THE DEVELOPMENT OF THREE-DIMENSIONAL ADJOINT …

163
The Pennsylvania State University The Graduate School Department of Aerospace Engineering THE DEVELOPMENT OF THREE-DIMENSIONAL ADJOINT METHOD FOR FLOW CONTROL WITH BLOWING IN CONVERGENT-DIVERGENT NOZZLE FLOWS A Dissertation in Aerospace Engineering by Nidhi Sikarwar c 2015 Nidhi Sikarwar Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy May 2015

Transcript of THE DEVELOPMENT OF THREE-DIMENSIONAL ADJOINT …

The Pennsylvania State University

The Graduate School

Department of Aerospace Engineering

THE DEVELOPMENT OF THREE-DIMENSIONAL ADJOINT

METHOD FOR FLOW CONTROL WITH BLOWING IN

CONVERGENT-DIVERGENT NOZZLE FLOWS

A Dissertation in

Aerospace Engineering

by

Nidhi Sikarwar

c© 2015 Nidhi Sikarwar

Submitted in Partial Fulfillmentof the Requirements

for the Degree of

Doctor of Philosophy

May 2015

The dissertation of Nidhi Sikarwar was reviewed and approved1 by the following:

Philip J. MorrisProfessor of Aerospace EngineeringDissertation AdviserChair of Committee

D. K. McLaughlinProfessor of Aerospace Engineering

Mark D. MaughmerProfessor of Aerospace Engineering

Daniel C. HaworthProfessor of Mechanical Engineering

George A. LesieutreProfessor of Aerospace EngineeringHead of the Department of Aerospace Engineering

1Signatures on file in the Graduate School.

iii

Abstract

The noise produced by the low bypass ratio turbofan engines used to power fighter air-craft is a problem for communities near military bases and for personnel working in closeproximity to the aircraft. For example, carrier deck personnel are subject to noise expo-sure that can result in Noise-Induced Hearing Loss which in-turn results in over a billiondollars of disability payments by the Veterans Administration. Several methods havebeen proposed to reduce the jet noise at the source. These methods include microjetinjection of air or water downstream of the jet exit, chevrons, and corrugated nozzleinserts. The last method involves the insertion of corrugated seals into the divergingsection of a military-style convergent-divergent jet nozzle (to replace the existing seals).This has been shown to reduce both the broadband shock-associated noise as well as themixing noise in the peak noise radiation direction. However, the original inserts weredesigned to be effective for a take-off condition where the jet is over-expanded. Thenozzle performance would be expected to degrade at other conditions, such as in cruiseat altitude. A new method has been proposed to achieve the same effects as corrugatedseals, but using fluidic inserts. This involves injection of air, at relatively low pressuresand total mass flow rates, into the diverging section of the nozzle. These “fluidic inserts”deflect the flow in the same way as the mechanical inserts. The fluidic inserts representan active control method, since the injectors can be modified or turned off dependingon the jet operating conditions. Noise reductions in the peak noise direction of 5 to 6dB have been achieved and broadband shock-associated noise is effectively suppressed.There are multiple parameters to be considered in the design of the fluidic inserts. Thisincludes the number and location of the injectors and the pressures and mass flow ratesto be used. These could be optimized on an ad hoc basis with multiple experiments ornumerical simulations. Alternatively an inverse design method can be used. An adjointoptimization method can be used to achieve the optimum blowing rate. It is shown thatthe method works for both geometry optimization and active control of the flow in orderto deflect the flow in desirable ways.

An adjoint optimization method is described. It is used to determine the blowing dis-tribution in the diverging section of a convergent-divergent nozzle that gives a desiredpressure distribution in the nozzle. Both the direct and adjoint problems and their asso-ciated boundary conditions are developed. The adjoint method is used to determine theblowing distribution required to minimize the shock strength in the nozzle to achieve a

iv

known target pressure and to achieve close to an ideally expanded flow pressure.

A multi-block structured solver is developed to calculate the flow solution and associatedadjoint variables. Two and three-dimensional calculations are performed for internal andexternal of the nozzle domains. A two step MacCormack scheme based on predictor-corrector technique is was used for some calculations. The four and five stage Runge-Kutta schemes are also used to artificially march in time. A modified Runge-Kuttascheme is used to accelerate the convergence to a steady state. Second order artificialdissipation has been added to stabilize the calculations. The steepest decent methodhas been used for the optimization of the blowing velocity after the gradients of thecost function with respect to the blowing velocity are calculated using adjoint method.Several examples are given of the optimization of blowing using the adjoint method.

v

Table of Contents

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiv

Chapter 1.. Introduction and Background . . . . . . . . . . . . . . . . . . . . . . 11.1. Motivation for Noise Reduction and Available Methods . . . . . . . . 11.2. Fluidic Inserts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71.3. Use of Adjoint Method . . . . . . . . . . . . . . . . . . . . . . . . . . 141.4. The Adjoint Approach . . . . . . . . . . . . . . . . . . . . . . . . . . 161.5. Passive Control Using Adjoint Method . . . . . . . . . . . . . . . . . 171.6. Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201.7. Original Contribution of the Thesis . . . . . . . . . . . . . . . . . . . 21

Chapter 2.. General Development of the Adjoint Method . . . . . . . . . . . . . 232.1. General Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242.2. Outline of the Design Procedure . . . . . . . . . . . . . . . . . . . . 272.3. Discrete and Continuous Approaches . . . . . . . . . . . . . . . . . . 292.4. Steady and Time Dependent Problems . . . . . . . . . . . . . . . . . 312.5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

Chapter 3.. Numerical Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343.1. Numerical Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

3.1.1. MacCormack Scheme . . . . . . . . . . . . . . . . . . . . . . . 353.1.2. Central Difference Schemes . . . . . . . . . . . . . . . . . . . 373.1.3. Time Marching Schemes . . . . . . . . . . . . . . . . . . . . . 373.1.4. Local Time Stepping . . . . . . . . . . . . . . . . . . . . . . . 41

3.2. Multi-Block Topology . . . . . . . . . . . . . . . . . . . . . . . . . . 433.3. Artificial Dissipation . . . . . . . . . . . . . . . . . . . . . . . . . . . 443.4. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

vi

Chapter 4.. Direct and Adjoint Characteristics Analysis . . . . . . . . . . . . . . 484.1. Grid singularities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 484.2. Characteristics and Interface Boundary Conditions for the Direct

Flow Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514.3. Adjoint Characteristics and Interface Boundary Conditions . . . . . 524.4. Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 544.5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624.6. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

Chapter 5.. Parameterization of the Control . . . . . . . . . . . . . . . . . . . . . 645.1. The Mathematical Development . . . . . . . . . . . . . . . . . . . . . 655.2. Optimization with Parameterization of Blowing . . . . . . . . . . . . 715.3. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 775.4. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

Chapter 6.. Adjoint Control of Nozzle Flow with Surface Blowing . . . . . . . . . 796.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 796.2. The Mathematical Development with Wall Blowing Control . . . . . 806.3. Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

6.3.1. Adjoint Outflow Boundary Condition . . . . . . . . . . . . . 896.3.1.1. Subsonic Outflow . . . . . . . . . . . . . . . . . . . 896.3.1.2. Supersonic outflow . . . . . . . . . . . . . . . . . . . 90

6.3.2. Adjoint Inflow Boundary Condition . . . . . . . . . . . . . . . 916.3.3. Adjoint Slip Wall Boundary Condition . . . . . . . . . . . . . 926.3.4. Sensor Boundary Condition . . . . . . . . . . . . . . . . . . . 936.3.5. Actuator Boundary Condition . . . . . . . . . . . . . . . . . . 94

6.4. The Optimization Procedure . . . . . . . . . . . . . . . . . . . . . . 966.5. Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 976.6. Two-dimensional Optimization of Blowing . . . . . . . . . . . . . . 97

6.6.1. Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 986.6.2. Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1016.6.3. Three-dimensional Optimization of Blowing . . . . . . . . . . 107

6.7. External Flow Calculations . . . . . . . . . . . . . . . . . . . . . . . 1186.8. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1246.9. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

Chapter 7.. Conclusions and Future Work . . . . . . . . . . . . . . . . . . . . . . 126

Appendix A.. The Solver Development . . . . . . . . . . . . . . . . . . . . . . . . 131A.1. Code Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131A.2. Input requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135A.3. Output files . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

Appendix B.. Governing and Adjoint Equations in Curvilinear Coordinates . . . 140

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

vii

List of Tables

6.1. The decay of the cost function with design cycles for two-dimensionalcalculations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

6.2. Three-dimensional calculations: The decay of the cost function with de-sign cycles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

A.1. Input Parameters: the values of fluid properties, convergence criteria,inlet and ambient conditions, initial conditions and numerical parameters. 135

