Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

AIAA AVIATION Forum 10.2514/6.

2020-2755
June 15-19, 2020, VIRTUAL EVENT
AIAA AVIATION 2020 FORUM

Flow Distortion Based S-Duct Optimization


using Adjoint Methodology
Vinayak Kamat1 and Vinod Rao2
ANSYS Software Pvt. Ltd., 34/1 Rajiv Gandhi Infotech Park, Hinjewadi, Pune, Maharashtra
411057, India

Min Xu3
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

ANSYS Inc.,10 Cavendish Court, Lebanon, NH 03766

Swati Saxena4
ANSYS Inc., 2645 Zanker Rd., San Jose, CA 95139, USA

In the present work, a duct geometry is optimized to reduce the flow distortion using
adjoint method. The S-duct geometry is from AIAA Propulsion and Aerodynamics Workshop
(PAW [1]). ANSYS Fluent Adjoint solver is used to run the design optimization study. The
workflow involves a 4-step process: CFD run, adjoint calculation, design calculation based on
the sensitivity data and environment constraints, and mesh update. This process is repeated
until the optimum geometry is obtained. The variance of total pressure at the Aerodynamic
Interface Plane (AIP) is chosen as the quantity of interest. It correlates well with the
circumferential and radial distortion index (CDI & RDI) as defined in [2], which indicates
flow distortion at the inlet/exit plane. After 3 design iterations, the variance of total pressure
and the average CDI at AIP was reduced by 20.91% and 8.36% respectively, while the
maximum CDI and RDI showed a drop of 2.15% and 12.37% respectively.

I. Nomenclature
AIP = Aerodynamic Interface Plane
CDI = Circumferential Distortion Index
CFD = Computational Fluid Dynamics
RDI = Radial Distortion Index
po,i = Average Total Pressure of clip i
po,mini = Vertex Minimum Total Pressure of clip i
po,i+1 = Average Total pressure of clip i+1
po,avg = Average Total Pressure of AIP

II. Introduction
Engine intakes are critical component for the stable operation of a propulsion system. S-ducts are among the most
common engine intake system in the modern propulsion system. S-duct often have highly curved centerline due to the
double bend structure with typical cross section growth for static pressure recovery. Curved geometric features give
rise to significant flow separation along with secondary flow vortices. Such flow distortions can cause significant

1
Lead Application Engineer, ANSYS Inc.
2
Senior Application Engineer, ANSYS Inc.
3
Senior Software Developer, ANSYS Inc.
4
Technical Manager, ANSYS Inc., and AIAA Senior Member

Copyright © 2020 by Ansys Inc.. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
impact on the engine performance downstream and induce mechanical vibrations [3]. Compact sizing can further
aggravate the flow distortion and its impact on the engine performance. Optimal design of the S-duct requires detailed
understanding of the flow behavior. Experimental testing offers limited sensor data and significant data reconstruction
is generally required. Design modification often requires multiple prototypes and is an expensive and time-consuming
process.

Physics based simulations can speed up the design process by providing critical performance data and guide
experimental testing and design iterations. CFD simulations can be done to quantify distortion coefficient, pressure
recovery and assess flow behavior [4]. Shape optimization can be a natural extension of the CFD study to minimize
penalties. Most of the optimization methodologies are driven by parameter-based simulations. The parameters can
either be based on the geometric features or on the operating conditions. These goal driven optimization simulations
are often run with large set of design of experiments before arriving at the final optimal design. Such optimization can
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

be done on models with well-defined geometric parameters as shown in [5-6]. For many component designs, it is
difficult to have geometric parameters defined for the complex geometric features such as in some S-Ducts.
In such complex cases, the adjoint solver accomplishes the remarkable feat of calculating the derivative of a single
engineering observation with respect to a very large number of input parameters simultaneously via a single
computation. The engineering observation could be a measure of the system performance, such as the lift or drag on
an airfoil, or the total pressure drop through a system. Most importantly, the derivatives with respect to the geometric
shape of the system are found. An adjoint solver is a specialized tool that extends the scope of the analysis provided
by a conventional flow solver by providing detailed sensitivity data for the performance of a fluid system.
Understanding such sensitivities in a fluid system can provide extremely valuable engineering insight. The sensitivities
of a fluid system provided by an adjoint solver satisfy a central need in gradient-based shape optimization. This makes
an adjoint solver a unique and powerful engineering tool for design optimization.
The work presented in this paper uses the adjoint based optimization approach to optimize the design of a S-duct
and minimize flow separation and flow distortion at a given location. The following sections outline the adjoint
methodology used, S-duct geometry, CFD setup and preliminary optimization results.

