You are on page 1of 14

JOURNAL OF AIRCRAFT

Vol. 42, No. 5, September–October 2005

Data Summary from Second AIAA Computational Fluid


Dynamics Drag Prediction Workshop

Kelly R. Laflin,∗ Steven M. Klausmeyer,† and Thomas Zickuhr‡


Cessna Aircraft Company, Wichita, Kansas 67218
John C. Vassberg§
The Boeing Company, Huntington Beach, California 92647
Richard A. Wahls¶ and Joseph H. Morrison∗∗
NASA Langley Research Center, Hampton, Virginia 23681
Olaf P. Brodersen†† and Mark E. Rakowitz††
DLR, German Aerospace Center, 38108 Brunswick, Germany
Edward N. Tinoco‡‡
The Boeing Company, Seattle, Washington 98124
and
Jean-Luc Godard††
ONERA, 92322 Chatillon Cedex, France

Results from the Second AIAA Drag Prediction Workshop are summarized. The workshop focused on absolute
and configuration delta drag prediction of the DLR, German Aerospace Research Center F6 geometry, which is
representative of transport aircraft designed for transonic flight. Both wing–body and wing–body–nacelle–pylon
configurations are considered. Comparisons are made using industry relevant test cases that include single-point
conditions, drag polars, and drag-rise curves. Drag, lift, and pitching moment predictions from several different
Reynolds averaged Navier–Stokes computational fluid dynamics codes are presented and compared to experimental
data. Solutions on multiblock structured, unstructured, and overset structured grids using a variety of turbulence
models are considered. Results of a grid-refinement study and a comparison of tripped transition vs fully turbulent
boundary-layer computations are reported.

Nomenclature c = mean aerodynamic chord


AR = aspect ratio i = solution index
CD = drag coefficient M = Mach number
CDP = idealized profile drag coefficient N = number of grid nodes
C D Pressure = drag coefficient pressure component NWB = number of grid nodes of wing–body (WB) grid
C D Viscous = drag coefficient viscous component NWBNP = number of grid nodes of wing–body–nacelle–pylon
Cf = skin-friction coefficient (WBNP) grid
CL = lift coefficient Re = Reynolds number
C Lα = lift curve slope Rec = Reynolds Number based on c
CM = pitching moment coefficient S = wing reference area
CP = pressure coefficient y+ = wall distance
α = angle of incidence
Presented as Paper 2004-0555 at the AIAA 42nd Aerospace Sciences  = difference in quantity
Meeting and Exhibition, Reno, NV, 5–8 January 2004; received 10 May η = fraction of wing semispan
2004; revision received 20 September 2004; accepted for publication σ = standard deviation
5 October 2004. Copyright  c 2004 by the authors. Published by the Ameri-
can Institute of Aeronautics and Astronautics, Inc., with permission. Copies Subscripts
of this paper may be made for personal or internal use, on condition that the
copier pay the $10.00 per-copy fee to the Copyright Clearance Center, Inc.,
bub = separation bubble
222 Rosewood Drive, Danvers, MA 01923; include the code 0021-8669/05 des = design flight condition
$10.00 in correspondence with the CCC. eyeb = eye separation bubble feature on body
∗ Specialist Engineer, Department of Aerodynamic and Product Analysis. eyew = eye separation bubble feature on wing
Member AIAA. ∞ = freestream condition
† Specialist Engineer, Department of Aerodynamic and Product Analysis.
Senior Member AIAA.
‡ Senior Specialist Engineer, Department of Aerodynamic and Product Introduction
Analysis. Senior Member AIAA.
§ Boeing Technical Fellow, Phantom Works, Department of Flight Sci-
ences and Advanced Design; currently Chairman Drag Prediction Workshop-
T HE Drag Prediction Workshop (DPW) series began with the
First AIAA Drag Prediction Workshop1 (DPW-I), which was
organized by members of a working group of the AIAA Applied
II Organizing Committee. Associate Fellow AIAA. Aerodynamics Technical Committee in an effort to assess the state
¶ Assistant Head, Configuration Aerodynamics Branch. Associate Fellow
of the art of computational fluid dynamics (CFD) drag prediction in
AIAA.
∗∗ Research Scientist, Department of Computational Aerosciences. Senior the aircraft industry. The premise of DPW-I was to solicit CFD pre-
Member AIAA. dictions of a common, industry relevant geometry and assess the re-
†† Research Engineer, Institute of Aerodynamics and Flow Technology. sults using statistical analysis techniques. Although the focus of the
‡‡ Boeing Technical Fellow, Department of Enabling Technology and Re- workshop was on drag prediction, lift and pitching moment predic-
search. Associate Fellow AIAA. tions were also evaluated. The DLR, German Aerospace Research
1165
1166 LAFLIN ET AL.

Center- (DLR-) F4 wing–body configuration was chosen as the sub- These comments guided the organizing committee to choose the
ject of DPW-I because of both its simplicity and the availability of DLR-F6 wing–body (WB) and wing–body–nacelle–pylon (WBNP)
publicly released experimental test data.2 The workshop commit- configurations. Test cases were chosen to focus on absolute drag and
tee provided a standard set of multiblock structured, unstructured, component drag increment prediction accuracy. Each participant
and overset structured grids for the DLR-F4 geometry to encour- was asked to perform grid-refinement studies of both configurations
age participation in the workshop and reduce variability in the CFD at a fixed-C L single-point condition and to submit computed drag
results. However, participants were encouraged to construct their polars for both configurations. Additionally, participants were en-
own best practices grids, so that learned knowledge concerning grid couraged to submit computed drag-rise curves and perform specified
generation and drag prediction might be shared among the work- tripped transition and fully turbulent boundary-layer computations
shop attendees. The test cases were chosen to reflect the interests for both configurations.
of industry and included a fixed-C L single-point solution, drag po- DPW-II was sponsored by the AIAA Applied Aerodynamics
lar, and constant-C L drag-rise data sets. There were 18 participants Technical Committee and held 21–22 June 2003 in conjunction
who submitted results using 14 different CFD codes; many submit- with the 21st AIAA Applied Aerodynamics Conference in Orlando,
ted multiple sets of data exercising different options in their codes, Florida. The DPW series has spawned considerable interest in the
for example, turbulence models and/or different grids. Because of CFD and applied aerodynamic communities, and this summary pa-
strong participation, DPW-I successfully amassed a CFD data set per is one of several DPW-II papers.9−23 Approximately 75 persons
suitable for statistical analysis.3 However, the results of that analysis from five continents representing academia, research laboratories,
were disappointing, showing, for example, a 270 drag count, that and industry attended the workshop. Of those attending, 25 partici-
is, drag coefficient unit = 0.0001, spread in the fixed-C L data, with pated in the workshop, submitting data for analysis. Those submit-
a 100:1 confidence interval of more than ±50 drag counts. ting data were given the opportunity to present their results. The
Despite the disheartening results of the statistical analysis, DPW-I workshop data and presentations may be viewed at the DPW-II
was a definitive success. It brought together CFD developers and website.24
practitioners and focused their efforts on a common problem. It fa- In the following sections, a description of the DLR-F6 geom-
cilitated an exchange of learned best practices and promoted open etry and the wind-tunnel experimental approach is included. The
discussions, identifying areas requiring further research or addi- standard multiblock structured, unstructured, and overset structured
tional scrutiny. Possibly most significantly, it employed statistical grids, provided to participants by the DPW-II organizing committee,
methods to assess CFD results objectively. Finally, it reminded the are briefly described. Test cases are defined, and an overview of the
CFD and applied aerodynamics communities that CFD is not yet a participation and submitted data is presented. The DPW-II results
fully mature discipline. are summarized for each of the four test cases, which include a grid-
In addition to the accomplishments just listed, DPW-I created a refinement study and configuration delta drag comparisons. Lift,
sustained interest in industry relevant drag prediction. Several of the drag, and pitching moment characteristics for both the DLR-F6 WB
participants presented their DPW-I results at a well-attended special and WBNP configurations are discussed. Drag data are compared
session of the 2002 AIAA Aerospace Sciences Meeting and Exhibit against grid size, grid type, turbulence model, and boundary-layer
in Reno, Nevada.3−8 transition specification. Finally, conclusions of DPW-II and recom-
The interest generated by DPW-I naturally led to the planning mendations for future drag prediction workshops are presented.
and organization of a Second AIAA Drag Prediction Workshop
(DPW-II). The DPW-II organizing committee, recognizing the suc- Geometry Description
cess of DPW-I, maintained the DPW-I objectives for DPW-II. These The configuration selected for DPW-II is a wing-mounted twin-
objectives are as follows. engine aircraft representing a typical modern wide-body transport.
1) State-of-the-art computational methods are to be assessed as This configuration, called DLR-F6, was designed by DLR more
practical aerodynamic tools for aircraft force and moment prediction than 20 years ago and is derived from the DLR-F4 configuration,25
of industry relevant geometries, focusing on drag prediction. which was the subject of DPW-I (Ref. 3). The DLR-F6 fuselage
2) An impartial forum is to be provided for evaluating the effec- remained unchanged from the DLR-F4 configuration. However, the
tiveness of existing computer codes and modeling techniques using outer three of the four wing defining airfoils were modified. The
Navier–Stokes solvers. airfoil at the wing kink is translated upward to achieve a smoother
3) Areas needing additional research and development are to be upper wing surface. The twist distribution is changed, and the span-
openly discussed and identified. Furthermore, the DPW-I format wise position of section three was moved from η = 0.7 to η = 0.84.
was retained, consisting of following key elements. The intention of these changes was to achieve a more elliptic lift
The first element is a common subject geometry, having publicly distribution to reduce or eliminate boundary-layer separation at the
available experimental data, that is simple enough to do high-quality rear upper wing surface of the DLR-F6 wing as compared to the
computations and that is relevant to industry. DLR-F4 wing; however, the goal of eliminating the trailing-edge
The second element consists of required and optional industry separation was not achieved. A list of the DLR-F6 quantities is
relevant test cases. These included a fixed-C L single-point condition, given in Table 1.
drag polars, and constant-C L drag-rise curves. Between 1990 and 1998 several measurement campaigns were
The third element is a standard set of provided grids to encourage carried out in the ONERA-S2MA wind tunnel to investigate the
participants and reduce variability in the CFD results. influence of different flow through nacelles on the DLR-F6 wing
The fourth key element is a rigorous statistical analysis of the flowfield. A model of 1.1713-m span was manufactured by DLR for
CFD results to establish confidence levels in the data.
The fifth element consists of scheduled open forum sessions Table 1 DLR-F6 quantities
to encourage discussion and interaction among participants and
attendees. Property Value
The DPW-II organizing committee relied heavily on discussions S/2 72,700.0 mm2
that took place during DPW-I in identifying what geometry and test c 141.2 mm
cases to use for DPW-II. These discussions focused on poor grid b/2 585.647 mm
quality/resolution and inconsistent boundary-layer transition spec- AR 9.5
ification, that is, tripped transition vs fully turbulent calculations, X ref 157.9 mm
as possible causes for the large scatter in the DPW-I results. Addi- Yref 0.0 mm
Z ref −33.92 mm
tionally, many attendees believed that a second workshop should
Mdes 0.75
focus on a more complex model. It was also hypothesized that C L des 0.5
CFD better predicts configuration delta drag than absolute drag Rec 3 × 106
values.
LAFLIN ET AL. 1167

