 to 10-7 depending upon the flow variable. The post-processing was done using the ANSYS package.
The inlet performance is characterized in terms of the three parameters, namely total pressure recovery (TPR),
mass flow ratio (MFR) and Flow distortion parameter (FD). As described in our previous study [25], TPR is the ratio

4
5

of total pressure (𝑃𝑜𝑒𝑥𝑖𝑡 ) at the inlet exit (compressor face) to the free stream total pressure (𝑃𝑜∞ ). Note that the total
pressure loss includes both the shock and viscous losses. The total pressure at the exit is calculated by the area-
weighted average. The mass flow ratio (MFR) is defined as the ratio of actual inlet mass flow rate to the maximum
capture mass flow rate.
𝑚̇ A
MFR = 𝑚̇ 𝑖 = A 𝑖 (3)
c c

The flow distortion parameter is a measure of flow non-uniformity at the exit, and is defined in terms of the total
pressure variation at the exit (i.e., the compressor face).
𝑃𝑜𝑚𝑎𝑥 −𝑃𝑜𝑚𝑖𝑛
FD = (4)
𝑃𝑜𝑎𝑣𝑔

The average total pressure in this equation is the same as the exit total pressure used in defining TPR. An inlet
with good performance implies as high TPR and MFR values as possible, and as low FD value as possible.
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

D. Grid-Dependency and Optimum Mesh


In order to examine grid independence and determine an optimum mesh, simulations were performed for eight
different cases using grid numbers of 3, 4, 5, 6, 7, 8, 9 and 10 million. The computational domain was divided into
12 blocks, which are connected to each other via their faces. Any change in the number of grid points in the axial or
radial direction in a given any block will lead to changes in the blocks with common faces. The local grid refinement
was implemented in the regions of high gradients as the number of grids was increased. Clearly, a coarse mesh yields
poor results, including incorrect shock locations, while a very highly refined mesh entails high computational cost.
Results for different grids are compared in terms of TPR, mass flow rate through outer boundary1, bleed mass flow
ratio, and MFR. These four parameters are shown in Figs. 6 (a), (b), (c), and (d), respectively, plotted with respect to
the number of grids. As indicated, computational results seem to become nearly grid independent for grid numbers
of 9 million and higher. Note that for all the cases, MFR was kept nearly constant by adjusting the backpressure so
as to bring the normal shock to the cowl lip. Thus, in the present study, detailed simulations were performed using 9
million nodes. In terms of the number of grid points in various sections, there were 125017 structured grid points in
the symmetry plane of the inlet, and 530 grid points distributed axially between the first cone apex and the compressor
face. In the bleed plenum, there were 13125 structured grid points. As noted earlier, the grid points were clustered
near the walls (such that 𝑦 + ≈ 1) and in regions where the shocks are expected to exist. Note that the 3-D
computational domain was generated by rotating the inlet symmetry plane about the central axis using 72 layers with
5o intervals.

V. Results and Discussion


Results from 3-D simulations are now presented for three different gust cases as outlined in Table 2. The base
case corresponds to no side gust, while the other two cases correspond to the side gust speed of 126 mph and gust
angles of α = 30o and 60o. In addition, we discuss the effect of bleed system on the inlet performance without and
with side gust.
Figure 6 presents the Mach number contours for three cases, which include the base case with no side gust (Fig.
6a), and other two cases with side gust angles α = 30o and 60o. From this figure, results indicate that the simulations
capture all the important shock and flow characteristics both outside and inside the inlet. As expected, for the base
case (no gust), the shock waves and flow fields in the upper and lower parts of the inlet are symmetric with respect to
the longitudinal axis. The major features captured by the simulations for this case include (i) two oblique shocks
impinging very close to the cowl lip according to the inlet design discussed earlier, with the second oblique shock
being curved due to the conical flow over the compression surface behind the first oblique shock, (ii) normal shock at
the cowl lip, (iii) local expansion at the bleed slot leading edge, as the approaching flow turns into the bleed region,
which is similar to the CFD results for flow over bleed region reported in Ref. [31], (iv) a shear layer formed from the
slot leading edge toward below the scooped edge, and supersonic jet flow along the bleed plenum’s back surface
starting just after the normal shock, and (v) subsonic flow in the diffusor section. In addition, the supersonic jet in the
bleed plenum, as described, seems to undergo expansion and contraction, similar to that discussed in Refs. [13, 44,
46]. It is also important to note that the bleed system almost completely removes the adverse effects of shock/boundary

