Sie sind auf Seite 1von 7

JOURNAL OF AIRCRAFT

Flow Analysis of the F-16XL Aircraft at Transonic Flow Conditions


Okko J. Boelens
Netherlands National Aerospace Laboratory/NLR, 1059 CM Amsterdam, The Netherlands

Downloaded by JOHNS HOPKINS UNIVERSITY on November 15, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.C033211

DOI: 10.2514/1.C033211
In the framework of the Cranked-Arrow Wing Aerodynamics Project International 2, Netherlands National
Aerospace Laboratory performed an analysis of a set of transonic flight conditions available in the Cranked-Arrow
Project database. Flight condition FC70 was used previously during the Cranked-Arrow Wing Aerodynamics Project
International to investigate transonic flow on the F-16XL aircraft. During this project, it was discovered that flight
condition FC70 was flown with a deflected leading-edge flap. To facilitate computational fluid analysis, a transonic
flight condition without deflected control surfaces was judged desirable by the Cranked-Arrow Project members.
Therefore, it was decided to search the Cranked-Arrow Project database for a transonic flight condition without any
control surface deflections. Because no information on the control surface deflections of the other transonic flight
conditions was available to the Cranked-Arrow Project partners, an alternative approach to scrutinize the available
flight-test sectional surface pressure measurements for indications of control surface deflections was adopted. This
analysis revealed that flight condition FC79 was the most likely transonic flight condition to be flown without any
control surface deflections. Flight conditions FC70 and FC79 were analyzed using National Aerospace Laboratorys
in-house-developed flow simulation system ENFLOW. Comparison of the measured and simulated surface pressure
coefficients confirmed that flight condition FC79 was flown without any control surface deflections, and that this flight
condition is thus much better suited for further comparisons between flight-test data and computational fluid
dynamics simulations.

I.

Introduction

OST-EFFECTIVE computational fluid dynamics (CFD) methods play an increasingly important role in simulating maneuver
conditions. Examples are conditions that cannot be simulated in a
wind tunnel properly or are too dangerous to be performed in flight
tests. Before simulation of such demanding maneuver conditions, the
CFD methods should be well validated by evaluation against state-ofthe-art wind-tunnel and/or flight-test data.
The Cranked-Arrow Wing Aerodynamics Project (CAWAP) [1]
provided the CFD community with an excellent database for validation and evaluation purposes. The CAWAP project focused on an
understanding of the flow phenomena appearing on a cranked-arrow
wing relevant to advanced supersonic fighter and transport aircraft.
The subject of investigation was the F-16XL aircraft. The CAWAP
database contains both subsonic and transonic data at (high) flight
Reynolds numbers. The data include surface pressure measurements,
both along butt line stations and fuselage stations, boundary-layer
measurements at four positions on the left wing, skin friction
measurements at the FS 330 station on the left wing, and surface flow
visualizations using tufts [1,2].
Along with the Vortex Flow Experiment 2 (VFE-2), the CrankedArrow Wing Aerodynamics Project International (CAWAPI) [39],
initiated by NASA as a follow-on project, was incorporated under the
NATO Research and Technology Organisation working group AVT113. The objective of CAWAPI was to allow a comprehensive validation and evaluation of current CFD methods against the CAWAP
flight database [1].
A second effort, CAWAPI-2, was established to focus on the unyielding flow conditions from the CAWAPI research program (i.e., a
subsonic high-angle-of-attack flight condition, dominated by multiple interacting vortices and a transonic low-angle-of-attack flight
condition, exhibiting a series of longitudinal shocks as well as a

Received 29 September 2014; revision received 23 March 2015; accepted


for publication 10 June 2015; published online 15 September 2016. Copyright
2015 by National Aerospace Laboratory, NLR, The Netherlands. Published
by the American Institute of Aeronautics and Astronautics, Inc., with
permission. Copies of this paper may be made for personal and internal use, on
condition that the copier pay the per-copy fee to the Copyright Clearance
Center (CCC). All requests for copying and permission to reprint should be
submitted to CCC at www.copyright.com; employ the ISSN 0021-8669
(print) or 1533-3868 (online) to initiate your request.
*Research and Development Engineer, Applied Computational Fluid
Dynamics, Department of Flight Physics and Loads, Aerospace Division.