the wind-tunnel tests. At this scale, the aerodynamic mean chord is


0.1412 m, and the reference area is 0.1453 m2 . The main geomet-
rical characteristics of the wing are 9.5 aspect ratio, 0.3 taper ratio,
27.1-deg leading-edge sweep angle, 25-deg quarter-chord sweep an-
gle, and 4.8-deg dihedral angle. The shape of the fuselage is defined
by a forebody, including a cockpit, a cylindrical part containing the
wing, and an afterbody. Design cruise conditions for DLR-F6 are
Mdes = 0.75 and C L des = 0.50.
Two variants of the DLR-F6 configuration are considered for
DPW-II: a WB configuration and a WBNP configuration. The
WBNP propulsion units are simulated by flow through nacelles. The
nacelles have axisymmetric shapes derived from the CFM56 long-
duct nacelles and are modified to achieve, in cruise conditions, a a) b)
similar mass flow coefficient as the real engine. The nacelle position
under the wing was selected to present significant engine/airframe Fig. 2 DLR-F6 experimental total force and moment measurements,
interference effects. The pylon is of symmetrical shape. M∞ = 0.75.
For DPW-II, a new CAD model was built based on the original
production coordinates and drawings. Great care was taken to re-
tain the original geometry and to account for wing deflection due
to aerodynamic load, while achieving smooth curvature of the sur-
faces. The wing aeroelastic effects include bend and twist, which
were derived by a finite element analysis. The required wing load-
ing was obtained from an unstructured TAU26 computation of the
DLR-F6 WB configuration at the design point using an old CAD
model of the F6. The wing tip is deflected 4 mm upward and has
a twist of −0.314 deg. Before the release of the geometry, struc-
tured grid computations of the WB configuration were carried out
using the FLOWer27 software for the geometries with and without
aerodynamic wing deflection. The influence of the deflected wing
shape on the pressure distribution was found to be small and limited
to the outer wing; the comparison of the computed pressures to the
experimental wing pressure distribution was good.
Fig. 3 DLR-F6 experimental wing section pressure profiles, M∞ = 0.75
Experimental Approach and CL = 0.5.
DLR-F6 wind-tunnel tests were performed in August 1990 in a based on the mean aerodynamic chord was kept constant and equal
cooperation between ONERA and DLR,28 in the S2MA pressurized to Re = 3 × 106 . The transition was fixed on the fuselage, wing, and
wind tunnel of the ONERA center in Modane.29 The transonic test nacelle by transition strips.
section has a height of 1.77 m and a width of 1.75 m. To have reduced Forces and moments were measured by the balance during con-
wall influence, the top and bottom walls are perforated. Two types of tinuous variations of incidence for a given Mach number and were
measurements were performed in these tests: total force and moment then interpolated vs C L and Mach number. The pressure distribu-
measurements using a six-component balance installed inside the tions were measured for a fixed Mach number and incidence. Forces
fuselage of the model and static pressure measurements over the and moments were corrected first for nonhomogeneous flow in the
surface of the model. wind tunnel and for wall and model support influence. The overall
The objective of these tests was the precise determination of drag. force and moment repeatability was checked for Mdes , and the scat-
The model was supported by a fin sting through the balance (Fig. 1). ter for C D was found to be less than ±1 drag count and less than
For a detailed analysis of the flow, the model is instrumented with ±0.01 deg for α.
344 pressure taps in 8 spanwise sections of the wing, in 3 radial Engine installation drag was also determined by subtracting the
sections of the nacelles and on the pylon. WB configuration and nacelle internal drag from the WBNP con-
The tests were carried out for the WBNP and WB configurations figuration. The nacelle internal drag was determined through cali-
for Mach numbers from 0.6 to 0.82 and sweeps of incidence corre- brations of the nacelles alone in a specific test bench.
sponding to lift coefficients from 0.00 to 0.60. The Reynolds number Some examples of measurements used for DPW-II comparisons
are shown in Figs. 2 and 3. The configuration with nacelles has,
as expected, a higher drag and a higher incidence at the same lift
compared to the WB configuration. The installation drag (Fig. 2b)
is quite large due to the location of the nacelles close to the wing.
Figure 3 shows wing section pressure distributions inboard and out-
board of the nacelle location for the two configurations at Mdes and
the cruise incidence of the WBNP configuration. The engine instal-
lation effects are clearly identified: On the lower side inboard of
the nacelle, a strong acceleration of the flow is visible, followed by
a shock wave, and on the upper side outboard of the nacelle, the
velocities are lower in the supersonic region.
These tests have produced high-quality measurements, in terms
of total forces and static pressures on the model, with a detailed
description of the engine installation effects. These measurements
constitute a good experimental database for the numerical assess-
ment of the drag of a civil transport aircraft configuration.

Standard Grids
To minimize variation in the results and encourage participation,
Fig. 1 DLR-F6 model in ONERA S2MA wind tunnel. a set of standard grids was generated for both the DLR-F6 WB
1168 LAFLIN ET AL.