A.2. Boundary Conditions: the number that specify the type of boundarycondition at a given boundary . . . . . . . . . . . . . . . . . . . . . . . . 136

viii

List of Figures

1.1. Far-field narrow-band noise spectra of a supersonic jet operating at Mj= 2.0 showing turbulent low frequency mixing noise, screeching tone andbroadband shock associated noise [40]. . . . . . . . . . . . . . . . . . . . 3

1.2. Typical subsonic broadband jet noise spectra Sound Pressure Level vs.Strouhl Number for a Mj = 0.9 jet [36]. . . . . . . . . . . . . . . . . . . 3

1.3. The corrugated seals for noise reduction as described by Seiner [41]. . . 61.4. The experimental set up for fluidic inserts. There are three inserts placed

symmetrically in the divergent section of the nozzle. There are two in-jectors per fluidic insert. . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.5. The CAD design of fluidic inserts. . . . . . . . . . . . . . . . . . . . . . 81.6. The acoustic spectrum of the baseline nozzle, a) corrugated seals and

b) fluidic inserts , Mj = 1.36, NPR = 3.0, TTR = 3.0 and injectionmass flow ratio = 3.8%. . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1.7. The schematic of fluidic inserts. Flow streamlines shift away from thewall to produce a new effective area ratio. . . . . . . . . . . . . . . . . . 11

1.8. The shifting of the streamlines in the nozzle with distributed blowing. . 111.9. Numerical shadowgraph with different fluidic inserts. Unheated, design

Mach number 1.65 and nozzle pressure ratio NPR = 4.58. The injectionpressure ratio = 4.58. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.10. The generation of counter rotating streamwise vortices at each actuator. 131.11. The adjoint design approach. . . . . . . . . . . . . . . . . . . . . . . . . 171.12. The distribution of pressure (with respect to total pressure Po) along

the nozzle centerline. The red and blue lines show the initial and finalpressure respectively. The desired pressure is shown by black symbols. . 19

1.13. Pressure contours inside the nozzle domain. The upper half shows theinitial flow and lower half shows the final flow. . . . . . . . . . . . . . . 19

1.14. The geometry of the nozzle. Calculations were performed for only halfthe domain. The red line shows the initial geometry. The green lineshows the final geometry. The black line shows the geometry that givesthe desired pressure distribution. . . . . . . . . . . . . . . . . . . . . . . 20

2.1. The main steps of the optimization procedure. . . . . . . . . . . . . . . 29

ix

2.2. Adjoint solution with time. The unsteady adjoint solution moves back-wards in time and the direct solution has to be stored at all times. . . . 32

3.1. The centerline singularity for circular grid and a multi-block ‘H’ typegrid generation to avoid centerline singularity. . . . . . . . . . . . . . . . 43

4.1. The generation of multi-blocks showing grid singularities at the blockinterface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4.2. Multi-block ‘H’ type grid generation to avoid the centerline singularity. . 514.3. The mesh and multi-block topology for propagation of a Gaussian pulse

in a cube. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 554.4. Propagation of a Gaussian pulse through block interfaces at a) t = 0s,

b) t = 0.001s, c) t = 0.002s and d) t = 0.004s. . . . . . . . . . . . . . . 574.5. Comparison of analytic (solid) and numerical (symbol) solutions for Gaus-

sian pulse propagation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 584.6. The cross-section of the convergent-divergent nozzle with five blocks to

avoid centerline singularity. . . . . . . . . . . . . . . . . . . . . . . . . . 604.7. The comparison of direct and adjoint pressure contours. . . . . . . . . . 604.8. The comparison of direct and adjoint axial velocity contours. . . . . . . 614.9. The comparison of direct and adjoint cross-stream velocity contours. . . 614.10. The comparison of direct and adjoint density contours. . . . . . . . . . . 61

5.1. Nozzle contour, sensor and actuator regions. . . . . . . . . . . . . . . . . 655.2. The variation in pressure contours and streamlines, in the divergent sec-

tion of the nozzle, as the amplitude of the blowing velocity vb is variedto 0, 10, 30 and 60 m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

5.3. Variation of the amplitude of the blowing vb in m/s with design cycles. 745.4. Decay of the cost function in Pa2 with design cycle. . . . . . . . . . . . 755.5. Removal of shock with adjoint design cycles. The sensor region is shown

as a box in the diverging section of the nozzle. . . . . . . . . . . . . . . 77

6.1. Variation of the blowing distribution with design cycles. . . . . . . . . . 996.2. Variation of the pressure distribution on the upper nozzle wall with design

cycles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1016.3. Two-dimensional nozzle domain and boundary conditions. . . . . . . . . 1026.4. The actuator and sensor regions on the divergent section of nozzle. . . . 1036.5. Pressure distibution for initial shocked flow. Area ratio = 1.12, NPR =

1.5, no blowing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1056.6. Ideally expanded ‘desired’ pressure distribution. NPR = 1.5, area ratio

= 1.04, no blowing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1056.7. Nozzle pressure distribution after two design cycles. NPR = 1.5, area

ratio = 1.12, with blowing. . . . . . . . . . . . . . . . . . . . . . . . . . 1066.8. Initial, first and desired pressure distributions on the nozzle centerline. . 1066.9. Nozzle showing surface mesh. . . . . . . . . . . . . . . . . . . . . . . . . 1086.10. Nozzle multi-block grid structure. . . . . . . . . . . . . . . . . . . . . . . 1096.11. The location of actuators (the blowing ports). . . . . . . . . . . . . . . . 109

x

6.12. The pressure distribution when there is no blowing in the divergent section. 1116.13. The pressure distribution with blowing in the divergent section. . . . . . 1116.14. Introduction of counter rotating stream-wise vortices due to blowing. . . 1126.15. Variation of the pressure difference on the nozzle wall between the current

and desired pressure distribution with design cycle. . . . . . . . . . . . . 1146.16. The decay of cost function with design cycles. . . . . . . . . . . . . . . . 1156.17. Pressure distribution on nozzle centerline with design cycles. . . . . . . 1166.18. Mach number distribution on nozzle centerline with design cycles. . . . 1166.19. Blowing velocity normal to the wall used to generate the desired pressure

distribution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1176.20. Blowing velocity normal to the wall as given by fourth design cycle. . . 1176.21. The placement of blocks when the external domain is included. . . . . . 1196.22. The mesh including the external domain. . . . . . . . . . . . . . . . . . 1206.23. The location of symmetrically placed actuators in the nozzle diverging

section. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1206.24. The Mach number contours for the nozzle operating with 4.58 nozzle

pressure ratio and no blowing. . . . . . . . . . . . . . . . . . . . . . . . . 1216.25. The Mach number contours for the nozzle operating with 4.58 nozzle

pressure ratio and the desired blowing. . . . . . . . . . . . . . . . . . . . 1226.26. The Mach number contours after the first design cycle for the nozzle

operating with 4.58 nozzle pressure ratio. . . . . . . . . . . . . . . . . . 1226.27. The Mach number contours after the second design cycle for the nozzle

operating with 4.58 nozzle pressure ratio. . . . . . . . . . . . . . . . . . 1236.28. The Mach number contours after third design cycle for the nozzle oper-

ating with 4.58 nozzle pressure ratio. . . . . . . . . . . . . . . . . . . . . 123

A.1. The code structure to find direct and adjoint solutions and optimizationof the cost function - continued to next figure . . . . . . . . . . . . . . . 133

A.2. The code structure to find direct and adjoint solutions and optimizationof the cost function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134

xi

Nomenclature

Symbol Description

ρ density

p static pressure

T temperature

ξ, η, ζ orthogonal coordinate system in computa-tional domain

δij Kronecker delta

γ ratio of specific heats

λ spectral radius

σ CFL number

θ angle of injection

∆t numerical time step

c speed of sound

D nozzle exit diameter

et total energy per unit mass

F1, F2, F3 inviscid fluxes in the governing equations

xii

M Mach number

Md design Mach number

Mj jet Mach number

Po total or stagnantion pressure

To total or stagnantion temperature

pa ambient pressure

R gas constant of air at STP

u1, u2, u3 axial, vertical and spanwise velocity

(x, y, z) three directions in the physical space

Q vector of adjoint variables

vb blowing velocity

α design parameters

J cost function or objective function

L Lagrange functional

Subscript

0 total value

i, j, k indices in three computational directions

u derivative with respect to u

Superscript

′ the perturbation with respect to the flowvariables

Abbereviations

NPR nozzle pressure ratio

xiii

BBSAN broadband shock associated noise

OASPL overall sound pressure level

SPL sound pressure level

TTR total temperature ration

IPR injection pressure ratio

CD convergent-divergent

xiv

Acknowledgments

First and foremost, I would like to express my sincere gratitude to my advisor Prof.Philip Morris for his continuous support of my Ph.D. study and research. His im-mense patience and his vast knowledge are the reasons this thesis was possible.His continuous guidance has helped me in research and writing of this thesis. Hisqualities like hard work and punctuality have lasting impact on me as a researcher.