(weak) wing leading-edge vortex). Partners participating in this new


effort are Airbus Defence & Space, Military Aircraft (Germany),
KTH (Sweden), Lockheed Martin (United States), NASA Langley
(United States), and National Aerospace Laboratory/NLR (The
Netherlands). Within this group, NLR focused on the analysis of the
differences between the CFD data and the flight-test data for the
transonic flight condition FC70.
This article describes the investigation performed at NLR. In
Sec. II, the results obtained for flight condition FC70 during
CAWAPI will be discussed. Next, an analysis of other transonic flight
conditions available in the CAWAP database will be presented,
resulting in the selection of a more suitable flight condition for CFD
analysis (Sec. III). In Sec. IV, the CFD result obtained with NLRs inhouse flow solver ENSOLV for the selected flight condition will be
presented. Finally conclusions are given in Sec. V and complete the
article.

II.

Results for Transonic Flight Condition FC70 During


CAWAPI

Flight condition FC70 was the only transonic flight condition


among the seven test cases investigated during CAWAPI. The transonic flight condition FC70 taken from [1] is as follows: 
3.6 deg;  0.0 deg; M  0.98; h  22; 300 ft; Reft 106 
3.60; load  1g unit; and flightrun number  1525b. The sectional surface pressure distribution for this case can be found in [8].
For this flight condition, it was found that : : : the computed data
agree very well among themselves but differ substantially from the
measurements, except at BL55 and BL95 where all results are in
fairly good agreement : : : [9]. The difference at the outboard butt
line stations (BL153.3 and BL184.5) were attributed to a leadingedge flap deflection (see also Fig. 1), which was not present in the
CFD simulations. This deflection of 9 deg, which at the time of
CAWAPI was unknown to its members, was later found in the postflight-test analysis [1,9]. To clarify the other discrepancies, a range
of possible causes were investigated in [9]. These include both
1) numerical effects, such as grid resolution (see the following discussion), turbulence modeling effects (no major differences observed),
and unsteady effects (no major differences observed for the inboard
stations; unsteady effects are mainly observed on the part of the wing
outside of the air dam and the actuator pod, see also [2]); and 2) flight

The butt line locations are graphically shown in Fig. 6.

Article in Advance / 1

Downloaded by JOHNS HOPKINS UNIVERSITY on November 15, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.C033211

Article in Advance

/ BOELENS

Fig. 1 F-16XL aircraft including control surfaces (taken from [10]).

test reasons, such as differences in angle of attack or sideslip angle


(differences are not caused by changes in these), differences in aircraft
geometry (the leading-edge flap deflection does not explain the
differences in the inboard stations), and effects of wing bending (no
effect visible).
A study on grid resolution [8] revealed that : : : It took a detailed
Euler simulation with a highly adapted mesh to reveal the occurrence
of a shock-vortex interaction that significantly changed the surface
pressure distribution : : : . If the goal is to resolve the pressure distribution over the aircraft, then this feature must be resolved : : : . In [8]
(Fig. 13), it is shown that this shock-vortex interaction mainly
influences the pressure distributions at butt line stations BL70 and
BL80. The grids used during CAWAPI lacked the grid resolution
required to resolve the weak inner wing primary vortex [2], and hence
agreed very well among themselves but differed substantially from
the measurements.

III.

Analysis of Other Transonic Flight Conditions


Available in the CAWAP Database

Fig. 2 Transonic flight conditions (nominal) in the CAWAP database


used for analysis.

At the start of CAWAPI-2, it was decided that, in addition to


transonic flight condition FC70, there was the need for a baseline
transonic flight condition without any control surface deflection to
facilitate the comparison of the CFD results with the flight-test data. It
should, however, be noted that, except for the leading-edge flap
deflection of 9 deg for FC70, which was shown by post-flight-test
analysis, no other information on the state of the control surface
deflections was available to the members of CAWAPI-2. Efforts to
obtain control surface configuration data turned out to be fruitless.
Therefore, NLR adopted the following simple alternative approach
to find a transonic flight condition for which the control surfaces were
not deflected during flight. For a range of flight conditions in the
transonic regime, the flight-test sectional surface pressure measurements were scrutinized for indications of control surface deflections.
The test conditions from the CAWAP database that were used during
this analysis are summarized in Table 1.In Fig. 2, the test conditions
are graphically presented.
Table 1

From this table and figure, it can be seen that the conditions
selected consist of three sets, that is, 1) the low-angle-of-attack
transonic flight conditions, including flight conditions FC68, FC69,
and FC70; 2) the high-angle-of-attack transonic flight conditions,
including flight conditions FC79, FC80, and FC81; and 3) the
reference flight condition FC43 at a comparable angle of attack and
slightly lower Mach than flight condition FC70.
Flight-test sectional surface pressure measurements will be
analyzed and discussed for these three sets in the following
sections. The F-16XL aircraft is equipped with the following
control surfaces on the wing (see also Fig. 1 and [10]): 1) a leadingedge flap outboard of the kink, 2) an elevon inboard of the actuator
pod, and 3) an aileron outboard of the actuator pod. As stated
earlier, the CAWAP database contained no information on the
deflections of these control surfaces.