Table 2 Medium density grid construction specifications Table 3 Multiblock structured grid sizes

Specification Value ICEM ZEUS


DLR-F6
Wing leading-edge chordwise spacing ∼0.1% Local chord grids NWB × 10−6 NWBNP × 10−6 NWB × 10−6 NWBNP × 10−6
Wing trailing-edge chordwise spacing ∼0.1% Local chord
Cells on blunt trailing edge 8 Coarse 3.0 4.6 2.1 ——
Wing spanwise spacing at root ∼1% Semispan Medium 5.5 8.5 3.9 6.2
Wing spanwise spacing at tip ∼0.1% Semispan Fine 10.0 13.7 8.9 8.7
Spacing at fuselage nose ∼2.0% Mean aerodynamic chord Finer —— —— 13.2 ——
y + (first grid cell distance) 0.001 mm
Table 4 Unstructured grid sizes (nodes, cells) VGRIDns
and WBNP configurations and made available for download via Node centered Cell centered
DLR-F6
the DPW website.24 Initial graphics exchange specification defini-
grids NWB × 10−6 NWBNP × 10−6 NWB × 10−6 NWBNP × 10−6
tions of the geometries were also made available for participants
who wished to construct their own grids. A grid-refinement study Coarse 1.1 1.8 1.4 2.2
was planned for the workshop, and so a series of three grids with Medium 3.0 4.8 3.9 5.9
varying grid density (coarse, medium, and fine) was constructed for Fine 9.1 10.3 11.3 16.8
each configuration. Multiblock structured, unstructured, and overset
structured grids were all provided. Table 5 Unstructured grid sizes ICEM and TAU plus Centaur
Guidelines as to the construction of the grids were published to
minimize grid density variations among the different grid types. ICEM TAU plus Centaur
DLR-F6
The medium grid was sized to be sufficient for industry drag predic- grids NWB × 10−6 NWBNP × 10−6 NWB × 10−6 NWBNP × 10−6
tion. Coarse and fine density grids were obtained from the medium
density grid through global coarsening and refining, respectively. Coarse 0.6 1.1 3.8 5.5
Key grid construction specifications for the medium density grid Medium 2.0 2.3 5.8 8.2
are listed in Table 2. Fine 5.7 5.9 8.8 12.3
It was desired that the total number of grid nodes grow by approxi-
mately a factor of three between the coarse and medium density grids DLR-F6 WB configuration grid-refinement study; however, they
and between the medium and fine density grids. This corresponds to were not globally successively refined. The refinement focused on
a growth factor, for structured meshes, of approximately 1.5 in each the region surrounding the wing. Only two sets of grids were gen-
grid dimension. The coarse and fine density grids were generated erated for the DLR-F6 WBNP configuration; the fine grid density
using this scaling guideline. The boundary-layer stretching ratio in is at least double the medium grid density in the spanwise (circum-
the direction normal to all viscous surfaces was specified to be less ferential direction) in the blocks surrounding the strut/nacelle. Grid
than 1.25, and the far-field boundary distance was specified to be sizes are listed in Table 3.
approximately 100 mean aerodynamic chord from the aircraft for
all three grids. Unstructured Grids
The described guidelines were used to construct grids for three VGRIDns
different grid types: multiblocked structured, unstructured, and over- Two sets of tetrahedral grids were generated for the current study:
set structured. However, not all supplied grids met these guidelines. grids for cell-centered solvers with wall-functions and grids for
In some cases, more than one set of coarse, medium, and fine grids node-centered solvers with integration to the wall. Additionally,
were generated for a particular grid type. For example, unstructured for the grid-convergence study with each solver type, two fam-
grids for both node-based and cell-based numerical schemes were ilies of coarse, medium, and fine grids were generated for the
generated. A brief description of each of the grids made available WB configuration and the WBNP configuration. All of the grids
by the DPW-II organizing committee is given subsequently. Images were generated with the VGRIDns advancing-layer and advancing-
of the grids and further details may be found within the DPW-II front grid-generation software package.33,34 The grids generated
presentations, posted on the DPW website.24 with VGRIDns were fully tetrahedral. However, VGRIDns used
an advancing-layer technique to generate the boundary-layer por-
Multiblock Structured Grids tion of the grid so that prisms can be reconstructed in the boundary
ICEM layer. In the boundary layer, two tetrahedral cells were combined
A set of multiblock structured grids with 1-to-1 point matching at to make up one prism. The different grids for each configuration
the block boundaries was produced by ICEMCFD using their HEXA were generated by a global coarsening/refinement of the inviscid
software package. Coarse, medium, and fine grids were generated spacing parameters and a global coarsening/refinement of the vis-
for both the WB and WBNP configurations. A summary of grid sizes cous wall spacing and stretching factors. Additional sources were
is given in Table 3. A growth factor of less than two was used between included in the WBNP grids to accommodate the nacelle and pylon.
grid levels. For both geometries, a C topology was used around the A comparison of the grid sizes for all 12 grids is given in Table 4.
wing to maintain viscous spacing in the wing wake. A total of eight
workshop participants utilized the ICEM structured grids, and as ICEM
such, the grids were exposed to a wide variety of flow solvers, both A set of unstructured mixed-element ICEM grids was also sup-
structured and unstructured. Before release, the grids were tested plied, composed of tetrahedra and pyramids with a thin layer of
with CFL3D, which exhibited satisfactory convergence. During the prismatic elements in the boundary layer. Grid sizes are given in
workshop, it was noted that the chordwise grid spacing on the wing Table 5.
for the medium and fine WBNP grids was nearly identical, a feature
that unfortunately slipped through the prerelease evaluation. TAU Plus Centaur
A set of grids was generated using the Centaur grid generation and
Zeus TAU solution-based grid-adaptation software packages.35,36 Cen-
A second set of 1-to-1 point-matched multiblock structured grids taur produces grids consisting of prismatic, tetrahedral, and pyra-
was generated using the Boeing Zeus and Aero Grid and Paneling midal elements. The triangles on the wing surface were stretched
system.30−32 Zeus grids used an H–H grid topology for the WB moderately in the spanwise direction by a factor of approximately
configuration. Additional H–H and H–O topology grids were used three. The resolution of the wake of the wing was improved down-
when the wing-mounted nacelle and pylon were added for the DLR- stream by introducing small tetrahedral elements. In addition to the
F6 WBNP configuration. Four sets of grids were generated for the gridding guidelines, nodes were slightly clustered in regions where
LAFLIN ET AL. 1169