1
Outer boundary is shown in Fig. 4

5
6

layer interactions, as evidenced by the fact that there is no separated boundary layer along the diffuser wall downstream
of the normal shock.
For the two gust cases, the Mach number contours in Figs. 6b-c clearly show that the shock waves and flow
characteristics are strongly affected by the presence of gust. The effect seems to be stronger for the gust angle of 30o.
As indicated in Table 2, the change in free stream Mach number caused by gust is also larger for this case. An obvious
effect of side gust is that the flow becomes significantly asymmetric with respect to the longitudinal axis. As indicated
in Figs. 6b-c, the oblique shock waves in the lower part become stronger compared to those in the upper part. This
results in a stronger normal shock in the upper part due to higher Mach number there. Consequently, the interaction
of stronger normal shock with the boundary layer causes boundary layer to rapidly thicken in the upper part and to
occur earlier than that in the lower part. However, in the diverging section, the separated flow region is larger in the
lower part due to larger pressure gradient there, which will be further discussed in the context of Figures 7-8. Another
effect of side gust, as indicated in Fig. 6b, is that the normal shock in the lower part is located closer to the cowl lip
than that in the upper part, due to a lower Mach number ahead of normal shock in the lower part. Moreover, because
of the higher free stream Mach number, the oblique shocks impinge downstream of the cowl lip after intersecting with
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

the cowl internal shock, and this causes a cowl boundary layer to develop.
An important observation from Fig. 6c (higher gust angle) is that the adverse pressure gradient effect, which
causes thickening of the boundary layer in figure 6b, has been removed by the bleed system, as the normal shock is
located at the bleed entrance, which also leads to increased shock stability, preventing the normal shock to move
further upstream. However, since there is no bleed system in the cowl surface, the adverse effects of strong oblique
shocks and the normal shock in the lower part cause rapid boundary layer thickening and subsequently separation in
the subsonic diffuser, as indicated by the dead zone in the lower part. Another observation from Fig. 6c is the absence
of expansion wave near the bleed slot in the upper part. This is due to the fact that the oblique shocks are weaker
there, and flow speed is high enough so as to keep the shear layer above the slot almost parallel to the local flow
direction. Finally, it is also important to note from Figs. 6a-c is that for all three cases, the flow at the bleed slot exit
is choked (M ≈ 1) as designed in order to prevent any back flow.
The asymmetric shock structures and flow asymmetry caused by the side gust are also depicted by the static
pressure contours Fig. 7 for the three cases. These contours complement the information provided by the Mach
number contours in discussed in the context of Fig. 6. Figure 7 also shows two representative streamlines that pass
through the shock waves and diffuser section in the upper and lower parts, and provide additional flow details,
especially on the effect of side gust. The pressure contours for the base case (no side gust) again indicate the symmetry
of flowfield with respect to the longitudinal axis. As expected, the static pressure increases as the flow moves through
the supersonic zone, which contains the oblique and normal shocks, and then through the subsonic zone, where the
subsonic diffuser works on increasing the pressure as the flow approaches the outlet section. More quantitative
information about the pressure distribution is provided in Fig. 8, which plots the pressure variation along the two
streamlines mentioned above. For the base case (Fig.8a), the pressure variations along the two streamlines are
identical due to flow symmetry. Here, the first two pressure jumps are due to the two oblique shocks, while the gradual
pressure rise in between the oblique shocks is due to the conical (irrotational) flow in that region. The third sharp
jump occurs at the normal shock located at the cowl lip, followed by a small drop and then rise in pressure due to the
convex curvature of the surface, then gradual pressure rises in the subsonic diffuser section.
Figures 7b-7c along with Figures 8b-8c provide further details about the effect of side gust on the shock waves
and flow characteristics. The pressure contours and pressure variation along the two streamlines clearly depict the
asymmetry caused by the presence of side gust. As discussed earlier in the context of Fig. 6, an obvious effect of side
gust is that the oblique shocks are stronger in the lower part compared to those in the upper part. Moreover, the
asymmetry in the oblique shock strength increases as the side gust angle is increased. This can be clearly seen in Figs.
8b-8c, which show greater pressure peaks along the lower streamline as the flow passes the two oblique shocks. This
leads to a significantly stronger normal shock in the upper part due to higher Mach number there. Another important
effect of side gust is a series of expansion and compression waves just beyond the cowl lip. These include (i) cowl lip
internal shock and its reflection on the compression surface, (ii) expansion waves at the bleed slot, due to convex
curvature of the compression surface, and (iii) compression waves from concave cowl surface (located just beyond
the cowl lip) in addition to the normal shock, especially in the upper part. These waves are clearly depicted in Fig. 7b
and Fig. 8b. Results in these figures also indicate that these effects of side gust are stronger at gust angle of 30 o
compared to those at higher gust angle. Another effect of side gust is the asymmetry in static pressure distribution in
the diffuser section, with the pressure being higher in the upper part compared to that in the lower part. This generates
radial flow in the diffuser section, with part of airflow going from the upper part to lower part. This also increases the
size of separation zone in the lower part compared to that in the upper part.