III. Gradient-based Shape Optimization using Adjoint Method

If a numerical solution to a flow problem can provide information about the quantities of interest (e.g. variance of
total pressure), then how these quantities are influenced by the shape of the underlying geometry, also called shape
sensitivity, can be obtained very efficiently using an adjoint method. The geometry can then be optimized following
the direction given by the shape sensitivity, design constraints and a gradient-based optimizer.

A. Adjoint Solver
There are two main categories of adjoint solvers: one based on a continuous approach [7,8] and the other on a
discrete approach [9,10,11]. For the continuous approach, the original partial-differential equations are linearized.
Their adjoint is then constructed, discretized and solved. For the discrete approach, the discretized form of the
nonlinear flow problem is linearized fully, giving a linear algebraic system that must be solved to give the adjoint
solution. The pros and cons of the two approaches have been discussed by Nadarajah & Jameson [12] and Collis et al.
[13]. ANSYS Fluent adjoint solver implemented the discrete approach since it offers scalability and reliability and
avoids the imposition of adjoint boundary conditions required in the continuous approach [14].
The governing equations for the fluid system can be expressed as:
𝑅𝜇𝑖 (𝒒𝒗 , 𝒙𝒍 ) = 𝟎 on 𝜇th cell, (1)
where 𝒒 denote the state of the flow in the 𝑣th cell (e.g. pressure and velocity) and 𝒙𝒍 denotes the coordinate of the
𝒗

𝑙th node in the mesh.


The observable, depends on the flow state and the geometry, and is formulated as
𝐽 = 𝐽(𝒒𝒗 , 𝒙𝒍 ). (2)
We can solve the following adjoint equation to determine the adjoint solution 𝒒̃ 𝝁,
𝜇
𝜕𝑅𝑖 𝜕𝐽
̃𝜇𝑖 = 𝑣 ,
𝑞 (3)
𝜕𝑞𝑣𝑗 𝜕𝑞 𝑗
with the Einstein summation convention being implied for repeated indices. After the adjoint solution is obtained, the
variation in the observable is correlated explicitly with changes to the shape by:

2
𝑙
Δ𝐽 = 𝑤
̃Δ𝑥
𝑖 𝑖 , (4)
where
𝜇
𝜕𝐽 𝜕𝑅𝑗
𝑤
̃𝑖 = − ̃𝜇𝑗 ,
𝑞 (5)
𝜕𝑥𝑙𝑖 𝜕𝑥𝑙𝑖
is the derivative of the observable with respect to the 𝑖th coordinate of the 𝑙th mesh node.
It should be noted that a frozen turbulence assumption is made in the current ANSYS Fluent adjoint solver, in
which the effect of changes to the state of the turbulence is not considered when computing sensitivities. It should be
noted that the turbulence qualities are updated in the next flow simulation, therefore the error induced by frozen
turbulence assumption is not accumulated over the optimization. Including the turbulence adjoint equation does not
only increase the number of equations to be solved but also makes the linear system highly poor conditioned [15].
Therefore, the frozen turbulence assumption has been widely used in the literature [16-21]. Despite this assumption,
the sensitivity calculated is still accurate enough for the most industrial application, especially when robust gradient-
based optimization algorithms are used, such as the steepest descent method used in the current study and the
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

conjugate-gradient method. R. Dwight and J. Brezillon have optimized the lift and drag of a transonic RAE2822 airfoil
using both the exact gradient by full linearized adjoint and the approximated gradient based on the frozen turbulence
assumption [21]. They found the two approaches result in almost identical optima with a robust optimizer.
When applied to problems with large mesh counts and complex geometry, adjoint solvers sometimes experience
stability issues. These instabilities can be associated with small-scale unsteadiness in the flow field and/or strong shear
[22], and tend to be restricted to small and isolated regions of the flow domain. Despite the spatial localization of these
instabilities, the linearity of the adjoint problem provides no intrinsic limit on their growth during solution
advancement. Their presence, if not handled, can disrupt the entire adjoint calculation despite the problem occurring
in sometimes just a few cells. Several stabilization schemes are available in ANSYS Fluent to overcome these stability
difficulties [15]. The residual minimization scheme is used in the current study, which uses an algebraic multigrid
preconditioned GMRES-like scheme to approximate the solution by the vectors in a Krylov subspace with minimal
residual. The overall adjoint residual is guaranteed to decrease and is more effective compared to the other stabilization
schemes in the ANSYS Fluent adjoint solver.