Table 6 Overset grid sizes (total nodes, compare nonblanked nodes) The drag polar, case 2, is typical of how industry collects and ex-
amines wind-tunnel experimental data. Participants were requested
DLR-F6 NWB × 10−6 NWBNP × 10−6 to trip the wing boundary layer, that is, specify the position where
Coarse 1.9 3.1 the boundary layer is forced to transition from laminar to turbulent,
Medium 6.8 10.7 at experimental trip locations. Fully turbulent computations were
Fine 23.1 35.8 accepted if a trip position could not be specified using a partic-
ular Navier–Stokes code. Case 2 results were requested to assess
it is likely that shocks occur. An attempt was made to achieve a the effects of grid type, turbulence model, boundary-layer transi-
smooth transition between the prismatic and tetrahedral elements. tion specification, and grid size on computed lift, drag, and pitching
The solution-based grid adaptation of TAU, using pressure and moment.
velocity gradients, has been applied successfully in the past for Optional case 3 was requested to compare the difference in con-
these kinds of configurations and flow conditions and, therefore, figuration delta drags when simpler fully turbulent boundary-layer
was used here to generate the three grid resolutions for each specification is used instead of the more physically correct tripped
configuration.36−38 The grid densities are listed in Table 5. transition boundary-layer specification.
Constant-C L drag-rise curves, case 4, are often used by industry
Overset Structured Grids to examine experimental results and make configuration decisions.
The overset structured grids for the DLR-F6 WB and WBNP con- This case was chosen to compare predicted absolute drag and con-
figurations were developed per specifications defined by the DPW-II figuration delta drag-rise curves to experimental data.
organizing committee. The first specification stated that the medium
grids should be representative of (or slightly more dense than) those Overview of Methods and Data Submitted
currently being used within industry for engineering-level accurate, Data from 22 different Navier–Stokes codes were submitted by
CFD-based drag predictions. To accomplish this, the Boeing Phan- 25 participants. Many participants submitted more than one set of
tom Works team created a 12-zone grid system for the WB configu- results, exercising different options in their codes, for example, tur-
ration that consisted of about 6.8 million grid nodes, then augmented bulence models and/or using different grids. All submitted data sets
this system with an additional 11 blocks to create a 23-zone grid can be found in their entirety at the DPW website.24
system for the WBNP configuration comprising nearly 10.7 million At the conclusion of DPW-II, participants were asked to complete,
grid nodes. A detailed description of these overset grids is provided recheck, and resubmit their data for inclusion in this summary pa-
in Ref. 39. The corresponding coarse/fine overset grid systems were per. This request was made to give participants the opportunity to
derived from the medium grids by decreasing/increasing each grid complete computations and correct errors discovered in their data
dimension by a factor of 1.5 in each zone. In this manner, the size entries. Regrettably, five participants withdrew their data submittals
of the grids in the coarse–medium–fine sequence grows by a factor from consideration or did not resubmit final data sets and so their
of about 3.375. This parametric variation of grid size conforms to results are not represented in this paper.
the DPW-II specifications. Table 6 provides the grid sizes. Tables 7 and 8 summarize the 29 data sets that were resubmitted
for analysis. Table 7 lists the organization submitting the data set,
Test Case Description the code used, and which cases were included in the set. The list
Participants were asked to submit data for two required test cases, is not in any particular order. Both complete and partially complete
case 1 and case 2, and were encouraged to complete two optional data sets are included. Additionally, a solution index i is assigned
test cases, case 3 and case 4 (WB and WBNP, respectively). Be- to each data set, so that the information in Table 7 can be cross
cause each test case involved computations for the WB and WBNP referenced with the information in Table 8.
configurations and either multiple grids or multiple test conditions, Table 7 Participants and submitted data
completion of all four test cases required a significant computational
effort. i Organization Code Cases
The required cases are described next. Case 1 is a fixed-C L 1 Airbus Industries ELSA 1, 2, 3
single-point grid refinement study (six solutions in total), where 2 CFX, Ansys CFX-5 1, 2
M∞ = 0.75 and C L = 0.500 ± 0.001. The boundary layer is fully 3 Boeing Commercial CFL3D 1, 2
turbulent. Coarse, medium, and fine grids are used. 4 Boeing Commercial CFL3D 1, 2, 3, 4
Case 2 is a drag polar study (14 solutions in total), where 5 Boeing Commercial TLNS3D 2
M∞ = 0.75 and α = −3, −2, −1.5, −1, 0, 1, and 1.5 deg. There is 6 Boeing Phantom Works OVERFLOW 1, 2, 3
boundary-layer transition at the lower surface quarter chord and up- 7 Cessna NSU3D 1, 2,
8 Cobalt Solutions COBALT 1, 2
per surface at 5% chord at root, 15% chord at kink, 15% chord at
9 DLR TAU 1, 2, 3, 4
η = 0.844, 5% chord at tip, or 10% chord if trip location cannot be 10 ONERA TAU 2
varied, or the boundary layer is fully turbulent if the trip location 11 Swedish Defence Agency EDGE 1, 2
cannot be specified. A medium grid from case 1 or the participants’ 12 Metacomp Technologies CFD++ 1, 2
own best practices grid is used. 13 Korean Advanced Institute KFLOW3D 1, 2, 4
A description of the optional cases is given next. of Science and Technology
Case 3 is a comparison of tripped and fully turbulent solutions 14 Kawasaki Heavy Industries UG3 1, 2, 3
(four solutions in total), where M∞ = 0.75 and C L = 0.500 ± 0.001. 15 National Aeronautical Laboratory UPACS 1, 2, 3
There is a fully turbulent boundary layer and the boundary-layer 16 NASA Langley Research Center USM3Dns 1, 2
17 NASA Langley Research Center FUN3D 1, 2, 3
tripped transition location as per case 2. The medium grid from
18 NASA Langley Research Center CFL3D 1, 2
case 1 or the participants’ own best practices grid is used. 19 NASA Langley Research Center CFL3D 1, 2, 3
Case 4 is a study of constant-C L drag rise (16 solutions in to- 20 NASA Langley Research Center CFL3D 1, 2
tal), where M∞ = 0.50, 0.60, 0.70, 0.72, 0.74, 0.75, 0.76, and 0.77 21 NASA Langley Research Center CFL3D 1, 2, 3
and C L = 0.500 ± 0.001. There is boundary-layer transition as per 22 NASA Langley Research Center OVERFLOW 1, 2, 3
case 2. The medium grid from case 1 or the participants’ own best 23 NASA Langley Research Center OVERFLOW 1
practices grid is used. 24 National Institute of Aerospace (NIA) NSU3D 1, 2, 3, 4
Participants were asked to perform case 1 to examine the effect of 25 National Aerospace Laboratory ENSOLV 1, 2, 3
grid refinement on absolute and delta drag predictions. In the interest 26 QinetiQ SAUNA 1, 2
27 NIA NSU3D 2
of improving the statistical basis of the data, it was requested that all
28 Science Applications International FEFLO 1, 2
calculations for case 1 be made using fully turbulent boundary-layer Corporation
specifications. Additionally, it was requested that the grids supplied 29 Fluent, Inc. FLUENT 6.1 1, 2
by the DPW-II organizing committee be used if possible.
1170 LAFLIN ET AL.

Table 8 Computational details of submitted data

Medium grid N × 10−6


Turbulence Boundary Average − σ <
i Key Grid modela Layer NWB NWBNP N < Average − σ N < Average + σ N > Average + σ
1 A S Wilcox Tr 5.83 8.43 X
2 B U SST Tr 5.83 8.43 X
3 C S SA FT 3.88 5.82 X
4 D S SST Tr 3.88 5.82 X
5 E S SST FT 3.88 5.82 X
6 F O SA Tr 6.86 10.78 X
7 G U SA FT 2.58 3.98 X
8 H U SST FT 4.88 6.04 X
9 I U SAE Tr 5.84 8.20 X
10 J U SAE FT 5.84 8.20 X
11 K U k–ω FT 3.16 4.79 X
12 L U k–ε FT 7.38 8.51 X X
13 M S k–ω FT 3.66 5.75 X
14 N U SA Tr 4.07 6.67 X
15 O S SA Tr 7.94 12.56 X
16 P U SA FT 3.90 5.91 X
17 Q U SA FT 3.01 4.75 X
18 R S EASM–ko Tr 5.72 8.29 X
19 S S SA Tr 5.72 8.29 X
20 T S SST Tr 5.72 8.29 X
21 U O SA Tr 6.59 10.33 X
22 V O SA Tr 6.59 10.33 X
23 W O SST FT 6.59 10.33 X
24 X U SA FT 3.01 4.75 X
25 Y S Wilcox Tr 5.50 8.30 X
26 Z S k–g Tr 1.63 2.86 X
27 a U SA Tr 3.01 4.75 X
28 b U SA FT 1.78 2.42 X
29 c S k–ε FT 5.72 8.29 X
a
Includes variant formulations.24

Table 9 WB and WBNP medium grid sizes Table 10 WB and WBNP coarse and fine grid sizes

Statistics NWB NWBNP Coarse Fine


Minimum 1,632,000 2,419,388 Statistics NWB NWBNP NWB NWBNP
Maximum 7,942,944 12,562,736
Average 4,827,148 7,162,311 Minimum 943,232 1,570,342 2,390,716 3,682,535
σ 1,690,857 2,468,449 Maximum 5,527,488 6,219,855 23,150,448 35,867,680
Average 2,359,351 3,701,039 10,295,438 14,130,000
σ 1,194,051 1,526,436 5,660,303 8,815,029

Participants were asked to submit computed C P distributions


along defined wing sections for specific conditions. These C P data Table 11 Number of data set submittals by case
sets can be found on the DPW-II website but are not presented in this
summary paper. Additionally, some participants included residual, Case Submittals
force, and moment convergence information, which are not pre- 1 26
sented here but can be found in the DPW-II website posted data 2 28
sets.24 3 12
Table 8 provides computational details about each data set in- 4 4
cluding the grid type, turbulence model used, boundary-layer tran-
sition specification, and sizes of both the WB and WBNP con-
figuration medium density grids. The turbulence models included pleteness, the coarse and fine grid size statistical values are given in
Wilcox, k–ω shear stress transport (SST), Spalart–Allmaras (SA), Table 10.
Spalart-Allmaras-Edwards (SAE), k–ω, k–ε, Explicit-Algebraic- Table 11 gives a breakdown of the total submittals for each case.
Stress-Model (EASM)-ko, and k–g. The boundary layer had tripped For the 26 case 1 contributions, the multiblocked structured and
transition (Tr) or was fully turbulent (FT). Table 8 also assigns a key unstructured grids are equally represented by 11 submittals, respec-
to each data set, which is used to identify the data presented in the tively. Only two different participants used the overset grids for a
summary shown in the next section. The boundary-layer transition total of four case 1 data sets.
specifications in Table 8 apply to cases 2 and 4 only because par- A general breakdown of the turbulence models used for the 26
ticipants were asked to run FT solutions for case 1 and both Tr and case 1 submittals is given in Table 12. The reader is referred to the
FT solutions for case 3. Despite an attempt to minimize grid size data sets and workshop presentations24 for a more detailed expla-
variability by the DPW-II organizing committee through the con- nation of the turbulence models.
struction of standard grids, a significant variation in grid size exists. Participants were asked to provide C L , C D , and C M predictions
Variations in the number of grid nodes for the WB and WBNP for each calculation. Additionally, participants were asked to de-
medium grids are given in Table 9. compose predicted drag coefficients into pressure drag and viscous
To compare case 2 drag polar results against grid size, the data sets drag components. These results are presented in the next section.
were grouped into three different categories based on the size of the
medium density grid. (See the last three columns of Table 8.) The Results and Discussion
colors red, blue, and green are associated to the three categories, Before DPW-II CFD results are presented, a brief discussion
respectively, and they are used as shown subsequently. For com- concerning the appropriateness of comparing CFD predictions to
LAFLIN ET AL. 1171