6
7

The asymmetry in terms of the size of separated flow and dead zones can be seen more clearly in Figs. 9 and 10,
which plot the total pressure (normalized) and Mach number contours, respectively, at the outlet section for the various
cases. For the base case, the symmetry in the flowfield is quite evident in both the total pressure and Mach number
contours. For the side gust cases, the asymmetry in the flow field caused by the gust is almost clearly shown by these
contours. In order to examine the effect of bleed system, the contours in Figs. 9c and 10c (with bleed) can be compared
with the corresponding contours (without bleed) in Figs. 9d and 10d. The comparison indicates that the bleed system
is able to partially remove the low momentum boundary layer from the centerbody surface. As further shown in Figs.
9c and 10c, there is still some boundary layer separation and dead zone formed in the lower part of subsonic diffuser,
indicating the need for an additional bleed system on the cowl surface.
Thus, the results indicate that the side gust significantly affects the shock waves and flow characteristics in
different parts of the inlet. It also has a noticeable effect on the flow in the bleed system, as illustrated in Fig. 11,
which plots the Mach number and static pressure along a vertical line (Fig. 11a) passing through the bleed plenum at
coordinates from (0.8, 0.3, 0) to (0.8, -0.3, 0). Again, as indicate by the Mach number and static pressure plots in
Figs. 11b and 11c, for the base case, the flow in the bleed system is symmetric with respect to the longitudinal axis,
while the presence of side gust causes significant asymmetry. Consistent with the description of flow in the bleed
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

plenum, Fig. 11b shows supersonic flow along the plenum scoop surface for all three cases.

Effects of Side Gust and Bleed System on the Inlet Performance


As mentioned earlier, the inlet performance is characterized in terms of three parameters, namely, the total
pressure ratio (TPR), mass flow ratio (MFR), flow distortion (FD). Figure 12 plots the TPR at the outlet as a function
of gust angle for the two inlets with and without a bleed system. As indicated, TPR values are the highest and lowest
for =0o and 30o, respectively. Thus, the total pressure loss is the minimum for the base case, while it is the highest
for the side gust angle =30o. Moreover, the use of a bleed system increases the TPR for both gust cases =30o and
60o, with the increase being higher for the latter case.
Figure 13 plots the mass flow ratio (MFR) as a function of gust angle with and without a bleed system. It is
interesting to note that for the base case, MFR is about 2% higher with bleed compared to that without bleed. This is
due to the fact that for the latter case, the intersection point of the first and second oblique shocks is not at the cowl
lip but upstream of it, which causes some flow to spill over the cowl, and therefore, the MFR is less than one. However,
with bleed and further optimization of second cone angle, the shock intersection point is pushed closer to cowl lip,
which increases MFR by 2%. It is also interesting to note that the MFR for the side gust case with =30o is higher
compared to the base case. This is due to the fact that for the gust angle of α=30o, the effective free stream Mach
number becomes high enough to push the shocks intersection point downstream of the cowl lip, and, consequently,
the MFR is higher. However, for the larger angle gust (α=60o), the effective free stream Mach number is reduced,
while the asymmetry in shock structures is increased as discussed earlier, compared to that for α=30o. As a
consequence, the shocks intersection point moves outside the cowl lip in the upper part, which leads to lower MFR.
Figure 14 shows the variation of FD (flow distortion parameter) with respect to gust angle for the two inlets with
and without a bleed system. As indicated, for all the three cases, the presence of a bleed system lowers the FD, i.e.,
improves flow uniformity at the compressor face. In addition, as expected, FD has the lowest value for the base case
(α= 0o), while it has the highest value for the side gust case with α=30 o. This is consistent with the results discussed
earlier regarding the shock wave and flow asymmetry caused by side gust, causing flow non-uniformity and
asymmetry in pressure distribution at the outlet.