B. Design Calculation and Mesh Morphing


Once the adjoint solution is converged, the derivative of the observable with respect to the position of every point
on the surface of the geometry (shape sensitivity) is available, and the sensitivity of the observation to specific
boundary condition settings can also be found. The shape sensitivity can be used to guide intelligent design
modifications to a system through the mesh morpher in the design tool of ANSYS Fluent adjoint solver. The mesh
morphing technology is used for both two- and three-dimensional systems in a two-fold role. The first role is as a
smoother for the shape sensitivity field. The second role is to provide smooth modifications of not only the boundary
mesh, but also of the interior mesh. This approach is very appealing since it avoids the turnover time of remeshing
due to geometry modification and it functions for arbitrary mesh cell types.

A Cartesian region is selected that encompasses all or part of the problem domain. Only the nodes of the
computational mesh that fall inside this region may move because of the morphing operation. Considering a 2D case,
a local coordinate system (u,v) is defined for the region, where 𝑢, 𝑣 ∈ [0,1]. The displacement of the mesh node inside
the design region is defined by the Bernstein polynomials because of their effectiveness in smoothing the adjoint
sensitivity field:
Δ𝑥𝑖𝜈 = Δ𝜂𝑗𝑘 𝑖
𝐵𝑗,𝑁𝑢 (𝑢 𝜈 )𝐵𝑘,𝑁𝑣 (𝑣 𝜈 ), (6)
𝑖
where 𝐵𝑖,𝑁 (𝑢) is the 𝑖th Bernstein polynomial of degree 𝑁 and Δ𝜂𝑗𝑘 denotes the displacement of the (𝑗, 𝑘)th control
point associated with the Bernstein polynomials in the 𝑖th coordinate direction.
During each design iteration of the single/multi-objective optimization, an augmented cost function is introduced
as,
𝑖 2
Δ𝐽 = 𝜆𝑚 𝑤 ̃𝑖 𝑚 Δ𝑥𝑖𝑙 + 𝜆 (Λ − (Δ𝜂𝑗𝑘 ) ), (7)
𝑚
where 𝑤 ̃𝑖 denotes shape sensitivity associated with the 𝑚th observable, 𝜆𝑚 is the weight for the 𝑚th observable, and
Λ is the user-specified scale of the freeform deformation. 𝜆𝑚 is determined explicitly by users or implicitly by the
design tool. The minimization of this augmented cost yields the optimal Bernstein polynomials coefficient Δ𝜂𝑖𝑗𝑘 , which
gives the optimal nodal deformation field Δ𝑥𝑖𝜈 that satisfies the design request. Both the surface and interior mesh are
then updated with the computed optimal nodal displacement.

3
C. Gradient-based Optimizer
A typical adjoint-driven design iteration using ANSYS Fluent adjoint solver consists of four steps: flow simulation,
adjoint simulation, design calculation, and mesh morphing. The gradient-based optimizer in ANSYS Fluent integrates
all these components and automates the optimization process, which uses the steepest decent optimization method in
the design space of the Bernstein polynomial constrained by user defined design constraints. At each design iteration,
the scaling can be determined adaptively along with the optimization process. The optimizer has been tested on various
industrial-strength problems and demonstrated its robustness.

IV. S-Duct Case Setup

A. Flow Simulation
The system is based on S-Duct geometry and is shown in Fig 1. The computational domain consists of S-Duct,
Bell Mouth Inlet, Aerodynamic Interface Plane (AIP) and the Exit. Total Pressure and Static Pressure at the bell mouth
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

inlet are taken as 99,254 Pa and 98,563 Pa respectively. The inlet Total Temperature is assumed to be at 293.95 K.
The total mass flow at the exit is maintained at 1.2547 kg/s. Symmetry plane is used to reduce the computational
domain by half. Walls of the S-duct are modeled with no-slip boundary condition. The computational grid was taken
from the AIAA Propulsion and Aerodynamics Workshop (PAW [1]) and has unstructured mesh with 7.864M cells or
3.398M points. Boundary layers generated near the walls have y+ ~1.