Table 12 Case 1: breakdown by turbulence modela

Times used; Model


12 SA, SAE
5 SST
2 Wilcox
1 EASM–ko
3 k–ω
2 k–ε
1 k–g
a
Includes variant formulations.24
wind-tunnel measurements is given. Additionally, experimentally
visualized separation regions on the two DLR-F6 configurations
are discussed. These two considerations are offered before the re-
sults presentation so that the reader may keep these issues in mind
when viewing the results.
The discussion of the appropriateness of comparing CFD predic-
tions to wind-tunnel measurements echos the discussion presented a) Upper planview, wing-root and trailing-edge flow separation
in the DPW-I summary paper.8 The CFD solutions are for free-
air models. Even though the tunnel data are corrected to a free-air
condition, the correction process introduces some error. To match
wind-tunnel data accurately, the CFD computations should include
the mounting hardware and tunnel walls (perhaps porous or slot-
ted) and the tunnel data should not include some of the corrections
normally applied, for example, blockage. However, a comparison
of this type is not usually done in practice and was not done in this
study. Therefore, it is not surprising or, it could be argued, even of
concern if CFD solutions do not precisely match wind-tunnel re-
sults. The greater concern in the present study is the scatter of the
CFD predictions.
In regards to the scatter of the CFD data, note that significant
regions of boundary-layer separation were noted on both the WB
and WBNP configurations using oil-flow visualization techniques
in the wind tunnel. Significant regions of boundary-layer separa-
tion exist on the wing upper surface near the trailing edge (Fig. 4a).
These separation regions are present on both configurations. Ad-
ditional pockets of separation near the pylon exist on the lower b) Lower planview wing lower-surface flow separation near pylon
surface of the wing of the WBNP configuration (Fig. 4b). Although Fig. 4 DLR-F6 wind-tunnel model with oil-flow patterns, M∞ = 0.75
Reynolds-averaged Navier–Stokes codes can predict such regions and CL = 0.5.
of boundary-layer separation, the size, shape, and temporal behavior
of the separation is greatly effected by local grid density/quality, nu-
merical dissipation, and the turbulence model employed. Because of
the variety of numerical schemes, grids, and turbulence models used
by the DPW-II participants, notable variations in the predicted sep-
aration regions, and the corresponding effect on force and moment
calculations, should be expected.
Figure 5 shows CFD predicted skin-friction coefficient surface
contours and streamlines in the area of the trailing-edge WB junc-
ture separation bubble, which is visible in the oil-flow visualizations
(Fig. 4a). To assess quantitatively variations in this bubble between
the various CFD solutions, participants were asked to provide geo-
metric coordinates of specific identifiable features of the separation
of the fine grid results (Fig. 5). Figure 6 shows predicted fuselage
station (FS) and butt line (BL) coordinates of the separation “eye”
feature. Other feature coordinates are contained within the original
data sets.24
The predicted chordwise position, FSeyew , of the wing eye sepa-
ration feature varies by about 10% of the wing mean aerodynamic
chord. The spanwise location, BLeyew , variation is roughly 5% of
the wing half-span length. Clearly, differences exist in the CFD-
predicted separation regions, which correspondingly result in differ- Fig. 5 Navier–Stokes code predicted DLR-F6 wing–body separation
ences in the CFD force and moment predictions. Note the preceding bubble, definition of measurement points.
two issues when examining the data. The former issue addresses dif-
ferences between CFD and experiment. The latter issue addresses Participants were asked to obtain FT solutions with C L values that
scatter among the CFD solutions. fell within the target range of C L = 0.500 ± .001. This tight range
was requested to obtain good comparisons of C D , which varies with
Case 1 C L . As can be seen in Figs. 7a and 7b, some submitted solutions
Case 1 (required) specified a single fixed-C L and Mach number did not meet the specified C L constraint. The remaining presenta-
condition on each of a series of coarse, medium, and fine density tion and discussion of the case 1 data does not include data sets
grids for both the WB and WBNP DLR-F6 configurations. The in which the C L value fell outside of the specified target range.
wind-tunnel experimental results for case 1 are given in Table 13. Statistical quantities for the WB coarse, medium, and fine grid
1172 LAFLIN ET AL.

Table 13 Case 1, interpolated experimental values Table 14 Case 1: WB coarse grid results

Parameters WBa WBNPa Parameters Average Minimum Maximum  Expmt


α, deg 0.52 1.00 α, deg 0.131 −0.151 0.500 −0.389
CL 0.500 0.500 CL 0.5001 0.4990 0.5010 0.0001
CD 0.0295 0.0338 CD 0.02985 0.02777 0.03258 0.00035
CM −0.1211 −0.1199 C D pressure 0.01662 0.01500 0.01892 ——
C D viscous 0.01302 0.00961 0.01400 ——
a
Interpolated from wind-tunnel data.25 CM −0.1379 −0.1633 −0.1195 −0.01680

Table 15 Case 1: WB medium grid results

Parameters Average Minimum Maximum  Expmt


α, deg 0.161 −0.059 0.500 −0.359
CL 0.5000 0.4993 0.5010 0.0
CD 0.02902 0.02799 0.03038 −0.00048
C D pressure 0.01609 0.01500 0.02001 ——
C D viscous 0.01293 0.00970 0.01416 ——
CM −0.1382 −0.1686 −0.1192 −0.0171

Table 16 Case 1: WB fine grid results

Parameters Average Minimum Maximum  Expmt


α, deg 0.1529 −0.128 0.540 −0.367
CL 0.4999 0.4990 0.5010 −0.0001
CD 0.02883 0.02768 0.03034 −0.00067
C D pressure 0.01590 0.01440 0.01913 ——
C D viscous 0.01294 0.00976 0.01434 ——
CM −0.1397 −0.1601 −0.1201 −0.0186

Table 17 Case 1: WBNP course grid results

Parameters Average Minimum Maximum  Expmt


α, deg 0.648 0.462 0.850 −0.352
CL 0.5001 0.4990 0.5010 0.0001
Fig. 6 Wing separation eye results for case 1: M∞ = 0.75 and CL = 0.5. CD 0.03497 0.03320 0.03792 0.00117
C D pressure 0.01962 0.01690 0.02192 ——
C D viscous 0.01570 0.01404 0.01661 ——
CM −0.1361 −0.1581 −0.1180 −0.0162

Table 18 Case 1: WBNP medium grid results

Parameters Average Minimum Maximum  Expmt


α, deg 0.716 0.466 1.225 −0.284
CL 0.5000 0.4990 0.5008 0
CD 0.03395 0.03235 0.03616 0.00015
C D pressure 0.01870 0.01708 0.02237 ——
C D viscous 0.01528 0.01077 0.01679 ——
a) WB CM −0.1339 −0.1567 −0.1125 −0.0140

Table 19 Case 1: WBNP fine grid results

Parameters Average Minimum Maximum  Expmt


α, deg 0.721 0.381 1.015 −0.279
CL 0.4999 0.4990 0.5010 −0.0001
CD 0.03387 0.03207 0.03631 0.00007
C D pressure 0.01852 0.01703 0.02118 ——
C D viscous 0.01528 0.01088 0.01703 ——
CM −0.1345 −0.1568 −0.1120 −0.0146
b) WBNP
Fig. 7 Target CL results for case 1: M∞ = 0.75. WB CFD fine grid solutions underpredict C D by 6.7 drag counts.
For the WBNP configuration, the difference in C D decreases with
CFD solutions are listed in Tables 14, 15, and 16, respectively. In- increased grid density. The average fine grid WBNP solutions over-
cluded in Tables 14–16 are the differences between the average CFD predict C D by only 0.7 drag counts. For both configurations and all
quantities and the experimental results; these terms are denoted by grids, C M is predicted as too nose down.
 Expmt. Similarly, WBNP quantities are listed in Tables 17–19. The DPW-II organizing committee originally envisioned the grid-
The CFD results generally underpredict α by about a one-third of refinement study as a grid-convergence study, in which force and
a degree for both configurations and all grids. A slight improvement moment data could be extrapolated to values that would be obtained
in the α predictions is seen with increased grid density. An increase in the limit of infinite grid density using Richardson extrapolation
in grid density has the effect, on average, of lowering the predicted (see Ref. 40). For this procedure to be valid, the series of computed
C D values. The difference between the average WB CFD and ex- solutions on the various grids must show asymptotic convergence.
periment C D increases with increased grid density. On average, the Unfortunately, the grids used were not dense enough to obtain this
LAFLIN ET AL. 1173