Transonic flight conditions (nominal) in F-16XL database used for analysis


(data taken from [1])

Flight condition , deg , deg


FC68
3.7
0.0
FC69
3.6
0.0
FC70
3.6
0.0
FC43
5.0
0.0
FC79
9.3
0.0
FC80
8.8
0.0
FC81
8.1
0.0

M
0.90
0.95
0.98
0.81
0.90
0.95
0.98

h, ft
Reft 106
19,600
3.60
21,300
3.60
22,300
3.60
24,000
2.82
19,600
3.60
21,300
3.60
22,300
3.60

Load, g units Flight/run no.


1
1465b
1
1524b
1
1525b
1
1512c
3.7 WUT
1465d
3.5 WUT
1466d
3.2 WUT
1524d

Article in Advance

The sectional surface pressure coefficient cp  p p 


12 u2  will be scrutinized for the following indications of
control surface deflections (see also [11]):
1) A deflection of the leading-edge flap will result in a distinct
suction peak for these transonic flow conditions.
2) A deflected aileron or elevon will result in a distinct signature in
the surface pressure close to the hinge line. For subsonic flow in front
of the hinge line of the respective control surface (cp < cpcritical ),
this signature is a local peak in the surface pressure. For supersonic
flow in front of the hinge line of the respective control surface
(cp > cpcritical ), this signature is either a discontinuity due to an
oblique wave (upward deflection) or an expansion fan (downward
deflection).

Downloaded by JOHNS HOPKINS UNIVERSITY on November 15, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.C033211

A. Low-Angle-of-Attack Transonic Flight Conditions

The sectional surface pressure data along the butt line stations
(BL50, BL70, BL80, BL95, BL153.5, and BL184.5) for the lowangle-of-attack transonic flight conditions FC68 (cpcritical 
0.1878), FC69 (cpcritical  0.0882), and FC70 (cpcritical 
0.0341) are shown in Fig. 3. In this figure and similar figures
discussed later the dashed orange vertical lines indicate the
approximate hinge line locations. Based on these figures, the
following observations can be made:
1) At butt line stations BL70 and BL95, for flight condition FC68,
a discontinuity in the surface pressure, indicative of a deflected
elevon, is observed (locally subsonic oncoming flow). Though at butt
line station BL95, flight condition FC70 also shows a discontinuity in
the sectional surface pressure at xclocal 0.88, this discontinuity is
thought not to stem from a deflected elevon but from an expansion fan
originating from the actuator pod (shown later). Note also that this
discontinuity is slightly aft of the hinge line.
2) At butt line stations BL153.5 and BL184.5, flight conditions
FC68 and FC70 show suction peaks in the sectional surface pressure
distribution at the location of the leading-edge (LE) flap. Note that, at
this location, a vortical structure may be present even without a
deflection of the leading-edge flap. In [2], this vortical structure due

/ BOELENS

to the kink in the wing leading edge was referred to as the outer wing
vortex. A deflection of the wing leading-edge flap will, however,
intensify this vortex.
3) At butt line stations BL153.5 and BL184.5, flight conditions
FC69 and FC70 both show a discontinuity in the surface pressure
(locally supersonic oncoming flow). Flight condition FC68 shows,
downstream of the leading-edge flap, the surface pressure
distribution of an undisturbed wing.
B. High-Angle-of-Attack Transonic Flight Conditions