Fig. 1 S-Duct

Air is used as the working fluid with ideal gas law formulation to model the compressibility effects. SST k-ω
model was used to model the turbulence effects. The 3D steady-state compressible CFD simulation was performed
with ANSYS Fluent 2019 version R3 with coupled pressure-velocity algorithm, Green-Gauss node-based gradient
calculation, 2nd order pressure discretization, and the 2nd order upwind discretization for the momentum, energy and
turbulence equations. The flow simulation converged within 4000 iterations. Mach number at the AIP and the static
pressure at the outlet was monitored to ensure complete convergence.

Fig. 2 Convergence of Flow and Energy equations for Baseline model

4
B. Distortion Index
When measuring the uniformity of the flow, a standard measurement arrangement for advanced engine intakes
uses a total pressure rake at the AIP comprising an array of 8 spokes with 5 probes each (Fig. 3). A range of descriptors
is typically considered to assess the total pressure distortion, and the calculations are based on a “ring and rake”
approach (Fig. 3).
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

Fig. 3 8 x 5 rake and ring AIP discretization for inlet flow distortion measurements [2]

The circumferential distortion index, CDI [2] assesses the uniformity of the circumferential total pressure
distribution and is defined as follows:

n −1 p0,i −p0,min p0,i+1 −p0,min


radius i i+1
CDI = Maxi=1 (0.5 [ + ])
p0,avg p0,avg
(8)
where po,avg is the average total pressure, po,i the average total pressure of the pressure distribution of the
i-th ring and po,min the minimal pressure of the i-th ring.

Similarly, the radial distortion can be assessed by the radial distortion index, RDI. RDI definition follows a similar
logic as CDI and is defined as follows:
p0,avg − p0,inner ring p0,avg − p0,outer ring
RDI = Max ( + )
p0,avg p0,avg
(9)
where po.inner ring is the average total pressure of the pressure distribution of the inner ring and p o,outer ring is the
average total pressure at the outer ring.
The Distortion Indices based on the discrete probe points can be inaccurate if the probe does not lie in the region
of secondary flow. In the current case, we would need a higher number of probes to capture the non-uniformity at the
AIP. In the CFD domain, iso-clips were created at different sections of the AIP as shown in Fig 4 and the minimum
and average total pressures were calculated. With this approach, the minimum total pressure at any region in the
domain can be captured.

Fig. 4 Iso-clips created at different sections of the AIP

5
C. Adjoint Based Optimization
In this study, the variance of the total pressure at AIP is selected as the observable instead of CDI and RDI for
simplicity. It is defined as an area-weighted standard deviation of the total pressure,
2
∑𝑓|𝐴𝑓 |(𝑝0,𝑓 −𝑝
̅̅̅̅̅̅)
0,𝑓
𝐽= ∑𝑓|𝐴𝑓 |
(10)
Where 𝑓 denotes facet, 𝐴𝑓 is the facet area, 𝑝0,𝑓 is the total pressure on the facet, and ̅̅̅̅̅
𝑝0,𝑓 is the averaged total
pressure over AIP.

The gradient-based optimizer in ANSYS Fluent can handle multiple objectives optimization problem at multiple
operating conditions. It updates the mesh gradually and automatically to an optimal shape in a series of design
iterations. Within each design iteration, after the adjoint solution is converged, the shape sensitivity is obtained very
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

efficiently, which provides a map across the entire geometry of the effects of moving the surface. Then the design tool
in ANSYS Fluent can use the computed shape sensitivity to find an optimal design change satisfying one or more
goals or constraints and predict the resulting changes in observable values. In current study, we are looking at reducing
the variance of total pressure by 10% for a single design iteration. Fig 5 shows the region undergoing deformation in
the design tool. The convergence history of 3 CFD design iterations is shown in Fig 6. The adjoint data set provides
rich information. Two post-processing variables are often analyzed and can provide better engineering insights into
the design process:

1. Shape Sensitivity Magnitude: Shape sensitivity magnitude is the magnitude of the sensitivity of the
observable with respect to a deformation applied to the mesh. In view of the large range of values possible
for the shape sensitivity magnitude, a convenience function that plots log 10 of the magnitude is provided.
This allows the importance of the surfaces in a domain to be ranked more easily based on how they affect
the observation of interest when they are reshaped.