Table 20 Case 1: configuration ∆CD results of 9, 6, and 8 drag counts for the coarse, medium, and fine grid
solutions, respectively. This result indicates that, even with the ex-
Grid Average Minimum Maximum  Expmt
istence of significant regions of separation, configuration delta drag
Coarse 0.00520 0.00410 0.00602 0.00090 predictions are largely insensitive to grid density, unlike absolute
Medium 0.00458 0.00342 0.00633 0.00063 drag predictions.
Fine 0.00498 0.00318 0.00600 0.00078
Case 2
Case 2 (required) is representative of a typical constant Mach
number sweep, that is, drag polar, performed in wind-tunnel testing.
Summary results for case 2 include drag coefficient, lift coefficient,
and pitching moment coefficient predictions. Also, drag predictions
are compared against grid size, grid type, turbulence model, and
boundary-layer transition specification to examine how each affect
drag prediction. The reader is reminded that Tables 10 and 11 should
be used as legends for the case 2 plotted results. In the summary
plots of case 2, (see Figs. 12–20) the drag values are presented in
terms of the idealized profile drag coefficient41
a) WB C D P = C D − C L2 /(πAR)

The idealized profile drag coefficient is used instead of the total


drag coefficient C D because it generally results in a more compact
presentation of the data, allowing expanded scales. This, in turn,
allows differences between data sets to be more discernable.
Composite lift curves are shown in Figs. 10a and 10b for the WB
and WBNP configurations, respectively. The CFD results consis-
tently overpredict lift at a given α. This behavior was also true of
the DPW-I (Ref. 8) results. C Lα values were obtained by a linear
fit of the data using values of α ranging from −3 to 1.5 deg. Both
the WB and WBNP predicted lift curve slopes are less than the
b) WBNP experimental values (Table 21).
Fig. 8 Drag results for case 1: M∞ = 0.75 and CL = 0.5.
Table 21 Case 2: drag polar CLα results

Configuration Average Minimum Maximum  Expmt


WB 0.1175 0.1075 0.1240 −0.0012
WBNP 0.1111 0.0966 0.1228 −0.0057

Fig. 9 Delta (WBNP − WB) drag results for case 1: M∞ = 0.75 and
CL = 0.5.

necessary behavior. Additionally, some of the supplied grids did not


maintain similarity between the coarse, medium, and fine density
grids.
Although a true grid-convergence study could not be performed,
it is interesting to examine the behavior of predicted C D values with
reported grid size, shown in Figs. 8a and 8b for the WB and WBNP a) WB
configurations, respectively. Coarse, medium, and fine grid results
are displayed together. The range in WB configuration predicted C D
values (Fig. 8a) narrows from about 70 drag counts to approximately
12.5 drag counts as the number of grid nodes is increased. Similarly,
the range of WBNP C D values (Fig. 8b) vary from approximately
50 drag counts to 7.5 drag counts as N is increased. Note that the
observation that the drag range decreases with increased grid density
cannot be substantiated statistically due to poor sampling at the
higher grid densities.
Configuration delta drag coefficients C D of all submitted data
sets are shown in Fig. 9. C D is defined as the DLR-F6 WBNP con-
figuration C D value minus the WB configuration C D value. Most of
the C D values are within a range of 20 drag counts, regardless of
grid size. Coarse, medium, and fine grid C D statistics are given in
Table 20. The values of Table 20 do not include data sets in which
the C L value fell outside of the requested target range for either
the WB or WBNP configurations. The Navier–Stokes codes consis- b) WBNP
tently overpredict the delta drag configurations by average values Fig. 10 Composite lift curve results for case 2: M∞ = 0.75.
1174 LAFLIN ET AL.

a) WB b) WBNP
Fig. 11 Composite pitching moment results for case 2: M∞ = 0.75.

a) WB b) WBNP
Fig. 12 Composite drag polar results for case 2: M∞ = 0.75.

Composite pitching moment coefficient curves are shown in


Figs. 11a and 11b. The majority of computed data sets overpre-
dict the nose-down pitching moment (more negative), as compared
to the experiment. DPW-I results also displayed a general overpre-
diction of nose-down pitching moment. The variation in the pitching
data is large, but appears to reduce with increased grid density.
Figures 12a and 12b show the composite idealized drag polars for
the WB and WBNP configurations, respectively. The CFD predic-
tions generally underpredict drag for the WB configuration. WBNP
drag is overpredicted for lower C L values and slightly underpre-
dicted at higher values. The crossover point is near the DLR-F6 de-
sign cruise condition, C L des = 0.5. The range in C D is about 35 drag
counts for the WB configuration. This range is fairly consistent
over the C L values examined. The WBNP configuration C D range
is about 55 drag counts for C L values lower than 0.4. Above this
point, the range becomes larger, presumably due to larger variations
in predicted separation regions as wing loading is increased. The
breakdown of the drag predictions into components of pressure drag
and viscous drag is not presented here but may be found in Ref. 9.
Figures 13–20 are the idealized profile polars compared against
grid size, grid type, turbulence model, and boundary-layer transition
specification. Figures 13 and 14 show drag polar trends with grid
size for the WB and WBNP configurations, respectively. Figures 13
and 14 indicate an apparent advantage of increased grid density.
The coarser grids have the largest range, 40 and 65 drag counts. For
the medium grids, this range is reduced to 25 and 35 drag counts,
and the finest grids exhibit the smallest range with 20 and 10 drag
counts for the WB and WBNP configurations, respectively. The
small range seen in the fine grid solutions may be a result of an
Fig. 13 WB, drag polar trends with grid size for case 2: M∞ = 0.75.
insufficient number of samples. Note that a statistical analysis of
the case 1 data10 did not show an improvement in drag scatter with grid solutions use refined (green) grids, and so an advantage cannot
grid refinement, suggesting that the case 2 results be examined in be attributed to the overset grid type.
further detail to assess properly the significance of grid density. No advantage can be attributed to the turbulence type (Figs. 17
In Figs. 15 and 16, drag polar trends for the different grid types and 18). Likely, an advantage in turbulence model or grid type, if
are shown (coarse grid, red; medium grid, blue; and fine grid, green). one exists, will not be differentiable unless true grid convergence is
No obvious advantage of grid type can be discerned from Figs. 15 achieved.
and 16. The range of the overset grid drag is small compared to the Finally, drag polar trends with boundary-layer transition spec-
multiblocked structured and unstructured grids, but all of the overset ification are shown in Figs. 19 and 20, for the WB and WBNP
LAFLIN ET AL. 1175

Fig. 14 WBNP, drag polar trends with grid size for case 2: M∞ = 0.75. Fig. 16 WBNP, drag polar trends with grid type for case 2: M∞ = 0.75.

Fig. 17 WB, drag polar trends with turbulence model for case 2:
Fig. 15 WB, drag polar trends with grid type for case 2: M∞ = 0.75. M∞ = 0.75.
configurations, respectively. As expected, specifying a boundary- Case 3
layer trip position, allowing a run of laminar flow, reduced the Case 3 (optional) is a single fixed-C L and Mach number con-
drag. Additionally, tripping the boundary layer generally reduced the dition using FT boundary-layer specification and fixed position
range in the drag data compared to the FT solutions. It is curious that, transition, that is, tripped boundary layer, specification on both
although the experimental wind-tunnel model used boundary-layer DLR-F6 configurations. This test case was chosen to determine
trips, the FT CFD solutions better predict the experimental results for two different delta drag values. The first delta drag of interest is the
the WB configuration, Fig. 19. However, the tripped CFD solutions difference between FT and tripped boundary-layer solutions. This
better match the WBNP configuration experimental data, Fig. 20. delta drag comparison is summarized in Table 22. Here, the tripped
1176 LAFLIN ET AL.

Table 22 Case 3: BL transition ∆CD results

FT-tr Average Minimum Maximum


WB 0.00092 0.00041 0.00119
WBNP 0.00071 −0.00045 0.00111

Table 23 Case 3: configuration ∆CD results

WBNP–WB Average Minimum Maximum  Expmt


Tr 0.00529 0.00464 0.00591 0.00099
FT 0.00507 0.00432 0.00553 0.00077

Fig. 18 WBNP, drag polar trends with turbulence model for case 2:
M∞ = 0.75.