Figure 4 shows the sectional surface pressure distribution along the


BL stations for the high-angle-of-attack transonic flight conditions
FC79 (cpcritical  0.1878), FC80 (cpcritical  0.0882) and FC81
(cpcritical  0.0341). These sectional surface pressure distributions
reveal the following:
1) Flight condition FC80 shows a discontinuity in the surface
pressure indicative of a deflected elevon. This discontinuity is,
however, not as clear as the one observed for the low-angle-of-attack
transonic conditions. The major indications for a deflected elevon are
present at butt line stations BL70 and BL95 for flight condition FC80
(locally supersonic oncoming flow). For FC81, which is governed by
supersonic flow on most of the wing upper surface, the situation on
the part of the wing inboard of the air dam and actuator pod is unclear.
This flight condition seems to have the inner wing primary vortex
located directly above butt line station BL80 (similar to what was
observed for FC70 in [8]), resulting in continuous high suction
along this butt line station. Though flight condition FC79 also shows
a peak in the pressure at BL70, at xclocal 0.75, the other data points
do not indicate that the elevon is deflected (compare, for example,
with flight condition FC68 at butt line station BL70; see also
Sec. IV.C).
2) Both flight condition FC80 (suction peak at butt line station
BL153.5) and FC81 (suction peak at butt line station BL184.5) show
some indication of a deflected leading-edge flap.
3) At butt line station BL184.5, the pressure distribution of flight
condition FC81 (locally supersonic oncoming flow) clearly shows

Fig. 3 Flight-test sectional surface pressure coefficient along butt line stations (see Fig. 6) for the low-angle-of-attack transonic flight conditions.

Downloaded by JOHNS HOPKINS UNIVERSITY on November 15, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.C033211

Article in Advance

Fig. 4

/ BOELENS

Flight-test sectional surface pressure coefficient along butt line stations (see Fig. 6) for the high angle-of-attack transonic flight conditions.

the effect of a deflected aileron. Though at butt line station BL153.5,


flight condition FC81 shows a discontinuity in the sectional surface
pressure as well, this discontinuity is thought not to originate from the
deflected aileron but from an expansion fan originating from the
actuator pod.
4) Note that outboard of the air dam and the actuator pod the
sectional surface pressure distribution for flight condition FC79 is fairly
smooth (i.e., no indications of any deflected surface are present).
C. Reference Flight Condition at a Comparable Angle of Attack and
Slightly Lower Mach than FC70

Finally, Fig. 5 compares the sectional surface pressure distribution


for flight conditions FC43 and FC68. Flight condition FC43
(cpcritical  0.4064) shows indications of both a deflected elevon (butt
line stations BL70 and BL95; locally subsonic oncoming flow) and a
deflected leading-edge flap (butt line stations BL153.5 and BL184.5;
similar to FC68). No indications of a deflected aileron were observed.
D. Summary

The preceding observations are summarized in Table 2. In this


table, for each flight condition, the control surface deflections as
obtained from the preceding analysis are indicated as either not
deflected (ND), deflected (D) or unclear (?). From this table, it
follows that flight condition FC79 is the most likely flight condition
from the selected transonic flight conditions to be flown without any
control surface deflections. It should be noted from the nominal flight
conditions shown in Table 1 that this flight condition represents a
wind-up turn (WUT) at a load of 3.7g units. Flight condition FC79
was chosen to be used for further CFD analysis of the transonic flow
on the F-16XL aircraft within CAWAPI-2.

solving the Euler and Reynolds-averaged NavierStokes (RANS)


equations on multiblock structured grids for arbitrary configurations.
The configuration can be either fixed or moving, relative to an inertial
reference frame, and can be rigid or flexible. The compressible equations in full conservation form are discretized in space by a secondorder-accurate cell-centered finite-volume method, using multiblock
structured grids, central differences, and matrix artificial dissipation.
The artificial dissipation consists of a blending of second- and fourthorder differences with a Jameson-type shock sensor for the basic flow
equations and a total variation diminishing discontinuity sensor for
the turbulence model equations.
For steady flow simulations, the discretized time-dependent
system of equations is integrated toward the steady-state solution
using a five-stage explicit RungeKutta scheme. Local time stepping
and multigrid acceleration techniques are applied. For time-accurate
simulations, the flow solver uses the dual time stepping scheme,
where for each time step, the time-dependent flow equations are
integrated in pseudotime toward a steady-state solution in a similar
way as in the steady flow simulation using the same acceleration
techniques.
Several turbulence models are present in the flow solver ENSOLV,
including the turbulent nonturbulent (TNT) k model [1315], the
explicit algebraic Reynolds stress model (EARSM) based on the
TNT k model, and a hybrid RANS/large-eddy simulation model
for extra-large-eddy simulation [16,17].
The structured (multiblock) grid has been generated using NLRs
Cartesian grid mapping technique [2,5,18]. The (semiautomatic) grid
generation algorithms have been developed at NLR and are part of
NLRs ENFLOW flow simulation system [12].
B. Details of Simulations

IV.