VI. Conclusions
The effects of side gust and bleed system on the performance of a double-cone external-compression supersonic
inlet have been numerically investigated. The inlet geometry is designed and optimized by using a combination of
Taylor-Maccoll method and existing methodology. The second cone angle is further optimized by using the method
of characteristics in order to get the oblique shocks intersection location very close to but not at the cowl lip, with the
objective of having a stable shock system with a relatively small flow spilling over the cowl. The bleed system is
located at the throat inner wall, and its shape is a combination of a ram scoop and a flush slot in order to provide
improvement in total pressure recovery at design and off-design Mach numbers. A 3-D structured mesh with 9 million
nodes is generated, and 3-D simulations are performed using the ANSYS Fluent 18.2 code. Results of grid-refinement
study are presented. The inlet performance is characterized in terms of three parameters, namely, total pressure
recovery (TPR), mass flow ratio (MFR), and flow distortion (FD). Important observations are:
1. For all the cases, simulations capture the important shock and flow characteristics both outside and inside the
inlet. For the base case (without side gust), the shock waves and flow fields are symmetric with respect to the

7
8

longitudinal axis, with the two oblique shocks originating from compression surfaces and impinging very close
to the cowl lip, along with a normal shock at the cowl lip, and finally subsonic compression in the diffusor section.

2. The shock wave and flow characteristics are strongly affected by the presence of gust. The major effect is that
the shock structures and the external and internal flows become asymmetric with respect to the longitudinal axis.
Thus, the oblique shocks in the lower part of the inlet are stronger compared to those in the upper part, and,
consequently, there is a stronger normal shock in the upper part compared to that in the lower part. The asymmetry
also leads to stronger shock-boundary layer interaction causing significant boundary layer separation, secondary
(radial) flow, and separated flow region in the diffuser section, especially in the lower half. As a consequence,
the inlet performance is adversely affected by the gust, as indicated by the performance parameters, TPR, MFR,
and FD. The overall effect seems to be the stronger at gust angle of 30o compared to that at 60o.
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

3. Results demonstrate the effectiveness of bleed system in removing the adverse effects of shock waves and their
interaction with the boundary layer from the centerbody walls, especially for the side gust cases. This leads to
improvement in the inlet performance, especially with respect to flow distortion at the inlet exit, which is
noticeably reduced. Results further indicate that the bleed system is not quite effective in removing the adverse
pressure gradient effect for the low gust angle case ( = 300), as the normal shock moves further downstream of
the cowl lip and gets stronger. This causes a significant degradation in TPR (more than 15%) and corresponding
increase in FD for the low gust angle case. This indicates the need for an additional bleed system near the cowl
surface in order to remove this adverse effect and provide further improvement in inlet performance.

4. Other possible effects of side gust include flow instabilities that could potentially lead to inlet buzzing. The
investigation of such scenarios, however, requires transient flow simulations using a time-accurate 3-D
compressible flow solver. A future study should also consider additional bleed systems and optimization of their
locations in order to further improve the inlet performance and alleviate the adverse effects of side gust.

Acknowledgments

Hussein K. Halwas would like to gratefully acknowledge his graduate scholarship from the Iraqi Ministry of
Higher Education and Scientific Research via University of Babylon.