Fig. 20 WBNP, drag polar trends with boundary-layer treatment for


case 2: M∞ = 0.75.

Fig. 21 Delta (FT-Tr) drag results for case 3: M∞ = 0.75 and CL = 0.5.

Fig. 19 WB, drag polar trends with boundary-layer treatment for


case 2: M∞ = 0.75.

C D values are subtracted from the FT C D values. The delta drag


due to boundary-layer transition specification for both the WB and
WBNP configurations shows a difference of only two drag counts Fig. 22 Delta (WBNP − WB) drag results for case 3: M∞ = 0.75 and
between the WB and the WBNP configurations. An average drag CL = 0.5.
increase of less than 10 drag counts is attributed to specifying an
FT boundary layer as opposed to specifying boundary-layer transi- computations are performed, the delta drag value is nearly the same,
tion trip positions. Individual delta drag comparisons are shown in varying, on average, by two drag counts. This result suggests that
Fig. 21. CFD configuration delta drag predictions can be made using the
The second delta drag of interest is the difference between WBNP less physically correct but simpler FT boundary-layer specification
and WB configurations. Predicted configuration delta drags are method when the configurations being considered are similar and
shown in Fig. 22 and summarized in Table 23. Here, WB C D val- the grids have similar density distributions. The CFD delta drag pre-
ues are subtracted from the WBNP C D values. Whether a tripped dictions over-predict the experimental delta drag values by less than
boundary-layer transition position is specified or FT boundary-layer 10 drag counts.
LAFLIN ET AL. 1177

documented in a companion paper,9 but key results will be stated.


Compared to DPW-I, DPW-II results show a 3:1 (or better) reduction
in the spread of the results, but an improvement in dispersion of
“core” solutions is more modest. The DLR-F6 WB medium density
grid results, which compare best to the DLR-F4 DPW-I results,
are considerably better than DPW-I results; however, the cause is
not known. Figures 24a and 24b show examples of the statistical
results for C D for the WB configurations DLR-F4 and DLR-F6
from DPW-I and DPW-II, respectively. The reduction in the spread
is obvious. With regard to the grid convergence for the collective,
a) WB there is no reduction in spread and there is no reduction in core
scatter. Incremental values, for example, configuration C D , tend
to be considerably better in both scatter and median.

Conclusions
A workshop was held to assess the state-of-the-art of CFD
Reynolds-averaged Navier–Stokes solvers to predict lift, drag, and
pitching moment of industry relevant aircraft configurations, fo-
cusing principally on drag calculations. Participants were asked to
compute a prescribed set of test cases on the DLR-F6 WB and
WBNP configurations. The test cases consisted of fixed-C L single-
b) WBNP point solutions, drag polars, and constant-C L drag rise curves. Stan-
dardized multiblock structured, unstructured, and overset structured
Fig. 23 Drag rise results for case 4.
grids were made available to workshop participants to encourage
participation and reduce grid variability. Grid refinement, configu-
ration delta drags, and the effect of boundary-layer transition speci-
fication were all examined. A large body of data was gathered from
20 international participants and presented in an objective manner.
In general, the CFD lift levels are higher than the wind-tunnel
results, drag is lower, and pitching moment is more negative (more
nose down) at a given angle of attack for the DLR-F6 WB config-
uration. For the DLR-F6 WBNP configuration, the CFD lift levels
are generally higher, and pitching moment is more negative (more
nose down) than the wind-tunnel results at a given angle of attack.
a) The WBNP configuration CFD-predicted drag is higher then experi-
mentally measured drag for C L values less than about 0.5; otherwise
it is lower.
Although the DPW-II organizing committee envisioned a grid-
convergence study as part of the workshop, the various series of
coarse, medium, and fine grids used by participants were of insuffi-
cient density to obtain asymptotic solution convergence. However,
it was noted that the range in the CFD predictions decreased with
increased grid density, indicating that proper grid resolution may
be the single most important factor in achieving quality CFD drag
predictions, compared to grid type or turbulence model. However,
this conclusion was not substantiated by statistical analysis.
b) In addition to grid size, the effects of boundary-layer transition
Fig. 24 Example of statistical results for CD (WB configuration) from specification, that is, Tr vs FT, turbulence model, and grid type on
Ref. 10. the range in the drag data were examined. None of these appear to
have as significant an effect as grid size. Specifying the boundary-
Case 4 layer trip location appeared to reduce the range in drag as compared
Case 4 (optional) is computed drag rise curves for both DLR-F6 to specifying FT boundary-layer computations. Tripped solutions
configurations. This test case required a significant computational were, on average, about 10 drag counts lower than FT solutions. No
effort because the solution must be converged to a specified C L value conclusive advantage as to grid type or turbulence model could be
for a number of different Mach number conditions. As such, only discerned from the data.
four participants submitted case 4 data; these are shown in Figs. 23a Delta drag comparisons between the DLR-F6 WB and WBNP
and 23b for the WB and WBNPs configurations, respectively. Three configurations were examined. The variation in these two configu-
of the four submitted data sets consistently underpredict drag for rations is more extreme than usually considered in CFD delta drag
both configurations over the range of Mach numbers for which ex- studies. Additionally, CFD drag comparisons that are conducted
perimental data is known. The general shape of the drag rise is for the purpose of making design decisions focus on configuration
captured in the WBNP computations but is generally too flat for the variants absent of large pockets of separated flow, unlike the config-
WB configuration. The tripped solutions show less variance to each urations considered for the workshop. Nevertheless, the workshop
other than do the two FT solutions. The increased variation for the results indicate that Tr boundary-layer treatment is not critical for
FT cases may be due to the unique numerical transition behavior delta drag comparisons. Additionally, CFD delta drag predictions
inherent in all turbulence models, that is, each turbulence model were less dependent on grid density than CFD absolute drag predic-
transitions at a different location even when allowed to run in the tions. These results tend to support the current industry practice of
FT mode. using FT boundary-layer calculations on coarse density grids to ob-
tain configuration delta drag values quickly. This procedure assumes
Statistical Analysis similar errors exist in both CFD solutions and that these errors can-
A rigorous statistical analysis of the DPW-II CFD results was cel each other, resulting in an adequate estimate of delta drag. Ide-
performed and presented during the workshop.24 These results are ally, this procedure should only be performed when configuration
1178 LAFLIN ET AL.

changes are small, flow separation is minimal, and all grids have 15 Wutzler, K., “Aircraft Drag Prediction Using Cobalt,” AIAA Paper

similar grid density distributions. 2004-0395, Jan. 2004.


16 May, G., van der Weide, E., Jameson, A., and Shankaran, S.,
It is recommended that future workshops consider including one
test case involving a grid-converged solution. The geometry for “Drag Prediction of the DLR-F6 Configuration,” AIAA Paper 2004-0396,
such a test case would need to be simple enough that the size of Jan. 2004.
17 Kim, Y., Park, S., and Kwon, J., “Drag Prediction of DLR-F6 Using the
the grid needed to obtain grid-converged solutions does not restrict Turbulent Navier–Stokes Calculations with Multigrid,” AIAA Paper 2004-
participation due to computational resource or time constraints. A 0397, Jan. 2004.
comparison of grid-converged solutions would eliminate the solu- 18 Yamamoto, K., “CFD Sensitivity to Drag Prediction on DLR-F6 Con-
tions dependency on grid size and allow true effects of grid type, figuration by Structured Method and Unstructured Method,” AIAA Paper
turbulence models, etc., to be assessed. Additionally, it would be 2004-0398, Jan. 2004.
19 Tinoco, E., and Su, T., “Drag Prediction with the Zeus/CFL3D System,”
very interesting to see how much the spread in the CFD predictions
can be reduced when grid density effects are eliminated. AIAA Paper 2004-0552, Jan. 2004.
20 Klausmeyer, S., “Drag, Lift, and Moment Estimates for Transonic
It is also recommended that future workshop test cases consider
geometries for which configuration flight data are available, in ad- Aircraft Using the Navier–Stokes Equations,” AIAA Paper 2004-0553,
Jan. 2004.
dition to wind-tunnel test data. A major advantage of CFD is that 21 Lee-Rausch, E., Frink, N., Milholen, W., and Mavriplis, D., “Transonic
it can model aircraft geometry at flight conditions, including flight Drag Prediction Using Unstructured Grid Solvers,” AIAA Paper 2004-0554,
Reynolds number, without the flow obstructions introduced dur- Jan. 2004.
ing wind-tunnel testing. A comparison between wind-tunnel data, 22 Pfeiffer, N., “Reflections on the Second Drag Prediction Workshop,”
corrected for flight Reynolds number, flight test data, and a large AIAA Paper 2004-0557, Jan. 2004.
23 Hemsch, M., “Statistical Analysis of CFD Solutions form the Drag Pre-
database of CFD predictions would be of extreme interest to indus-
try and helpful in further establishing the credibility of CFD. diction Workshops,” CFD-based Aircraft Drag Prediction and Reduction,
von Kármán Inst. Lecture Series, Brussels, 2003.
24 “2nd AIAA CFD Drag Prediction Workshop,” URL: http://ad-www.
Acknowledgments
larc.nasa.gov/tsab/cfdlarc/aiaa-dpw/index.html [cited 20 Sept. 2004].
The authors wish to thank the AIAA Applied Aerodynamics 25 Redeker, G., Schmidt N., and Müller R., “Design and Experimental
Technical Committee for sponsoring the Drag Prediction Workshop Verification of a Transonic Wing for Transport Aircraft,” Proceedings of
(DPW). We also wish to thank our respective companies and re- the FDP Symposium on Subsonic/Transonic Configuration Aerodynamics,
search organizations for their support and acknowledge Elizabeth CP-285, AGARD, 1980, pp. 13.1–13.7.
Lee-Rausch of NASA Langley Research Center for providing con- 26 Gerhold, T., Friedrich, O., Evans, J., and Galle, M., “Calculations of