CFD Analysis

A. CFD Method

The flow solver ENSOLV, which is part of NLRs in-housedeveloped flow simulation system ENFLOW [12], is capable of

The (multiblock) structured grid generated for CAWAPI [2,5] has


also been used in CAWAPI-2 for consistency. This grid around the
half-span full-scale model of the F-16XL aircraft consists of 1903
blocks and 14,750,720 grid cells. Further details on this grid can be
found in [5].

Downloaded by JOHNS HOPKINS UNIVERSITY on November 15, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.C033211

Article in Advance

/ BOELENS

Fig. 5 Flight-test sectional surface pressure coefficient along butt line stations (see Fig. 6) for the transonic flight conditions.

All simulations were performed as steady flow simulations in fully


turbulent mode. A full multigrid scheme (grid sequencing) was used
to simulate the flow on the three grid levels using the TNT k
turbulence model to obtain an initial solution. Next, the solution on
the finest grid level was computed using the EARSM model. The
number of iterations used on each grid level equals 5000. The full
approximation storage (FAS) multigrid scheme is used to compute
the solution on a specific grid level. One FAS multigrid level was
used. Three orders of convergence were obtained for the root mean
square norms.
The boundary conditions are the same as those reported in [2],
except for the boundary conditions at the engine inlet and the engine
exit. Because engine data for the transonic conditions other than
flight condition FC70 (see [2]) were not available, a simple through-

Table 2 Summary of control surface deflections


as obtained from analysis of sectional surface pressure
distributions of transonic flight conditions
Flight condition Aileron Elevon Leading-edge flap
FC68
ND
D
D
FC69
D
ND
ND
FC70
D
ND
D
FC43
ND
D
D
FC79
ND
ND
ND
FC80
ND
D
D
FC81
D
?
D

Table 3

flow boundary condition was used on these planes during CAWAPI2. Note that, in [8,9], no significant effect of different boundary
condition formulations for the engine inlet or exit has been observed.
C. Assessment of Results

Simulations have been performed for the actual flow conditions


of flight conditions FC70 and FC79 (see Table 3), however, with
the sideslip angle set to zero. Figure 6 shows the simulated
surface pressure coefficient for flight conditions FC70 and FC79.
For a detailed description of the flow phenomena observed for
these flight conditions, the reader is referred to [2]. The black
lines indicate the butt line stations along which flight-test pressure
measurements were performed. A comparison between the
measured and simulated surface pressure coefficients along these
lines is shown in Fig. 7.
Regarding these results, the following can be remarked:
1) Whereas flight condition FC70 was characterized by a weak
inner wing primary vortex [8], flight condition FC79 is characterized
by a very distinct inner wing primary vortex. Though the grid used
was insufficient to resolve the weak inner wing primary vortex (and
the associated shock-vortex interaction) for flight condition FC70
[8], this vortex is well resolved for the high-angle-of-attack flight
condition FC79. At butt line station BL55, the suction peak associated with this vortex is slightly underpredicted (probably due to
insufficient grid resolution in the leading-edge region of butt line
station BL55), but at all the other stations inboard of the air dam and
the actuator pod, the agreement between the flight-test data and the
simulation is good.

Transonic flight conditions (actual) in the F-16XL database used for


analysis

Flight condition , deg , deg


M
Re 106
FC70
4.366 0.310 0.969
88.7
FC79
9.521 0.642 0.901
87.8

Load, g units Flight/run no.


1
1525b
3.7 WUT
1465d

Downloaded by JOHNS HOPKINS UNIVERSITY on November 15, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.C033211

Article in Advance

/ BOELENS

Fig. 6 Surface pressure coefficient for flight conditions FC70 and FC79. Black lines indicate butt line stations along which flight-test pressure
measurements were performed.

Fig. 7 Flight-test and simulated surface pressure coefficient along butt lines (see Fig. 6) for the transonic flight conditions FC70 and FC79.