References

8
9

[1] Farokhi, S., “Aircraft Propulsion,” 2nd Edition, John Wiley and Sons Ltd, 2014.
[2] Archer, R. D., and Saarlas M., “An Introduction to Aerospace Propulsion,” Prentice-Hall Inc, 1996.
[3] Kim, S. D., and Song, D. J., “Numerical Study on Performance of Supersonic Inlets with Various Three-
Dimensional Bumps,” Journal of Mechanical Science and Technology, Vol. 22, 2008, pp. 1640–1647.
doi: 10.1007/s12206–008–0503–9.
[4] Kim, H., Kumano, T., Liou, M. S., Povinelli, L. A., “Optimal Shape Design of Supersonic Bypass Inlet,” 13th
AIAA/ISSMO Multidisciplinary Analysis and Optimization Conference, AIAA, Fort Worth, Texas, 13–15
September 2010.
[5] Anderson, J. Th., “How Supersonic Inlets Work_Details of the Geometry and Operation of the SR–71 Mixed
Compression Inlet,” Copyright Lockheed Martin Skunk Corporation, August 2013.
[6] Oswatitsch, K., “Pressure Recovery for Missiles with Reaction Propulsion at High Supersonic Speeds (The
Efficiencies of Shock Diffusers),” NACA TM 1140 (translation), 1947.
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

[7] Ferri A., and Nucci, L. M., “Preliminary Investigation of a New Type of Supersonic Inlets,” NACA Report 1104,
1951.
[8] Kantrowitz, A., and Donaldson, C. D., “Preliminary Investigation of Supersonic Diffusers,” NACA ACR L5D20,
1945.
[9] Seddon, J., and Goldsmith, E. L., “Intake Aerodynamics,” AIAA Educational Series, 2nd Edition, 1999.
[10] Ran, H., and Mavris, D., “Preliminary Design of a 2D Supersonic Inlet to Maximize Total Pressure Recovery,”
5th AIAA ATIO Conference, AIAA, Arlington, Virginia, 26–28 September 2005.
[11] Rolston, S. C., and Raghunathan, S., “Passive Control of Pre-Entry Shock in Supersonic Intakes,” Aeronautical
Journal, Vol. 98, No. 971, Jan. 1994, pp. 1–8.
[12] Soltani, M. R., Younsi, J. S., and Daliri, A., “Performance Investigation of a Supersonic Air Intake in the Presence
of Boundary Layer Suction,” Journal of Aerospace Engineering, Vol.0, No. 0, I Mech E 2014, pp. 1–15.
[13] Faro, Ione D. V., “Supersonic Inlets,” AGARDograph 102, Advisory Group for Aerospace Research and
Development, NATO, May 1965.
[14] Loth, E., and Babinsky, H., “A Representative Flowfield of External Compression Inlets and Diffusers,” 47th
AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition, AIAA–2009–
0032, Orlando, Florida, 5–8 January 2009.
[15] Slater, J., “CFD Methods for Computing the Performance of Supersonic Inlets,” 40th AIAA/ASME/SAE/ASEE
Joint Propulsion Conference and Exhibit, AIAA–2004–3404, Fort Lauderdale, Florida, 11–14 July 2004.
[16] Ryu, K. J., Lim, S., Song, D. J., “A Computational Study of the Effect of Angle of Attack on a Double–Cone Type
Supersonic Inlet with a Bleeding System,” Computer & Fluids Journal, Vol. 50, 2011, pp. 72–80.
doi: 10.1016/j.compfluid.2011.06.019.
[17] Slater, J. W., and Saunders, J. D., “Computational Fluid Dynamics (CFD) Simulation of Hypersonic Turbine-
Based Combined-Cycle (TBCC) Inlet Model Transition,” 16th International Space Planes and Hypersonic
Systems and Technologies Conference, AIAA–2009–7349, Bremen, Germany, 19–22 October 2009.
[18] Connors, J. F., Lovell, J. C., Wise, G. A., “Effects of Internal-Area Distribution, Spike Translation, and Throat
Boundary-Layer Control on Performance of a Double-Cone Axisymmetric Inlet at Mach Numbers from 3.0 to
2.0,” NACA RM E57F03, August 1957.
[19] Anderson, B. H., Tinapple, J., Surber, L., “Optimal Control of Shock Wave Turbulent Boundary Layer Interactions
Using Micro-Array Actuation,” 3rd AIAA Flow Control Conference, San Francisco, California, 5–8 June 2006.
[20] Srinivasan, K. R., Loth, E., Dutton, J. C., “Aerodynamics of Recirculating Flow Control Devices for Normal
Shock/Boundary-Layer Interactions,” AIAA Journal, Vol. 44, No. 4, April 2006, pp. 751–763.
[21] Kim, S. D., “Aerodynamic design of a Supersonic Inlet with a Parametric Bump,” Journal of Aircraft, Vol. 46,
No. 1, January–February 2009, pp. 198–202.