Besides my advisor, I would like to thank the rest of my thesis committee: Prof. D.K. McLaughlin, Prof. Mark D. Maughmer, and Prof. Dan Haworth for reviewingmy research work and providing with their valuable inputs.

I would especially like to thank the Penn State aeroacoustics experimental group,Dr. D. K. McLaughlin and Russell Powers for providing with the experimentaldata. My sincere thank goes to cluster admin Kirk Heller for his promptnessin troubleshooting the computer and cluster related problems. I would also liketo thank the Aerospace Department staff for their helpful attitude and friendliness.

I thank my labmates Yongle Du and Swati Saxena for their help in making meunderstand the concepts better and helping me with their deep knowledge of nu-merical implementation. I would also like to thank my fellow labmates VeeraManek and Monica Christiansen for making the lab a cheerful place.

Thanks to my special friends: Vishesh Karwa for endless discussions on how todo better research and Arnab Sengupta for his support and for lending me hiscomputer for finishing parts of my Ph.D., my awesome roommates RushmithaThiyagrajan and Rashmi Shukla for their dealing with my frustration and theirsupport in the most difficult times of my Ph.D. I would also like to thank Tanushree,Ragini, Himanshu, Kadappan, Nikhil, Ashish, Rakesh and Pankaj for making mystay at Penn State enjoyable. I would like to give a special mention of my lifelongfriends: Shraddha, Richa, Sangeeta and Vijaya for always believing in me and fortheir constant encouragement.

xv

My gratitude towards my husband Ashutosh Chauhan is inexpressible. This workwould not have been possible without him. His constant love and support comesin the most appropriate ways. He has helped me in all possible ways: encouragingme, pushing me, proof reading for me, cooking for me and just listening to mewhenever needed.

I would like to thank my best friend of all time, my sister Dr. Nimisha Singh whohas been an integral part of my life. Her guidance and support throughout my lifehas been invaluable to me. I dedicate this thesis to my parents: Mr. J. P. Sikarwarand Mrs. Sampada Sikarwar. My mother’s encouragement and belief are the solereasons for me being able to even imagine achieving anything in my life.

1

Chapter 1Introduction and Background

1.1 Motivation for Noise Reduction and Avail-

able Methods

The noise produced by the low bypass ratio turbofan engines used to power tactical

fighter aircraft is a problem for communities near military bases and for personnel

working in close proximity to the aircraft. For example, carrier deck personnel

are subject to noise exposure that can result in Noise-Induced Hearing Loss. This

results in disability payments that are made by the Veterans Administration each

year, costing over a billion dollars. Scientists have been struggling to achieve noise

reduction from tactical aircrafts for decades. The higher noise levels from tactical

aircraft put restrictions on the airbase operations and expansions. The number

of hours that Navy personnel can be exposed to such high levels is restricted.

According to military standards and regulations, a decrease in noise generation

by 3 dB can double the allowable exposure time. However, these restrictions are

not necessarily enforced. A recent study performed at Harvard School of Public

Health [7] shows that older people exposed to aircraft noise, especially at high

levels, may face increased risk of being hospitalized for cardiovascular disease. The

Federal Aviation Administration (FAA) pursues a program of aircraft noise con-

trol in cooperation with the aviation community. Noise control measures include

noise reduction at the source. That is, the development and adoption of quieter

2

aircraft, soundproofing and buyouts of buildings near airports, operational flight

control measures, and land use planning strategies. Similar noise reduction efforts

are being pursued by the Department of Defense.

Aircraft noise originates from three main sources: 1. Aerodynamic noise, 2. En-

gine and other mechanical noise and 3. Noise from aircraft systems. One of the

major contributors to the noise generated from the aircraft on take-off is the en-

gine exhaust noise. Other sources of sound from jet engines include fan noise,

combustion noise (low frequency and non-directional); and internal engine compo-

nent noise such as the turbine, struts and splitters (peaks at 60o from downstream

jet axis). The noise generated by an aircraft in the approach configuration has

two main contributions, the airframe noise resulting from turbulent flow over solid

structures like wings, slats, flaps and landing gears and the engine noise generated

from the jet and fan flows. The latter propagates in both the upstream and the

downstream directions. The dominant component of jet noise is due to the mixing

of large-scale turbulent eddies. The noise characteristics of supersonic jets are very

different from that of subsonic jets. Supersonic jets operating at off-design con-

ditions contain distinct components of turbulent mixing noise, broadband shock

associated noise and secreech. Figure 1.1 shows the different noise components in

an imperfectly expanded supersonic jet. On the other hand, subsonic jet noise is

mainly due to turbulent mixing and has a uniform broadband spectrum as shown

in figure 1.2.

3

Fig. 1.1 Far-field narrow-band noise spectra of a supersonic jet operating at Mj = 2.0

showing turbulent low frequency mixing noise, screeching tone and broadband shock

associated noise [40].

Fig. 1.2 Typical subsonic broadband jet noise spectra Sound Pressure Level vs. Strouhl

Number for a Mj = 0.9 jet [36].

Military aircraft engines are low bypass ratio turbofan engines that generate thrust

by exhausting a very high velocity jet. This jet is highly turbulent and hence there

4

are high levels of noise associated with the turbulent mixing of the jet plume with

the ambient air. Another noise component that exists is the broadband shock asso-

ciated noise (BBSAN) that occurs due to the fact that the engine does not always

operate at on-design conditions. When there is a difference between the ambi-

ent pressure and the nozzle exit pressure, there exist shock cells in the jet plume

through which the pressure adjusts to the ambient pressure. The interaction of

the turbulence in the jet shear layer with these shock cells (regions of alternating

pressure and temperature) results in the occurrence of BBSAN. In imperfectly ex-

panded supersonic jets, two additional components of noise arise: screech tones

and broadband shock associated noise.

Several methods have been proposed to reduce noise generated by high speed jets.

A recent summary of methods for high speed jet noise reduction is given by Morris

and McLaughlin [29]. The methods include microjet injection of air [1] or water [23]

downstream of the jet exit, chevrons [5, 13, 4], fluidic chevrons [21], plasma actua-

tors [39], beveled nozzles [47], conical and contoured nozzles with plasma actuators

[19] and corrugated nozzle inserts [42, 41]. The chevrons [5, 13] are used to reduce

noise from separate-flow turbofan engines. Mechanical chevron serrations at the

nozzle trailing edge generate axial vorticity that enhances jet plume mixing and

consequently reduces far-field noise. Fluidic chevrons as described by Kinzie et al.

[21] generated with air injected near the nozzle trailing edge create a vorticity field

similar to that of the mechanical chevrons and allow more flexibility in controlling

acoustic and thrust performance than a passive mechanical design. In addition, the

design of such a system has the future potential for actively controlling jet noise

by pulsing or otherwise optimally distributing the injected air. Plasma actuation

represents another active method to control aircraft noise as shown by Samimy

[39]. Some advantages of these methods are simplicity, the absence of mechanical

moving parts, and fast response. Experiments have shown that plasma actuators

can be used for flow control and airframe noise reduction. It is shown by Munday

5

et al. [32] that conical nozzles can exhibit reduction in far field noise generation.

Kim et al. [19] showed that contoured and conical nozzles with plasma actuators

can be helpful in noise reduction from Mach 1.65 jets. The reduction at shallow

polar angles is related to the decrease in the peak mixing noise level in both jets.

Experiments with microjets were performed with and without chevrons by Alk-

islar [1] and the effect of microject injection on vorticity generation and turbulent

characteristics was studied. It was observed that at an early location in the jet,

the influence of the microjet injection at the chevron tips is considerable. At the

injection locations within the shear layer, strong counter rotating vortex pairs were

created in addition to, but separate from, the ones created by the chevrons at the

chevron base. Their strength was comparable to those produced by the chevrons

alone; however, the pairs have an opposite sense of translation; their induced ve-

locities are opposite, with the chevron’s being away from the main jet and the

microjet’s being towards the main jet, at a location in the jet away from the noz-

zle exit both microjets and chevrons start to dissipate and lose their signature on

the flow.