Downloaded by JOHNS HOPKINS UNIVERSITY on November 15, 2016 | http://arc.aiaa.org | DOI: 10.2514/1.C033211

Article in Advance

2) For flight condition FC79, both the flight test and the simulations show a supersonic region (cpcritical  0.1878) at about 75% of
the chord on the part of the wing inboard of the air dam and the
actuator pod. The surface pressure distributions in this region (see
especially butt line stations BL55 and BL70) and aft of it agree well
between the flight-test data and the simulations. This supersonic
region seems to decelerate smoothly and is not terminated by a shock
wave. This delicate flow feature is known to be difficult to model.
3) On the part of the wing outboard of the air dam and the actuator
pod flight, condition FC79 shows a much better agreement between
flight-test data and simulation than flight condition FC70. At butt line
station BL153.5, the qualitative agreement is good, whereas at butt
line station BL184.5, the results also agree well on a quantitative
base. Figure 6 shows that, in between butt line station BL153.5 and
the air dam and actuator pod, the pressure coefficient displays a
(steep) spanwise pressure gradient toward lower pressures. Changes
in the flight conditions (for example, in the sideslip angle) will have a
large effect on the local distribution of surface pressure coefficient in
this region. Therefore, the sensitivity of the simulations to such
changes needs to be further investigated.
4) The results confirm that flight condition FC79 has been flown
without any control surface deflections (see also Sec. III). Therefore,
this flight condition constitutes a much better defined transonic test
case than flight condition FC70 and should be used for further
transonic CFD analysis.

V.

Conclusions

In the framework of CAWAPI-2, NLR has performed an analysis of


a set of transonic flight conditions available in the Cranked-Arrow
Wing Aerodynamics Project (CAWAP) database. This simple analysis,
based on scrutinizing the flight-test surface pressure measurements for
indications of control surface deflections, provided a lot of insight in
the otherwise unknown actual state of the control surfaces during the
flight test (i.e., whether a control surface was deflected or not).
Computational fluid dynamics (CFD) simulations were used to
confirm the outcome of the analysis, showing that flight FC79 is a
much better candidate for further CFD analysis than flight condition
FC70.
The outcome of the CFD analysis of flight conditions FC70 and
FC79 also has provided the steps that need to be taken in follow-on
transonic F-16XL CFD investigations:
1) Improve the grid in critical regions, for example, near the wing
leading edge where the inner wing primary vortex is formed and on the
part of the wing outboard of the air dam. The grid used in CrankedArrow Wing Aerodynamics Project International (CAWAPI), and also
for consistency in CAWAPI-2, was generated using a Cartesian grid
mapping technique that had become available just before the start of
the CAWAPI project [2,5]. In the meantime, these grid generation
methods have been improved so that higher-quality grids can be
generated now.
2) Investigate the sensitivity of the simulation with respect to
changes in flow conditions (Mach number, angle of attack, sideslip
angle, etc.). For such an analysis, the steady-state approach employing the explicit algebraic stress model adopted in CAWAPI and
CAWAPI-2 suffices, because such an approach will already uncover
the trends.
3) Perform unsteady flow simulations (also using more advanced
turbulence models). Tuft images have, for example, shown that the
flow outboard of the air dam and the actuator pod [2] is unsteady. To
more realistically resolve the flow in such regions, the introduction of
unsteady effects is required.
Finally, CAWAPI-2 has provided NLR with an excellent platform to
evaluate its ENFLOW flow simulation system against the high-quality
flight measurements contained in the unique CAWAP database.

Acknowledgments
This work has been conducted under National Aerospace Laboratorys programmatic research funding Kennis voor Vermogen. The

/ BOELENS

author would like to thank all members of the Cranked-Arrow Wing


Aerodynamics Project International 2 and, in particular, James M.
Luckring of NASA Langley for the interesting and fruitful discussions during the course of this work.