2. Normal Optimal Displacement: Normal Optimal Displacement is the projection of the shape sensitivity into
the design space. Therefore, it can be considered as a smoothed/surrounding-averaged shape sensitivity.

We will discuss these two fields in the result section.

Fig. 5 Region selected for Morphing

6
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

Fig. 6 Convergence of all Flow Iterations

V. Results

A. Baseline Geometry
The postprocessing is done on the full model taking advantage of the symmetry boundary condition. Baseline
model indicates strong separation upstream of AIP giving rise to the secondary flow. The location of AIP is plotted in
Fig 7. Fig 8 shows the plot of total pressure at the AIP and Fig 9 shows the plot of total pressure overlaid with
streamlines. We can see a gradual increase of velocity profile along the S-Duct from Fig 10. This non-uniform
distribution of velocity can be attributed to the flow separation.

Fig. 7 Baseline Design: AIP location in S-Duct

7
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

Fig. 8 Baseline Design: Total Pressure contour plot at the AIP

Fig. 9 Baseline Design: Total Pressure Contour plot at AIP overlaid with streamlines from Inlet

8
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

Fig. 10 Baseline Design: Velocity Contour plot at different sections along the S-Duct

B. Shape Optimization
The Gradient based optimizer went through three design iterations to arrive at the final optimized shape. Fig 11
displays contours of large sensitivity regions on the S-Duct walls.

Fig. 11 Log10 Shape Sensitivity Magnitude on the S-Duct

Fig 12 shows the contours of Normal Optimal Displacement on the S-Duct walls. Positive displacement values
indicate node movement into the fluid volume. Negative values indicate node movement away from the fluid volume.
The Normal Optimal Displacement vectors are displayed in Fig 13.

9
Fig. 128 Contours of Normal Displacement on the S-Duct
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

Fig. 13 Vectors of Normal Displacement on the S-Duct

The Optimal Design showed smaller Total Pressure loss region at AIP in Fig 14 indicating reduction in variance
of the Total Pressure. As shown in Fig 15, the Optimal Design showed uniform distribution of velocities at sections
along the S-Duct compared to the Baseline scenario. Fig 16 shows how the geometry has morphed at different sections
along the S-Duct.

Fig. 14 Baseline vs Optimal Design: Total Pressure Distribution at AIP

10
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

Fig. 15 Baseline vs Optimal Design: Velocity distribution at different sections along the S-Duct

Fig. 16 Baseline vs Optimal Design: Geometry at different sections along the S-Duct

Fig 17 showing quantitative results of variance of the total pressure, CDI & RDI for Baseline & Optimal Design.
The variance of total pressure & average CDI at AIP was reduced by 20.91% & 8.36% resp. The maximum CDI &
RDI showed a drop of 2.15% and 12.37% respectively.

11
Fig. 17 Baseline vs Optimal Geometry: Variance of Total Pressure, CDI & RDI

VI. Conclusion
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

Industry standard pressure distortion metrics like CDI & RDI are typically used to compare the performance of
engine intakes. It is found that the total pressure variance constraint correlates well with both CDI and RDI and can
be used as an indicator to run these kinds of optimization studies to minimize flow distortion at the inlet/exit plane.
Adjoint Gradient-based Shape Optimizer was used to optimize the shape of S-Duct to reduce Variance of Total
Pressure at AIP. The results demonstrated significant reduction in variance of total pressure, CDI & RDI values at the
AIP.