The final method involves the insertion of corrugated seals into the diverging sec-

tion of a military-style convergent-divergent jet nozzle (to replace the existing

seals) as described by Seiner [42]. Figure 1.3 show the seals in a military style

nozzle divergent section. Seiner et al. [42] presented experimental and computa-

tional data for the noise changes with the use of two passive design modifications

that can be implemented with very little change in the nozzle geometry. They are

corrugated internal divergent secondary flap seals and external divergent primary

flap chevrons. Experimental measurements indicated that the noise was reduced

by 4dB relative to the baseline nozzle when corrugated seals were used and by 2

dB when chevrons were used. This technique has been shown to reduce both the

broadband shock-associated noise as well as the mixing noise in the peak noise ra-

diation direction. The corrugated seals work in two ways - they change the effective

6

area ratio of the nozzle and they introduce streamwise vortices in the jet plume.

The change in the effective area ratio helps to achieve nearly ideally expanded flow

even when the nozzle is working at an off-design condition at the time of take off

when the nozzle generally operates under-expanded. A reduction in broadband

shock associated noise is then observed. The introduction of streamwise vortices

helps to increase the turbulent mixing and thus reduces the mixing noise. The

original inserts were designed to be effective for take-off conditions where the jet

is over-expanded. The nozzle performance would be expected to degrade at other

conditions, such as in cruise at altitude.

Fig. 1.3 The corrugated seals for noise reduction as described by Seiner [41].

Recently, Morris et al.[30, 24] built on the corrugated seal concept, but instead

used fluidic inserts. This involves injection of air, at relatively low pressures and

total mass flow rates, into the diverging section of the nozzle. These fluidic inserts

deflect the flow in the same way as the mechanical inserts. But the fluidic inserts

represent an active control method, since the injectors can be modified or turned

off depending on the jet operating conditions. Details of the development of fluidic

inserts are given in the next section.

7

1.2 Fluidic Inserts

Fluidic inserts are noise reduction devices proposed at Penn State by Morris et

al.[30]. The idea is to achieve similar effects as the corrugated seals by using

carefully distributed blowing on the divergent section of the nozzle wall. The

bypass air would be used at relatively low pressure and low mass flow rates. It

is expected that there will not be much difference in the engine performance due

to the low mass flow rate. The experimental set-up and CAD design of fluidic

inserts is shown in figures 1.4 and figurecad respectively. The experiments have

been conducted at Penn State by McLaughlin et al.. An efficient methodology for

the simulation of heated jet is made possible since helium-air mixtures are used to

simulate the heated jets [8, 27]. Noise reductions in the peak noise direction of 4

to 5 dB have been achieved at model scale and broadband shock-associated noise

is effectively suppressed. The reduction in noise is dependent on the ratio of mass

flow rates for the injection and the core jet flow. Experiments on fluidic inserts

with simulated forward flight were performed at Penn State by Powers et al. [38].

Fluidic inserts in the presence of forward flight streams have shown a peak mixing

noise reduction of 4 dB and broadband shock associated noise reduction of 3 dB.

Comparisons of the acoustic spectra of the fluidic inserts nozzle and corrugated

seals nozzle have been made. Figure 1.6 shows the typical acoustic spectrum for

the baseline nozzle, the nozzle with three corrugated seals placed symmetrically

and the nozzle with three fluidic inserts places symmetrically on the nozzle wall.

The design Mach number of the nozzle is 1.56 and mass flow rate of the fluidic

inserts with respect to the core mass flow rate is 3.8%. The acoustic reductions

generated by the fluidic inserts are same as that of the corrugated seals. The

reduction in OASPL is of the order of 3 - 4 dB. The screech tones are removed

completely for both fluidic inserts and corrugated seal nozzles.

8

Fig. 1.4 The experimental set up for fluidic inserts. There are three inserts placed

symmetrically in the divergent section of the nozzle. There are two injectors per fluidic

insert.

Fig. 1.5 The CAD design of fluidic inserts.

9

(a) Corrugated seals (b) Fluidic inserts

Fig. 1.6 The acoustic spectrum of the baseline nozzle, a) corrugated seals and b)

fluidic inserts , Mj = 1.36, NPR = 3.0, TTR = 3.0 and injection mass flow ratio =

3.8%.

There are many advantages to the fluidic insert noise reduction technique. The

rate of blowing can be changed during flight. Thus, it would be very flexible.

Various blowing rates and distribution could help in making the engine operate

at near on-design conditions for all flight regimes. This would not only reduce

the noise but would also help in improving the performance. It is planned to use

the bypass air as the blowing fluid. Hence the implementation of the technique

would be convenient. Few new mechanical changes in the existing nozzle would be

10

required, apart from the installation of the flow control.

A schematic of how the fluidic inserts affect the flow is shown in figure 1.7. The

incoming flow due to the blowing on the nozzle wall shifts the streamlines away

from the nozzle wall, thus changing the effective exit to throat area ratio, as shown

in figure 1.7. The change in exit to throat area ratio changes the effective operating

conditions for the nozzle. Figure 1.8 shows the shifting of the flow by the shift

in streamlines caused by the two actuators. Early numerical experiments were

performed by Morris et al. [31] to see the effects of the inserts on the jet flow.

The modification in the jet plume due to the placement of the fluidic inserts is

shown in a numerical shadowgraph in figure 1.9 [34]. The nozzle pressure ratio

NPR is 4.58 and the design Mach number is 1.65. Figure 1.9a shows the baseline

unheated jet. Figure 1.9b shows a shadowgraph with one fluidic insert on the nozzle

wall. The shock cell structure is tilted from the centerline in the case with one

fluidic insert. Figure 1.9c shows the shadowgraph with two fluidic inserts placed

symmetrically. The shock cell structure for the two fluidic inserts is narrowed

towards the centerline. The placement of fluidic inserts in both cases change the

shock cell structure in the nozzle jet flow significantly. An optimized location of

the fluidic inserts with an optimum mass flow rate (or injection pressure ratio)

can be incorporated to achieve shock free flow field even when the nozzle is over-

expanded. This technique can be used to achieve a reduction in the broad band

shock associated noise from the nozzles operating at off-design conditions.

11

Fig. 1.7 The schematic of fluidic inserts. Flow streamlines shift away from the wall to

produce a new effective area ratio.

Fig. 1.8 The shifting of the streamlines in the nozzle with distributed blowing.

12

(a) The baseline nozzle [31]

(b) With one fluidic insert [31]

(c) With two fluidic inserts [31]

Fig. 1.9 Numerical shadowgraph with different fluidic inserts. Unheated, design Mach

number 1.65 and nozzle pressure ratio NPR = 4.58. The injection pressure ratio = 4.58.

13

The fluidic inserts are also known to introduce streamswise vortices into the jet

plume. Figure 1.10 shows the generation of counter-rotating vortices. There are

two actuators (or fluidic inserts) placed symmetrically shown by the black lines

and the vorticity generation from these actuators is shown by the counter-rotating

vortices. The generation of those counter rotating vortices is seen by the positive

and negative streamwise vorticity on the two sides of each actuator. The generation

of counter rotating vortices produce the effects similar to the corrugated seals.

The turbulent mixing in the jet shear layer is enhanced due to the introduction of

streamwise vortices and a reduction in mixing noise is achieved due to the increased

turbulent mixing in the jet shear layer.

Fig. 1.10 The generation of counter rotating streamwise vortices at each actuator.

An optimum placement of the fluidic inserts with optimum mass flow rates could

result in a greater reduction in the broadband shock associated and turbulent mix-

ing noise. Optimization of the placement and rate of injection is a challenge to

14

obtain the best effects from the fluidic inserts. There are multiple parameters to

be considered in the design of the fluidic inserts. These include the number and

location of the injectors and the pressures and mass flow rates to be used. The

optimization of these parameters is the next goal in the development of fluidic

inserts. These could be optimized on an ad hoc basis with multiple experiments

or numerical simulations. Alternatively an inverse design method could be used.

The adjoint approach is an inverse design optimization method that can be used

to optimize a problem with many design parameters. The objective of this the-

sis is to develop an adjoint method to find that rate of blowing in the nozzle to

achieve desired effects. The desired effects are quantified by the pressure distri-

bution associated with the nozzle. It is shown that an adjoint design approach

can be used to achieve a desired pressure distribution. The primary advantage

of using an adjoint method is that computational cost associated with the adjoint

approach is considerably less than that of the traditional inverse design approaches.

The next section provides some background on adjoint design methods. The origi-

nal use of adjoint methods in the area of aerodynamic design have been associated

with shape modification. A study of the use of adjoint methods for shape opti-

mization is followed by a description of the adjoint approach to incorporate active

flow control optimization [44, 45].