References
[1] Lamar, J. E., Obara, C. J., Fisher, B. D., and Fisher, D. F., Flight, WindTunnel, and Computational Fluid Dynamics Comparison for Cranked
Arrow Wing (F-16XL-1) at Subsonic and Transonic Speeds, NASA
TP-2001-210629, Feb. 2001.
[2] Boelens, O. J., Badcock, K. J., Elmiligui, A., Abdol-Hamid, K. S., and
Massey, S. J., Comparison of Measured and Block Structured Simulations for the F-16XL Aircraft, Journal of Aircraft, Vol. 46, No. 2, 2009,
pp. 377384.
doi:10.2514/1.35064
[3] Lamar, J. E., Cranked Arrow Wing (F-16XL-1) Flight Flow Physics
with CFD Predictions at Subsonic and Transonic Speeds, NATO
Research and Technology Organisation MP-069, 2003, Paper 44.
[4] Obara, C. J., and Lamar, J. E., Overview of the Cranked-Arrow Wing
Aerodynamics Project International, Journal of Aircraft, Vol. 46, No. 2,
2009, pp. 355368.
doi:10.2514/1.34957
[5] Boelens, O. J., Badcock, K. J., Grtz, S., Morton, S., Fritz, W., Karman,
S. L., Jr., Michal, T., and Lamar, J. E., Description of the F-16XL
Geometry and Computational Grids Used in CAWAP, Journal of
Aircraft, Vol. 46, No. 2, 2009, pp. 369376.
[6] Grtz, S., Jirsek, A., Morton, S. A., McDaniel, D. R., Cummings, R.
M., Lamar, J. E., and Abdol-Hamid, K. S., Standard Unstructured Grid
Solutions for CAWAPI F-16XL, Journal of Aircraft, Vol. 46, No. 2,
2009, pp. 385408.
[7] Fritz, W., Davis, M. B., Karman, S. L., and Michal, T., RANS Solutions
for the CAWAPI F-16XL Using Different Hybrid Grids, Journal of
Aircraft, Vol. 46, No. 2, 2009, pp. 409422.
[8] Rizzi, A., Jirsek, A., Lamar, J. E., Crippa, S., Badcock, K. J., and
Boelens, O. J., Lessons Learned from Numerical Simulations of the
F-16XL Aircraft at Flight Conditions, Journal of Aircraft, Vol. 46,
No. 2, 2009, pp. 423441.
doi:10.2514/1.35698
[9] Rizzi, A., Jirsek, A., Lamar, J. E., Badcock, K. J., Boelens, O. J., and
Crippa, S., What was Learned from Numerical Simulations of F-16XL
(CAWAPI) at Flight Conditions, NATO RTO-TR-AVT-113 AC/323
(AVT-113)TP/246, Oct. 2009.
[10] Stachowiak, S. J., and Bosworth, J. T., Flight Test Results for the F16XL with a Digital Flight Control System, NASA TP-2004-212046,
2004.
[11] Berglind, T., Brunet, V., Caballero Rubiato, V., Ceresola, N., Heinrich,
R., Leicher, S., and Prananta, B. B., Computations of Unsteady Aerodynamics Due to Body Motion, CEAS European Air and Space
Conference; Century Perspectives, German Soc. for Aeronautics and
Astronautics Paper CEAS-2007-201, Sept. 2007.
[12] Boerstoel, J. W., Kassies, A., Kok, J. C., and Spekreijse, S. P.,
ENFLOW, a Full-Functionality System of CFD Codes for Industrial
Euler/.Navier-Stokes Flow Computations, National Aerospace Lab.,
NLR-TP-96286U, Amsterdam, The Netherlands, 1996.
[13] Brandsma, F. J., Kok, J. C., Dol, H. S., and Elsenaar, A., Leading Edge
Vortex Flow Computations and Comparison with DNW-HST Wind
Tunnel Data, National Aerospace Lab.,
NLR-TP-2001-238,
Amsterdam, The Netherlands, 2001.
[14] Dol, H. S., Kok, J. C., and Oskam, B., Turbulence Modeling for
Leading-Edge Vortex Flows, AIAA Paper 2002-0843, 2002.
[15] Kok, J. C., Resolving the Dependence on Free-Stream Values for the
k Turbulence Model, National Aerospace Lab., NLR-TP-99295,
Amsterdam, The Netherlands, 1999.
[16] Kok, J. C., Dol, H. S., Oskam, B., and van der Ven, H., Extra-Large
Eddy Simulation of Massively Separated Flows, National Aerospace
Lab., NLR-TP-2003-200, Amsterdam, The Netherlands, 2003.
[17] Kok, J. C., Soemarwoto, B. I., and van der Ven, H., X-LES
Simulations Using a High Order Finite-Volume Scheme, Notes
on Numerical Fluid Mechanics and Multidisciplinary Design,
Vol. 97, edited by Peng, S. H., and Haase, W. (eds.), Advances in
Hybrid RANSLES Modelling, SpringerVerlag, Berlin, 2008,
pp. 8796.
[18] Boelens, O. J., CFD Analysis of the Flow Around the X-31 Aircraft at
High Angle of Attack, AIAA Paper 2009-3628, 2009.

Das könnte Ihnen auch gefallen