9
10

[22] Baydar, E., Lu, F. K., Slater, J. W., “Vortex Generators in a Two–Dimensional, External–Compression Supersonic
Inlet,” Journal of Propulsion and Power, Vol. 34, No. 2, 2018, pp. 521-538.
[23] Burnham, J., "Atmospheric Gusts, A review of the Results of Some Recent Research at The Royal Aircraft
Establishment," Royal Aircraft Establishment, Bedford, England, Vol. 98, No. 10, October 1970.
[24] Barry, F. W., "Development of Atmospheric Gust Criteria for Supersonic Inlet Design," NASA Report, Hamilton
Standard Division of United Aircraft Corporation, Windsor Locks, Connecticut, December 1968.
[25] Halwas, H. K., Aggarwal, S. K., “Effect of Side Gust on the Performance of an External Compression Supersonic
Inlet,” Journal of Aircraft, Online, https://doi.org/10.2514/1.C035093.
[26] "Info note No.58 — World Record Wind Gust: 408 km/h". World Meteorological Association. 2010–01–22.
Archived from the original on 2013-01-20.
[27] Ferri A., and Nucci, L. M., “The Origin of Aerodynamic Instability of Supersonic Inlets at Subcritical
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

Conditions,” NACA Report RM L50K30, January 1951.


[28] Stewart, D. G., “Supersonic Intake Instability–Further Investigations on Intakes of 250 Cone Semi–Angle at
Mach Numbers up to 2.14 with and without Boundary layer Bleed,” Australian defense scientific service,
aeronautical research laboratories, Report ARI/M.E. 112, January 1964.
[29] Connors, J. F., and Woollett, R. R., “Performance Characteristics of Several Types of Axially Symmetric Nose
Inlets at Mach Number 3.85,” NACA RM E52115, November 1952.
[30] Gawienowski, J. J., “The Effects of Boundary–Layer removal Through Throat Slots on the Internal Performance
of a Side Inlet at Mach Numbers of 2.0 and 2.3,” NASA TM-502, March 1961.
[31] Slater, J. W., and Saunders, J. D., “Modeling of Fixed-Exit Porous Bleed Systems,” 46th AIAA Aerospace
Sciences Meeting and Exhibit, AIAA–2008–0094, Reno, Nevada, January 7–10, 2008.
[32] Kwak, E., Lee, H., Lee, S., “Numerical Simulation of Flows Around Axisymmetric Inlet with Bleed Regions,”
Journal of Mechanical Science and Technology, Vol. 24, No. 12, 2010, pp. 2487-2495.
doi: 10.1007/s12206–010–0927–x.
[33] Shoemaker, C. J., and Henry, J. R., “Effects of Suction Boundary-Layer Control on the Performance of a Short
Annular Diffuser with an Upstream Terminal Normal Shock,” NASA TN D–1241, April 1962.
[34] Syberg, J., and Koncsek, J. L., “Bleed System Design Technology for Supersonic Inlets,” Journal of Aircraft,
Vol. 10, No. 7, July 1973, pp. 407–413.
[35] Eichorn, M. B., Barnhart, P. J., Davis, D. O., Vyas, M. A., Slater, J. W., “Effect of Boundary–Layer Bleed Hole
Inclination Angle and Scaling on Flow Coefficient Behavior,” 51st Aerospace Science Conference, AIAA –2013–
0424, Grapevine, Texas, January 7–10, 2013.
[36] Taylor, G. I., and Maccoll, J. W., “The Air Pressure on a Cone Moving at High Speeds,” Proceedings of the Royal
Society of London, Series A, Vol. 139, No. 838, 1933, pp. 298–311.
[37] Maccoll, J. W., “The Conical Shock Wave Formed by a Cone Moving at High Speed,” Proceedings of the Royal
Society of London, Series A, Vol. 159, No 898, 1937, pp. 459–472.
[38] Slater, J. W., “Design and Analysis Tool for External-Compression Supersonic Inlets,” 50th AIAA Aerospace
Sciences Meeting including the New Horizons Forum and Aerospace Exposition, AIAA–2012–0016, Nashville,
Tennessee, January 9–12, 2012.
[39] Mahoney, J. J., “Inlets for Supersonic Missiles,” AIAA Educational Series, Washington DC, 1990.
[40] Sritharan, S. S., and Seebass, A. R., “Finite Area Method for Nonlinear Supersonic Conical Flows,” AIAA
Journal, Vol. 22, No. 2, February 1984, pp. 226–233.
[41] Rodriguez, D. L., “Multidisciplinary Optimization of a Supersonic Inlet Using a Cartesian CFD Method,” 10th
AIAA/ISSMO Multidisciplinary Analysis and Optimization Conference, AIAA–2004–4492, Albany, New York,
30 August–1 September, 2004.