1.3 Use of Adjoint Method

Adjoint equations have been used for some time in optimal control theory. Li-

ons [26] used adjoint methods to develop an optimization technique for systems

that are governed by partial differential equations. Pironneau [37] used the adjoint

method for the first time in fluid dynamics for design work, but Jameson [14, 15]

revolutionized the use of adjoint methods for aerodynamic design. He developed

continuous adjoint methods for different governing equations including the poten-

tial, the Euler, and the Navier-Stokes equations. An optimal design is one that

15

optimizes the defined cost function within given constraints. Jameson showed that

the adjoint method can be used to modify the shape under consideration such as

airfoil, wing or full aircraft to achieve a reduction in the cost function. The cost

function could be either the lift or drag coefficients or some difference relative to

a desired flow behavior. Joslin et al.[16] developed an automated methodology for

active flow control using adjoint methods. A method for the suppression of two-

dimensional instability waves for a flat plate boundary layer was incorporated, by

implementation of blowing and suction at the wall. The problem was treated as

a control problem with the rate of blowing (or suction) acting as the control. It

was found that no prior knowledge of the instability characteristics was required

to optimize the control (blowing). Noise control based on adjoint methods has

also been explored by Wei and Freund [49]. Freund [10] attempted to understand

the workings of the jet noise mechanism with the use of adjoint methods. These

studies were based on the aerodynamic optimization approach by Jameson and

turbulence control by Bewley [2]. Wei and Freund minimized the noise radiated

by a two-dimensional mixing layer for a line of observers above the layer. The

time dependent Navier-Stokes equations and the associated adjoint equations were

solved to determine the time history of various controllers near the origin of the

mixing layer. Wei and Freund showed that a two dimensional mixing layer can be

made quieter based on the concept that the flow also works as a source of sound.

The actuation was taken as a general inhomogeneity in the flow equations with

support near the inflow boundary. It did not correspond to any particular actuator.

The near-nozzle jet was modeled by a randomly excited two-dimensional mixing

layer. Mixing layers between streams with Mach numbers 0.2 and 0.9, which is

a subsonic case, were studied. The controls were implemented as general source

terms. That is, body forces, mass sources and internal energy sources. The internal

energy control was found to be the most effective with a reduction of 11 dB in the

noise intensity. It was shown that the noise reduction was not achieved by noise

cancellation, but from a genuine change in the flow as a noise source. The changes

16

observed in the flow gross features were very subtle. However, the decomposition

of the flow into empirical eigenfunctions showed that the downstream advection of

turbulent structures takes place more uniformly with the excitation. Kim et al.[18]

extended these ideas to a Mach 1.3 jet with some success. Kim tried to reduce

the turbulent jet noise by implementing various flow controls that would change

the flow turbulence. The benefit of this approach is that the ‘actuator region’ can

exist anywhere in the domain. They found the gradient of the cost function with

respect to the control using adjoint methods. This gradient was then used to find

the local optimum of the cost function with respect to the control. The conjugate

gradient method was used to minimize the cost function. Far field noise, beyond

the computational domain, was predicted using a solution to the Ffowcs Williams

and Hawkings equation [3]. The total number of optimization parameters was 280

x 106. A reduction of 2.6% was observed in the value of cost function over the

entire domain.

1.4 The Adjoint Approach

The adjoint optimization method is a constrained optimization technique. The

adjoint approach works as a ‘feedback - control’ approach where the feedback from

the sensor is communicated back to the control. Then the control is modified based

on the feedback and generates a new flow. The same process is repeated until the

change in the flow is small. Figure 1.11 illustrates this cycle. The adjoint equations

are developed in such a way that the flow governing equations are considered as

the constraints for the optimization of a given cost function. The resultant adjoint

equations produce the gradients of the cost function with respect to the design

parameters without having to calculate the change in flow variables with respect

to the design parameters. This saves considerable computational cost. The adjoint

method works as a control technique in which a sensor provides a measure of the

cost function. The adjoint equations are then solved for the sensitivity of the

17

respective cost function to the control. This information is then passed to an

actuator or controller, so that the flow is driven towards the desired state.

Fig. 1.11 The adjoint design approach.

1.5 Passive Control Using Adjoint Method

The use of an adjoint design method for nozzle shape design is illustrated in this

section, the details of this analysis are given in Sikarwar [43]. Adjoint methods were

implemented to find the optimum change in the geometry such that a defined cost

function was minimized. The formulation was based on Jameson’s [15] approach.

The geometry was dependent on a set of design parameters and the cost function

was defined as the difference between a desired and calculated nozzle centerline

pressure. The adjoint approach was validated by calculating the desired pressure

difference using a known set of design parameters. That is, for a known geometry

the centerline pressure was calculated and was considered as the desired pressure

distribution. Then the geometry was perturbed from this desired geometry by

changing the design parameters. Then the adjoint method was used to recover the

18

desired pressure distribution. Figure 1.12 shows the distribution of pressure on the

nozzle centerline. The initial pressure distribution is shown by a red line. The final

pressure distribution is shown by a blue line. The desired pressure distribution is

shown by the symbols. All these calculations are performed for a fixed value of

nozzle pressure ratio NPR = 1.5. Initially, the shock is at the nozzle exit and keeps

moving inside the nozzle with each design cycle. The final and desired pressure

distributions do not agree exactly but they are very close. The initial value of the

cost function is 4.39 x 106 Pa which drops to a value of 2.13 x 105 Pa after 32 design

cycles for a decrease of 95.14% in the cost function. The full flow solutions given

by pressure contours, for the initial (upper) and final (lower) geometries inside the

nozzle domain, are shown in Figure 1.13. The difference in the two solutions is the

shock location. The nozzle is sonic at the throat, hence, the flow upstream of the

throat remains the same; whereas, the flow downstream changes with each design

cycle.

19

Fig. 1.12 The distribution of pressure (with respect to total pressure Po) along the

nozzle centerline. The red and blue lines show the initial and final pressure respectively.

The desired pressure is shown by black symbols.

Fig. 1.13 Pressure contours inside the nozzle domain. The upper half shows the initial

flow and lower half shows the final flow.

20

The nozzle geometries are shown in figure 1.14. The red line shows the initial

geometry. The black line shows the desired geometry and the green line shows the

final geometry given by the adjoint design method.

Fig. 1.14 The geometry of the nozzle. Calculations were performed for only half the

domain. The red line shows the initial geometry. The green line shows the final geometry.

The black line shows the geometry that gives the desired pressure distribution.

1.6 Thesis outline

The objective of this thesis is to develop an adjoint approach for the optimization

of blowing in the diverging section of a convergent-divergent (CD) nozzle. The

adjoint approach developed in this thesis considers the blowing velocities as the

control parameters that are optimized. The next chapter describes the general

theory and development of adjoint optimization. The details of the discrete and

continuous adjoint approaches and steady and unsteady problems are given in this

chapter. The adjoint optimization procedure for a typical design cycle is described.

Chapter 3 describes the numerical methods implemented in the development of the

solver. The details of the numerical schemes, artificial dissipation and convergence

acceleration techniques are given in this chapter. The solver developed for the

21

proposed research is a multi-block structured solver. The multi-block mesh topol-

ogy includes block interfaces where there may be grid singularities between two

adjacent blocks. These grid singularities introduce numerical errors into the solu-

tion. A characteristics-based block interface boundary condition is implemented

to overcome this issue for both direct and adjoint equations. The details of the

boundary conditions as well as discussion on the nature of the adjoint variables are

given in Chapter 4. Two different adjoint approaches have been developed in the

course of this study. The difference in the two approaches is in the consideration

of the control. Chapter 5 describes the first approach where the blowing velocity

is given by a combination of basis functions multiplied by weighting coefficients.

These coefficients are considered as the design parameters for the optimization.

The second approach where the blowing velocities are considered as free controls,

is described in Chapter 6. The velocity components at each grid point in the actu-

ator region serve as the control parameters for the adjoint approach. The details

of the solver developed for this research are given in Appendix A. The conclusions

and ideas for future work are given in chapter 7.

1.7 Original Contribution of the Thesis

The traditional uses of adjoint methods have been in the area of shape optimiza-

tion. There has been little work on the use of adjoint methods for active flow

control. Though some examples exist in the literature where an adjoint optimiza-

tion technique has been used for the optimization of blowing, those problems are

restricted to two-dimensional incompressible calculations for simple geometries.

Adjoint methods have not been used to optimize control in nozzle flows. The work

presented in this thesis is novel in the sense of the application and also in the sense

of the development of the method. A new adjoint method to optimize distributed

blowing on the nozzle wall to achieve a desired pressure distribution in the nozzle

has been developed. Two different adjoint approaches have been developed and

22

implemented.

An adjoint approach, where distributed blowing in a nozzle is considered as control,

has been developed. The blowing velocity at each grid point of an actuator region

has been considered as an independent control parameter. New adjoint boundary

conditions based on the direct flow boundary conditions have been developed. For

the numerical implementation of the new approach for complex geometries, two-

and three-dimensional direct and adjoint solvers have been developed. A solver to

calculate the sensitivities of the cost function with respect to the control parame-

ters using the direct and adjoint solutions has been developed for the optimization.