tent on the VGRIDns unstructured grids for this paper. Finally, Complex Three-Dimensional Configurations Employing the DLR TAU-
a special thanks goes to all of the DPW-II participants because Code,” AIAA Paper 97-0167, Jan. 1997.
27 Kroll, N., Rossow, C.-C., Becker, K., and Thiele, F., “The MEGAFLOW
without their contributions this workshop would not have been
possible. Project,” Aerospace Science Technology, Vol. 4, No. 4, 2000, pp. 223–237.
28 Rossow, C.-C., Godard, J.-L., Hoheisel, H., and Schmitt, V., “Investiga-
tions of Propulsion Integration Interference Effects on a Transport Aircraft
References Configuration,” Journal of Aircraft, Vol. 31, No. 5, 1994, pp. 1022–1030.
1 Proceedings of the First AIAA CFD Drag Prediction Work- 29 Pierre, M., and Fasso, G., “The Aerodynamic Test Center of Modane-
shop, URL: http://aaac.larc.nasa.gov/tsab/cfdlarc/aiaa-dpw/Workshop1/ Avrieux,” ONERA TN 166E, March 1972.
workshop1.html, email: dpw@cessna.textron.com [cited 20 Sept. 2004]. 30 Tinoco, E. N., and Bussoletti, J. E., “Minimizing CFD Uncertainty
2 Redeker, G., “DLR-F4 Wing-Body Configuration,” A Selection of Exper- for Commercial Airplane Applications,” AIAA Paper 2003-0407, Jan.
imental Test Cases for the Validation of CFD Codes, AGARD Rept. AR-303, 2003.
Vol. 1, Aug. 1994, pp. 88–89. 31 Capron, W. K., and Smit, K. L., “Advanced Aerodynamic Applications
3 Hemsch, M., “Statistical Analysis of CFD Solutions from the Drag Pre- of an Interactive Geometry and Visualization System,” AIAA Paper 91-0800,
diction Workshop,” AIAA Paper 2002-0842, Jan. 2002. June 1991.
4 Rakowitz, M., Eisfeld, B., Schwamborn, D., and Sutcliffe, M., “Struc- 32 Su, T. Y., Wilkinson, W. M., and Yu, N. J., “Structured Navier–Stokes
tured and Unstructured Computations on the DLR-F4 Wing-Body Configu- Grid Generations for Propulsion-Integrated Airplane Configurations,” 5th
ration,” Journal of Aircraft, Vol. 40, No. 2, 2003, pp. 256–264. International Conference on Numerical Grid Generation in Computational
5 Mavriplis, D. J., and Levy, D. W., “Transonic Drag Prediction Using an Field Simulations, edited by B. K. Soni, J. F. Thompson, J. Häuser, and P.
Unstructured Multigrid Solver,” AIAA Paper 2002-0838, Jan. 2002. Eiseman, NSF Engineering Research Center, Mississippi State, Mississippi,
6 Pirzadeh, S. Z., “Assessment of the Unstructured Grid Software TetrUSS 1996, pp. 1193–1203.
for Drag Prediction of the DLR-F4 Configuration,” AIAA Paper 2002-0839, 33 Lohner, R., and Parikh, P., “Generation of Three Dimensional Unstruc-
Jan. 2002. tured Grids by the Advancing Front Method,” International Journal of Nu-
7 Vassberg, J. C., Buning, P. G., and Rumsey, C. L., “Drag Prediction for merical Methods in Fluids, Vol. 8, No. 10, 1988, pp. 1135–1149.
the DLR-F4 Wing/Body Using OVERFLOW and CFL3D on an Overset 34 Pirzadeh, S., “Three-Dimensional Unstructured Viscous Grids by
Mesh,” AIAA Paper 2002-0840, Jan. 2002. the Advancing Front Method,” AIAA Journal, Vol. 34, No. 1, 1996,
8 Levy, D. W., Zickuhr, T., Vassberg, J., Agrawal, S., Wahls, R., Pirzadeh, pp. 43–49.
S., and Hemsch, M., “Summary of Data from the First AIAA CFD Drag Pre- 35 CentaurSoft, http://www.centaursoft.com [cited 20 Sept. 2004].
diction Workshop,” Journal of Aircraft, Vol. 40, No. 5, 2003, pp. 875–882. 36 Gerhold, T., and Evans, J., “Efficient Computation of 3D-Flows for
9 Laflin, K., Klausmeyer, S., Zickuhr, T., Vassberg, J., Wahls, R., Morrison, Complex Configurations with the DLR-Tau Code Using Automatic Adap-
J., Brodersen, O., Rakowitz, M., Tinoco, E., and Godard, J., “Summary of tation,” Notes on Numerical Fluid Mechanics, edited by W. Nitsche, H.-J.
Data from the Second AIAA CFD Drag Prediction Workshop,” AIAA Paper Heinemann, and R. Hilbig, Vol. 72, Vieweg, Brunswick, Germany, 1998,
2004-0555, Jan. 2004. pp. 178–185.
10 Hemsch, M., and Morrison, J., “Statistical Analysis of CFD Solutions 37 Brodersen, O., “Drag Prediction of Engine-Airframe Interference Ef-
from 2nd Drag Prediction Workshop,” AIAA Paper 2004-0556, Jan. 2004. fects Using Unstructured Navier–Stokes Calculations,” Journal of Aircraft,
11 Brodersen, O., Rakowitz, M., Amant, S., Destarac, D., and Sutcliffe, Vol. 39, No. 6, 2002, pp. 927–935.
M., “Airbus, ONERA, and DLR Results from the 2nd AIAA Drag Prediction 38 Brodersen, O., “Numerische Analyse der aerodynamischen Triebw-
Workshop,” AIAA Paper 2004-0391, Jan. 2004. erksinstallationseffekte an Transportflugzeugen,” DLR, German Aerospace
12 Langtry, R., Kuntz, M., and Menter, F., “Drag Prediction of Engine- Research Center, DLR Technical Rept. 2003-10, Cologne, Germany,
Airframe Interference Effects with CFX 5,” AIAA Paper 2004-0392, May 2003.
Jan. 2004. 39 Vassberg, J., DeHaan, M., and Sclafani, T., “Grid Generation Require-
13 Sclafani, A., DeHaan, M., and Vassberg, J., “OVERFLOW Drag Pre- ments for Accurate Drag Predictions Based on OVERFLOW Calculations,”
diction for the DLR-F6 Transport Configuration: A DPW-II Case Study,” AIAA Paper 2003-4124, June 2003.
AIAA Paper 2004-0393, Jan. 2004. 40 Roache, P. J., “Verification and Validation in Computational Science
14 Rumsey, C., Rivers, S., and Morrision, J., “Study of CFD Variation and Engineering,” Hermosa, Albuquerque, NM, 1998.
on Transport Configurations from the Second Drag-Prediction Workshop,” 41 Tinoco, E. N., “An Assessment of CFD Prediction of Drag and Other
AIAA Paper 2004-0394, Jan. 2004. Longitudinal Characteristics,” AIAA Paper 2001-1002, June 2002.

You might also like