10
11

[42] Kennedy, E. C., “Calculation of the Flow Fields Around a Series of Bi-Conic Bodies of Revolution Using the
Method of Characteristics as Applied to Supersonic Rotational Flow,” OAL/CM 873, June 1956.
[43] Connors, J. F., and Meyer, R. C., “Design Criteria for Axisymmetric and Two-Dimensional Supersonic Inlets and
Exists,” NACA TN 3589, January 1956.
[44] Shih, T. I-P., Rimlinger, M. J., Chyu, W. J., “Three-Dimensional Shock-Wave/Boundary-Layer Interactions with
Bleed,” AIAA Journal, Vol. 31, No. 10, 1993, pp. 1819-1826.
[45] Menter, F. R., “Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications,” AIAA Journal,
Vol. 32, No. 8, August 1994, pp. 1598–1605.
[46] Slater, J. W., “Improvements in Modeling 900 Bleed Holes for Supersonic Inlets,” 47th AIAA Aerospace Sciences
Meeting and Exhibit, AIAA–2009–0710, Orlando, Florida, January 5–8, 2009.
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

Table 1: Inlet geometric data for the design conditions.


First cone semi-angle 20o
Second cone semi-angle 26o
Length of the first cone 37.89cm
Overall length of on-design supersonic part 60.67 cm
Interior surface cowl initial angle 13o
Exterior surface cowl initial angle 17o
Inlet radius 53.14 cm
Outlet radius (Engine face radius) 53.34 cm
The engine spinner radius (semi-minor axis) 10.67 cm
The overall axial length of the engine spinner (major axis) 30.48 cm
The overall length of the supersonic inlet 220.29 cm
On-Design free stream Mach number 1.8
Air model Ideal gas
Altitude (H) 13716 m
Captured mass flow rate (𝑚̇𝑐 ) 154.60 kg/sec
Inlet mass flow rate (𝑚̇𝑖 ) 153.78 kg/sec
Bleed mass flow rate (𝑚̇𝑏 ) 4.8 kg/sec
Outlet mass flow rate (𝑚̇𝑜𝑢𝑡 ) 148.98 kg/sec
Back to free stream static pressure ratio (𝑝𝑜𝑢𝑡 ⁄𝑝∞ ) 4.69

Table 2: Free stream conditions for the three side gust cases studied.
Effective
Vgust Net Flow Angle Effective 𝑃𝑜∞
Cases α (deg.) V∞
(m/sec) (deg.) M∞ (pa)
(m/sec)
1 0 0 508.84 0 1.800 112380.7
2 56 30 558.04 2.876 1.972 146612.2
3 56 60 536.84 5.162 1.906 132323.2

11
12

)
Total pressure recovery, (Poy/Po 0.99

0.98

0.97
1st semi-cone angle, (deg)

16
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

0.96 17
18
19
20
0.95
21
22

0.94

4 6 8 10 12 14
Second semi-cone angle, (deg.)

Fig. 1: Total pressure recovery for the double-cone inlet design for M = 1.8.

Fig. 2: Geometry of the supersonic inlet design (upper half shown). All dimensions are in centimeters.

12
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

13
Fig. 3: Geometry of supersonic inlet at section z-zero.
13
14

Outer boundary
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

(a) Three-dimensional grid and computational domain.

Outer boundary

Compressor face
or
outlet section

(b) Grid in the vertical plane.

Fig. 4: Computational grid and domain.