These sensitivities are used to determine the blowing distribution that minimizes

the cost function.

To perform three-dimensional calculations for more complicated geometries a multi-

block grid topology has been used. A new technique to treat the block interfaces

while performing the adjoint calculations has been developed. This characteristics-

based block interface boundary condition for the adjoint equations has been de-

veloped and implemented for the first time. The numerical implementation of the

direct and adjoint block interface boundary conditions has been incorporated into

the direct and adjoint solvers. The direct and adjoint solvers are equipped to work

with multi-block structured grids with arbitrary orientations. The communication

of data at arbitrarily indexed block boundaries has been incorporated.

The next chapter provides a general development of the adjoint optimization

method.

23

Chapter 2General Development of the AdjointMethod

The scope for the uses of the adjoint method is very wide. As the general interest

in the use of the adjoint method is increasing, a variety of problems are being

solved using this approach. Some of the many areas where the adjoint method has

been used for the optimization include statistics [22], weather predictions [17], ge-

ometry optimization [11, 15], mesh improvements [35] and error analysis [46]. The

use of the adjoint approach varies by disciplines, goals etc. However, the general

approach remains the same. In this chapter the basic theory of the adjoint method

is described. Since there are two ways to develop the adjoint equations - discrete

and continuous - an introduction to both these approaches is given in this chapter.

This chapter provides the basis for the research work presented in this thesis. The

particular approaches used in this thesis can be better understood once the basic

approach is clear. The approach can be modified to solve specific problems using

the flow model and cost function specific to the problem. If the desired pressure

distribution inside a nozzle is known, an adjoint method can be used to find the

design parameters corresponding to the desired pressure distribution. For the su-

personic case, when there are shocks in the nozzle, this method can be used to

find a set of design parameters such that the shock strength is minimized, and by

inference, so would be the broadband shock associated noise. The various goals

can be facilitated by changing the cost function, which is a measure of the desired

24

behavior. Various design problems can be addressed by changing the control pa-

rameters or design parameters, but the basic approach for a general cost function

and general design parameters remains the same and this is described in the next

section.

2.1 General Approach

The basic theory of the adjoint method explained in this section is based on Jame-

son’s [14] approach. It is a constrained optimization method where the governing

equations are considered as constraints. The theory presented in this section is

generic to the development of the adjoint method and can be implemented for any

set of governing equations, for any design parameters, and for any choice of cost

function (within the mathematical admissible limits). The optimization technique

is based on the gradient descent method. However, it is computationally expen-

sive to calculate the gradients directly. Adjoint methods can be used to calculate

these gradients using a significantly reduced computational cost. The adjoint equa-

tions are developed such that the changes in the flow variables with respect to the

control/design parameter are not required to be calculated while determining the

sensitivity of the cost function to the design parameters. The elimination of the

need to calculate the change in the flow variables with respect to the change in

design parameters reduces the computational cost required to calculate the gradi-

ents with respect to the design parameters significantly.

The starting point is a system of nonlinear partial differential equations governing

the flow in the computational domain. Design parameters are denoted by α. The

flow equations and the cost function are dependent on the flow solution vector U

and the design variables α.

The governing equations are given by,

R = R(U, α) = 0. (2.1)

25

The cost function is a measure of the desired behavior. The adjoint method is

used to minimize or maximize the cost function. It could be a measure of lift or

drag over an airfoil or it could be a measure of pressure drop in a pipe. There are

several choices for the cost function based on the statement of the problem. The

cost function is also a function of the flow variables U and design parameters α.

In general, the cost function J is given by,

J = J(U, α) (2.2)

Any change in the design parameters would cause a change in the flow variables

as well, which would cause a change in the cost function. The changes in the cost

function and the governing equations with respect to the flow and design variables

are given by,

δJ =∂JT

∂αδα +

∂JT

∂UδU (2.3)

and,

δR =∂R

∂αδα +

∂R

∂UδU = 0 (2.4)

The augmented change in the cost function is the change in cost function when

the change in the governing equations is considered a constraint. The augmented

change in cost function when relation (2.4) is used is given by,

δJ =∂JT

∂αδα +

∂JT

∂UδU −QT

(∂R

∂αδα +

∂R

∂UδU

)(2.5)

Here, Q is the vector of adjoint variables and acts as a Lagrange multiplier. It

should be noted that augmentation of the cost function does not change the value

of the cost function since the governing equations are always satisfied and are equal

to zero. Expression (2.5) can be rearranged such that the terms dependent on the

change in the flow variables δU are brought together as are the terms dependent

on the change in the design parameter δα.

26

δJ =

(∂JT

∂α−QT ∂R

∂α

)δα +

(∂JT

∂U−QT ∂R

∂U

)δU (2.6)

The goal is to eliminate the requirement to calculate the change in flow variables

δU . This is achieved by choosing the adjoint variables to satisfy the equation,

∂JT

∂U−QT ∂R

∂U= 0 (2.7)

This equation is known as the “adjoint equation”, and can be solved to calculate

the value of the vector Q. The boundary conditions for the adjoint equations are

derived based on the boundary conditions for the direct flow equations.

The sensitivity of the objective function δJ is independent of the flow solution

perturbation δU and is given by,

δJ =

(∂JT

∂α−QT ∂R

∂α

)δα (2.8)

or, the sensitivity of cost function with respect to the design parameters is given

by,δJ

δα=

(∂JT

∂α−QT ∂R

∂α

)(2.9)

Once the values of the adjoint variables are known, equation (2.9) is used to cal-

culate the sensitivities of the cost function with respect to the design parameters.

These sensitivities are used in any gradient optimization method to find the opti-

mum values of the design parameters α.

This section has explained the general formulation of the adjoint method. The

approaches for the particular cases of interest in this thesis are explained in more

detail in chapters 5 and 6.

27

2.2 Outline of the Design Procedure

An outline of the optimization procedure is summarized in figure 2.1. More steps

can be added to the process based on requirements such as when there is shock

in the solution. In this case an extra step can be added for smoothing the cost

function. When geometry optimization is considered, an extra step to generate

a new mesh is added, because a new mesh has to be generated for a new set of

design parameters. In the present research the geometry has been kept fixed and

it is not required to generate a new mesh for a new set of design parameters. The

basic steps of the design process are,

1. Fix the initial set of design parameters:

First the initial values of the design parameter are chosen. In theory, this

choice is arbitrary but the convergence of the adjoint design cycle is dependent

on it so a wise guess is advisable. In the case when the blowing velocity is

considered as the design parameter, the cycle is started either with a zero

initial blowing velocity or a very small value.

2. Solve the flow equations for this set of design parameters:

The flow equations are then solved for the design parameters fixed in the

previous step. The flow equations are solved by marching in time to a steady

state and an appropriate initial condition for the flow variables is required.

3. Solve the adjoint equations:

Once the flow solution is obtained, the adjoint equations are solved for the

fixed set of design parameters. The adjoint equations are also solved by

marching in time to a steady state, and again it is important to chose an

appropriate initial condition for the adjoint variables.

4. Calculate the gradients:

The gradients are dependent on both the adjoint and flow solutions and can

be calculated using direct algebraic relations.

28

5. Find the new set of design parameters using the gradients:

The gradients calculated in the previous step are used in a gradient based

optimization technique such as steepest decent or conjugate gradient methods

to calculate the next set of design parameters.

6. Repeat the cycle until convergence:

The cycle is repeated until the cost function reduces to a required minimum.

29

Solve flow equations

STOP

Solve adjoint equations

New design parameters

Calculate gradients

Calculate Cost Function

If C.F. < Tolerance

Fig. 2.1 The main steps of the optimization procedure.

2.3 Discrete and Continuous Approaches

The adjoint equations are dependent on the direct flow equations. As the form of

direct flow equations considered varies, the development of the adjoint variables

also varies. There are two main approaches in which the adjoint equations can

30

be developed. These are the discrete and continuous approaches. The discrete

approach is such that the discretized direct flow equations are considered first

and then the set of adjoint equations is developed. The adjoint equations derived

in this manner are already discretized and no further discretization is required.

The continuous approach is the one where the continuous direct flow equations

are considered. The continuous adjoint equations are then discretized using the

discretization scheme of choice. In most cases the continuous adjoint equations

are discretized using the same numerical scheme as the direct equations. The

differences in the two approaches can be summed up as calculating the exact

gradient of the inexact cost function (discrete adjoint) or the inexact gradient of

the exact cost function (continuous) [33].