VII. References

1. 5th AIAA Propulsion Aerodynamics workshop ( https://paw.larc.nasa.gov/)


2. Zachos, P.K., Frascella, M., MacManus, D.G., and Prieto, D.G., “Pressure Flow Field and Inlet Flow Distortion Metrics
Reconstruction from Velocity Data,” AIAA Journal, Vol. 55, Issue 9, 2017, pp. 2929-2941, DOI:10.2514/1.J055585.
3. Society of Automotive Engineering, Inlet Total-Pressure-Distortion Considerations for Gas-Turbine Engines, Warrendale,
PA, 1999.
4. J. Selvanayagam,C. Aliaga, J. Stokes and B. Sasanapuri. “Numerical Simulation of an Aircraft Engine Intake S-Duct
Diffuser”, 53rd AIAA/SAE/ASEE Joint Propulsion Conference.
5. F. Furlan, N. Chiereghin, T. Kipouros, E. Benini and M. Savill, "Computational design of S-Duct intakes for distributed
propulsion",," Aircraft Engineering and Aerospace Technology: An International Journal, vol. 86, no. 6, pp. 473 - 477, 2014.
6. N. Chiereghin, L. Guglielmi, M. Savill, T. Kipouros, E. Manca, A. Rigobello, M. Barison, and E. Benini. “Shape
Optimization of a Curved Duct with Free Form Deformations”. 23rd AIAA Computational Fluid Dynamics Conference
7. Xu, M., Wei, M., Li, C. & Dong, H. 2015 Adjoint-based optimization of flapping plates hinged with a trailing-edge flap.
Theoretical and Applied Mechanics Letters, vol. 5, pp. 1–4.
8. Xu, M. & Wei, M., 2016 Using adjoint approach to study flapping wings. Journal of Fluid Mechanics, vol. 799, pp. 56-99.
9. Hill, D.C., The Automatic Generation of Adjoint Solutions for a General Purpose Flow Solver. 46th AIAA Aerospace
Sciences Meeting and Exhibit, vol. AIAA-2008-904, Reno, Nevada (US)
10. Caridi, D., Hill, D.C., Xu, M. & Braun, M., Optimization of Confined Internal Flows, NAFEMS, adjoint CFD methods in
industrial application and research, 2016 ,Wiesbaden, Germany.
11. Verma, I., Hill, C., and Xu, M., “Multi-Objective Adjoint Optimization of Flow-Bench Port Geometry,” SAE Technical
Paper 2018-01-0772, 2018, doi:10.4271/2018-01-0772.
12. Nadarajah, S. & Jameson, A. 2000 A comparison of the continuous and discrete adjoint approach to automatic aerodynamic
optimization. AIAA Paper 2000-667
13. Collis, S. S., Ghayour, K., Heinkenschloss, M., Ulbrich, M. & Ulbrich, S. 2001 Towards adjoint-based methods for
aeroacoustic control. AIAA Paper 2001-0821.
14. ANSYS Fluent user’s guide. Release 2019R3. ANSYS, Inc., 2019.
15. Peter, J. E., & Dwight, R. P. (2010). Numerical sensitivity analysis for aerodynamic optimization: A survey of approaches.
Computers & Fluids, 39(3), 373-391.
16. E. Nielsen, W. Anderson, Aerodynamic design optimization on unstructured meshes using the Navier-Stokes equations,
AIAA Journal 37 (11) (1999) 185–191
17. C. Kim, C. Kim, O. Rho, Effects of constant eddy viscosity assumption on gradient-based design optimization, in: AIAA
Paper Series, Paper 2002-0262, 2002.
18. J. Peter, J. Mayeur, Improving accuracy and robustness of a discrete direct differentiation method and discrete adjoint
method for aerodynamic shape optimization, in: Proceedings of ECCOMAS, Egmond ann Zee, 2006.
19. A. Jameson, N. Pierce, L. Martinelli, Optimum aerodynamic design using the Navier-Stokes equations, in: AIAA Paper
Series, Paper 97-101, 1997.
20. B. Soemarwoto, The variational method for aerodynamic optimization using the Navier-Stokes equations, Tech. Rep. 97-71,
ICASE (Dec 1997).

12
21. R. Dwight, J. Brezillon, Effect of approximations of the discrete adjoint on gradient-based optimization, AIAA Journal 44
(12) (2006) 3022–3071.
22. Joshua A. Krakos and David L. Darmofal, Effect of Small-Scale Output Unsteadiness on Adjoint-Based Sensitivity, AIAA
Journal 48 (2010).
Downloaded by UNIVERSITY OF NEW SOUTH WALES on June 22, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2755

13

You might also like