14
15

0.956

0.954
TPR

0.952
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

0.950

0.948

2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0
Grid number (million)
(a) TPR for various grids used in simulations.

0.30
Mass flow rate through outer boundary

0.25

0.20

0.15

0.10

0.05

0.00

3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0


Grid number (million)
(b) Mass flow rate through outer boundary.

15
16

3.45

3.40
m. b / mc
Bleed mass flow rate (%)

3.35

3.30

3.25
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

3.20

3.15

3.10

3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0


Grid number (million)
(c) Bleed mass flow ratio.

0.997

0.996

0.995
MFR

0.994

0.993

0.992

3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0


Grid number (million)
(d) Inlet MFR.

Fig. 5: Grid refinement results and optimum mesh used in simulations.

16
17
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

(a) Vgust=0 mph, α=0o (b) Vgust=126mph, α=30o

(c) Vgust=126mph, α=60o

Fig. 6: Mach number contours for flow in double-cone inlet for three cases; (a) without side gust, (b) and (c)
with side gust angles of 30o and 60o.

17
18
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

(a) Vgust=0 mph, α=0o (b) Vgust=126mph, α=30o

(c) Vgust=126mph, α=60o

Fig. 7: Static pressure contours for the three cases as discussed in the context of Fig. 6. Two representative
streamlines are also shown.

18
19
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

(a) Vgust=0 mph, α=0o (b) Vgust=126mph, α=30o

(c) Vgust=126mph, α=60o

Fig. 8: Static pressure variation along the upper and lower streamlines (shown in Fig. 7), for the three side
gust cases discussed in the context of Fig. 6.

19
20
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

(a) Vgust=0 mph, α=0o (b) Vgust=126mph, α=30o

(c) Vgust=126mph, α=60o (d) Vgust=126mph, α=60o


without bleed system
from Ref. [25]
Fig. 9: Total pressure contours (normalized) at the outlet section of inlet for four cases; (a) without side gust,
(b) and (c) with side gust angles of 30o and 60o, (d) without bleed system; from Ref. [25]

20
21
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

(a) Vgust=0 mph, α=0o (b) Vgust=126mph, α=30o

(c) Vgust=126mph, α=60o (d) Vgust=126mph, α=60o


without bleed system;
from Ref. [25]
Fig. 10: Mach number contours at the outlet section of inlet for the four cases discussed in the context of
Fig. 9.

21
22
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

(a) A yellow vertical line passing through the bleed plenum at coordinates
(0.8, 0.3, 0) to (0.8, -0.3, 0).

1.75 45000

1.50 base case 42500 base case


30 deg gust 30 deg gust
40000
Static pressure (Pa)

1.25
60 deg gust 60 deg gust
Mach number

37500
1.00
35000
0.75
32500
0.50
30000

0.25 27500

0.00 25000

-0.30 -0.20 -0.10 0.00 0.10 0.20 0.30 -0.30 -0.20 -0.10 0.00 0.10 0.20 0.30
Vertical position (m) Vertical position (m)

(b) Mach number variation for the three cases; red: (c) Static pressure variation for the three cases; red:
Vgust=0, black: α=30o, blue: α=60o. Vgust=0, black: α=30o, blue: α=60o.

Fig. 11: Mach number (Fig. b) and static pressure (Fig. c) variations along a vertical line (shown in Fig. a)
passing through the bleed plenum for three different cases: without side gust, and with side gust angles of 30 o
and 60o.

22
23

1.00

Inlet with bleed


0.95
Inlet without bleed

0.90
TPR

0.85
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

0.80

0.75

0 30 60
Side gust angle (deg.)

Fig. 12: TPR variation with the side gust angle for the two inlets with and without bleed.

1.01

1.00
MFR

0.99

0.98 Inlet with bleed


Inlet without bleed

0.97

0 30 60
Side gust angle (deg.)

Fig. 13: MFR variation with the side gust angle for the two inlets with and without bleed.

23
24

55

50
Downloaded by CARLETON UNIVERSITY on March 28, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2019-4422

45

40
FD (%)

35

30

25

20 Inlet with bleed


Inlet without bleed
15

10

0 30 60
Side gust angle (deg.)

Fig. 14: FD variation with the side gust angle for two inlets with and without bleed.

24

You might also like