In the appropriate limits, when space and time intervals go towards zero, the

discrete adjoint equations converge to the continuous adjoint equations and the

discrete adjoint boundary conditions converge to the continuous adjoint boundary

conditions.

lim∆x,∆t→0

Discrete Adjoint Equations⇒ Continuous Adjoint Equations

lim∆x,∆t→0

Discrete Adjoint BC⇒ Continuous Adjoint BC(2.10)

There are pros and cons to both approaches. The continuous approach provides

an insight into the behavior and nature of the equations but it may be difficult to

obtain a stable solution. The boundary conditions for the continuous approach are

derived separately. The discrete approach is relatively easy to implement especially

because no direct derivation of the boundary conditions is required for the discrete

adjoint variables, but it is difficult to infer the meaning of the adjoint variables.

The present work is based on the continuous approach, since it is intended to

obtain an insight into the form of the adjoint equations and the behavior of adjoint

variables. A characteristic analysis of the adjoint equations was possible because

of the choice of the continuous approach. The nature of the adjoint characteristics

31

and the development of boundary conditions based on the adjoint characteristics

are given in chapter 4.

2.4 Steady and Time Dependent Problems

The adjoint approach can be used for both, steady and unsteady problems. As

expected, the adjoint variables corresponding to the unsteady flow variables are

unsteady in nature. The time dependent adjoint solution can be obtained as shown

in figure 2.2. The adjoint equations are solved backwards in time, the details of

the mathematical development of the unsteady equations are given in chapter 6.

Adjoint variables are dependent on the direct flow variables, when the adjoint

solution is marched backwards in time from the final time of calculation, t = tf , to

the initial time of calculation, t = ti, the direct flow solution has also to march from

the initial time, t = ti to final time, t = tf . Therefore, before starting the adjoint

calculation, it is required to have the direct flow solution available at all times. This

condition increases the memory requirement tremendously, especially for problems

where the grid sizes are large. Researchers have tried several manipulations to

handle the storage of solutions better. These include storing the solution at few

time intervals and solving for the remainder [48, 25]. But, none of the known

manipulations are very effective. Keeping the memory requirements in mind, only

the steady state solutions of adjoint and direct equations are obtained in this thesis.

Even though artificial time marching is incorporated to march to the steady state,

only the steady state solution for the direct flow equations is stored and is kept

fixed when marching to the steady state of the adjoint equations.

32

t = ti

t = ti+Δt

t = ti+2Δt

Flow solution U(ti)

Flow solution U(ti+Δt) Adjoint solution Q(ti+Δt)

Adjoint solution Q(ti)

Flow solution

t = ti+nΔt

t = tf

Flow solution U(ti+2Δt)

Flow solution U(ti+nΔt)

Flow solution U(tf)

Adjoint solution Q(ti+2Δt)

Adjoint solution Q(ti+nΔt)

Adjoint solution Q(tf)

Fig. 2.2 Adjoint solution with time. The unsteady adjoint solution moves backwards in

time and the direct solution has to be stored at all times.

2.5 Summary

In this chapter, the general approaches to the theory of the adjoint equations have

been explained for the purpose of providing the reader with an overview. A generic

approach is described that can be used for a generic set of equations, as it is im-

portant to ascertain the well posedness of the problem first. The design procedure

for one typical cycle has been described. The step-wise procedure of optimization

includes the solution of the flow and adjoint equations. The sensitivities of the cost

function with respect to the design parameters are calculated using the flow and

adjoint solutions. These gradients are then used in a gradient-based optimization

technique to calculate an improved set of design parameters. An introduction to

the discrete and continuous adjoint equations has been given. This thesis is based

on the continuous adjoint equations with the intention of obtaining knowledge

about the nature of adjoint variables. Both steady and unsteady problems can

33

be solved using the adjoint approach. However, the unsteady adjoint equations

have vast memory requirements to store the flow solution at all times. This is the

primary reason why steady problems are solved in the present thesis.

The development of the adjoint equations for specific problems and their solutions

are given in chapters 5 and 6. The next chapter describes the numerical method

used to find the solution of the adjoint and direct problems. Details of time

marching and discretization techniques are given. Several schemes are used to

accelerate the convergence and to stabilize the solution. A detailed description of

these schemes is also given. The grid indexing is arbitrary at the block interfaces

mentioned above and there can be numerical issues if differences are not accounted

for. A technique to generalize the grid indexing for each block is discussed in the

next chapter.

34

Chapter 3Numerical Method

The goal of the research described in this thesis is to use an adjoint optimization

method for the optimization of the blowing velocity in the divergent section of

a nozzle to achieve a modification of the flow with the ultimate goal of reducing

noise. The nature of the adjoint and flow equations is very similar as discussed in

more details in chapter 4. This property is the basis of making the assumption

that both adjoint and direct flow equations can be solved on the same mesh. The

direction of propagation of the adjoint characteristics are opposite to the direction

of propagation of the direct flow equations and the adjoint equations are solved

backwards in time. The time step is negative for the time integration of the adjoint

equations but the numerical scheme for time integration is same as for the direct

flow equations. This chapter describes in detail the numerical techniques that have

been implemented for the solution of the direct and adjoint problems.

A direct flow solver and a corresponding adjoint solver have been developed for

implementing the adjoint optimization. The solver works on multi-block structured

grids. An example of the grid topology is shown in figure 4.2. The block interfaces

of multi-block meshes have grid singularities that can introduce numerical errors

in the solution. Such block interfaces are treated with the use of characteristics-

based boundary conditions. These conditions have been developed for the direct

as well as the adjoint flow equations. The adjoint flow equations behave differently

35

from the direct flow equations so the treatment is different for the adjoint block

interfaces. These boundary conditions are discussed in detail in chapter 4. The

main numerical attributes of the solver include the implementation of local time

stepping for better convergence rate and the use of implicit residual smoothing

for additional faster convergence. Various numerical schemes have been used for

the discretization of the spatial derivatives and for time marching. These schemes

include a MacCormack scheme, second and sixth order central differencing and

traditional and modified Runge-Kutta time marching algorithms. The details of

these schemes are given in the next section.

3.1 Numerical Schemes

This section describes the spatial discretization and time marching schemes. There

are several options for spatial discretization. These options include MacCormack

and central differencing schemes. These schemes are described in the following

sections.

3.1.1 MacCormack Scheme

It has been mentioned that one of the goals of this thesis is to provide a method for

generating shock-free flows for the purpose of reducing broadband shock associated

noise. This has been a key point to consider while choosing the numerical scheme.

The MacCormack scheme is very good for capturing shocks in a solution. The

governing flow equations are solved in conservative form for the same reason. A

second order MacCormack scheme includes two steps: the predictor step where

the solution is predicted using a forward difference in space at a half time step and

the corrector step, where the solution is corrected using the predicted half step

solution and a backward difference scheme. These steps are described below.

The governing flow equations in conservative form are given by,

∂U

∂t+∂Fi∂xi

= 0 (3.1)

36

The following two steps explain how the MacCormack scheme is used to discretize

the above set of equations.

Predictor step:

Predicted values of the time derivatives are calculated in this step and this value

is then used to correct the values in the next step.

∂U

∂t

∗= −

(∂Fi∂xi

)FD

(3.2)

Where, (∂Fi∂xi

)FD are the spatial derivatives calculated using a forward differencing

scheme. The discretization is illustrated at the point ‘j′ with grid spacing ‘h′ as,

(∂Fi∂xi

)FD

=F j+1i− F j

i

h(3.3)

Corrector step:

The time derivatives calculated above are used to correct the values of the time

derivative as follows,∂U

∂t

∗∗= −

(∂F ∗

i

∂xi

)BD

(3.4)

Where, (∂F ∗

i

∂xi)BD are the spatial derivatives calculated using a backward differencing

scheme. The discretization is illustrated at the point ‘j′ with grid spacing ‘h′ as,

(∂F ∗

i

∂xi

)BD

=F ∗ji− F ∗j−1

i

h(3.5)

Note that the predicted values of fluxes F ∗i

have been used here to calculate the

corrected time derivatives.

These two time derivatives are now used to calculate the average time derivative,

given by,∂U

∂t=

1

2

(∂U∗

∂t+∂U∗∗

∂t

)(3.6)

This value of time derivative is used to calculate the values of the solution variables

at the next time step. The MacCormack scheme is second order in time and space

37

and is very good at capturing discontinuities in the flow. The above discretiza-

tion is illustrated here for the direct equations of motion. The same process is

implemented for the discretization of the adjoint equations. The adjoint equations

can not be written in conservative form and are used as given for discretization

purposes.

3.1.2 Central Difference Schemes

There are several central difference schemes that have been developed by re-

searchers for various numerical requirements. A central difference scheme can

be used with an appropriate time marching scheme for marching iteratively to the

steady state. There are a few advantages of using these schemes over using the

MacCormack scheme. The most important advantage is that local time stepping

can be used for a faster convergence to steady state solution. It was observed

that when combined with MacCormack scheme, local time stepping does not al-

ways converge to the correct solution. To increase the rate of convergence while

maintaining the order of accuracy of the solution, several more schemes have been

implemented. There are options available for using either a second or sixth or-

der central difference scheme based on the specific requirement of the simulation.

When acoustic predictions are not being performed, a second order scheme can be

used. The stencil for this scheme is given by,

∂Fii∂xi

=1

2

Fii+1− Fii−1

h, (3.7)

where h is the uniform grid spacing between i− 1, i and i+ 1 grid points.

3.1.3 Time Marching Schemes

Even though the equations and solutions considered for this work are steady, artifi-

cial time marching has been implemented to march the solution to a steady state.

A number of time marching schemes have been implemented and a comparison

between them has been carried out as described in the following section. First, an

38

introduction of these schemes is given followed by a comparison of the convergence

rates for the various time marching schemes. Finally, local time stepping is used

to increase the rate of convergence to a steady state. An introduction to local time

stepping is provided in the final part of this section.

Traditional Runge-Kutta Time Marching

Runge-Kutta methods are a family of implicit and explicit time marching schemes

that were developed in 1900’s by the German mathematicians C. Runge and M.

W. Kutta. These time marching schemes were developed for finding approximate

numerical solutions of ordinary differential equations. These methods are high

order non linear methods that can be used for marching the solution from one

time step to the next. A generalization of the method is possible for k stage time

marching where there are k stages between two time steps n and n+ 1. Consider

an equation of the type,

U ′ = H (3.8)

where,

H = H(U, t) (3.9)

and U ′ is the derivative of U with respect to time t. To find the solution at the n+1

time step, when the solution at the n time step is know, the following algorithm

39

is followed,

U (1) = Un (3.10)

U (2) = Un + ∆tα2H(1) (3.11)

U (3) = Un + ∆tα3H(2) (3.12)

. (3.13)

. (3.14)

U (k) = Un + ∆tαkH(k−1) (3.15)

U (n+1) = Un + ∆tk∑k=1

βkH(k) (3.16)

Here, H(k) depicts the time derivative at the stage k, and is calculated using the

right hand side of the differential equation calculated using the solution at that

stage: H(k) = H(U (k)). The coefficients βk must satisfy the consistency equation,

k∑i=1

βi = 1 (3.17)

A traditional Runge-Kutta method is the 4 stage classical Runge-Kutta method,

often referred to as the RK4 method. The coefficients for this method are given

by,

α2 =1

2, α3 =

1

2, α4 = 1

β1 = β4 =1

6, β2 = β3 =

1

3

(3.18)

Relaxed Runge-Kutta Time Marching

For any number of stages, there are an infinite number of methods possible for time

accurate time marching with maximum order of accuracy but when a steady state

40

solution is sought, the convergence rate and stability are more important than ac-

curacy and a modified method can be used. It is also important to have the largest

allowable time step when performing steady calculations. The stability region can

be extended by relaxing the coefficients of the RK method. The coefficients for a

relaxed 4 stage Runge-Kutta method are given by,

α1 = 0.375, α2 =1

2, α3 = α4 = 1 (3.19)

A compact scheme does not use the derivatives at the various stages to calculate the

derivative at final stage. Only the previous step derivative is required to calculate

the derivative at any stage. This helps to reduce the required memory as the same

array can be overwritten when calculating the derivative at any stage.

A five stage RK compact method can also be used for higher allowable time

steps. The five stages are given by,

U (1) = Un (3.20)

U (2) = Un + ∆tα2H(1) (3.21)

U (3) = Un + ∆tα3H(2) (3.22)

U (4) = Un + ∆tα4H(3) (3.23)

U (5) = Un + ∆tα5H(4) (3.24)

The coefficients αi for this method are given as,

α1 = 0, α5 = c2, α4 =c3

α5

, α3 =c4

α5α4

, α2 =c5

α4α4α3(3.25)

41

Where ci are,

c1 = 1,

c2 = 1/2,

c3 = 0.166558,

c4 = 0.0395041,

c5 = 0.00781071,

(3.26)

3.1.4 Local Time Stepping

The stability criterion provides a relationship between the cell size and time step at

any grid point. This criterion is used to calculate the time step based on the local

cell size. This method helps to increase the convergence rate. Steady calculations

are performed using local time stepping. This can not be used to perform time

accurate calculations as the time step varies from grid point to grid point. Using

this method, larger cells will move faster towards convergence because the time

step will be larger in such cells. The local time step is chosen to be the minimum

of the allowable time steps based on viscous and inviscid calculations (though only

the inviscid criterion is required for Euler equations).

First, the contravariant velocities in the computational domain are calculated,

contraU =∂ξ

∂xu+

∂ξ

∂yv +

∂ξ

∂zw (3.27)

contraV =∂η

∂xu+

∂η

∂yv +

∂η

∂zw (3.28)

contraW =∂ζ

∂xu+

∂ζ

∂yv +

∂ζ

∂zw (3.29)

Cell dimensions are given by,

ξmag =

√∂ξ

∂x

2

+∂ξ

∂y

2

+∂ξ

∂z

2

(3.30)

42

ηmag =

√∂η

∂x

2

+∂η

∂y

2

+∂η

∂z

2

(3.31)

ζmag =

√∂ζ

∂x

2

+∂ζ

∂y

2

+∂ζ

∂z

2

(3.32)

The speed of sound is given by,

c =√γp/ρ (3.33)

The spectral radius, λ in ξ, η and ζ directions are calculated as,

λξ = |contraU |+ cξmag (3.34)

λη = |contraV |+ cηmag (3.35)

λζ = |contraW |+ cζmag (3.36)

Now the inviscid time step is given by,

dtI =CFL

λξ + λη + λζ, (3.37)

where CFL is the Courant-Friedrichs-Lewy number, given by the ratio of distance

traversed by the information in one time step, ∆t and the cell size, ∆x (for waves

moving in x direction). The viscous time step (dtV ) is based on the local Reynolds

number (Re), CFL number and Mach number M .

dtV =ρReCFL

Mµ(|ξ|+ |ζ|+ |η|) (3.38)

where |ξ| =√

∂ξ∂x

2+ ∂ξ

∂y

2+ ∂ξ

∂z

2, |ζ| and |η| are calculated in a similar manner. In

general, the viscous flow time step is given by,

dt = min(dtI , dtV ) (3.39)

43

3.2 Multi-Block Topology

In the multi-block approach used here the grid boundaries at the division line of

the blocks are considered as independent boundaries and appropriate boundary

conditions are incorporated at these boundaries. The details of these boundary

conditions are given in chapter 4. The grid indexing at such block boundaries

may not be consistent and will cause numerical errors if not treated properly. The

details of the indexing and the numerics of dealing with multi-block grids are given

here.

Figure 3.1 shows the way multiple blocks are created and how the grid indexing

could be different at a block interface. As shown in the figure, the block interface

could correspond to and ‘i′ increasing index in one block but a ‘j′ decreasing index

in the neighboring block.

Fig. 3.1 The centerline singularity for circular grid and a multi-block ‘H’ type grid

generation to avoid centerline singularity.

The boundary condition file generated by Gridgen stores the initial and final in-

dices of the boundary surfaces. The indices are same for the surface at which

the boundary condition is applied. A matrix of the correspondence between the

44

indices is generated using the information from the boundary condition file. This

matrix is formed of elements ii, ij etc where the value 1 of an element mean a

correspondence between those two indices at a block interface and a value 0 means

they are not aligned. A negative value means that the direction of increment is

opposite for the two indices. The matrix is given by,ii ij ik

ji jj jk

ki kj kk

(3.40)

For the example shown in figure, increasing ‘i’ correspond to decreasing ‘j’, so the

value of ij is −1 in this case, other values will depend on the index correspondence

in other two directions. When one block indices are running in the i, j and k

directions, the second block indices ineighbor, jneighbor and kneighbor will run as,

ineighbor = ineighbor + ii(i) + ij(j) + ik(k)

jneighbor = jneighbor + ji(i) + jj(j) + jk(k)

kneighbor = kneighbor + ki(i) + kj(j) + kk(k)

(3.41)

3.3 Artificial Dissipation

Artificial dissipation is implemented to increase the stability of the solution. How-

ever, too much artificial dissipation results in a smoothing of the solution and

causes the loss of important information about the flow physics. The coefficients

of artificial dissipation are chosen such that the stability is increased but the infor-

mation about flow physics is not lost. The artificial dissipation is a combination

of second and fourth order terms. The second order term is switched on when

there is a shock in the solution and fourth order terms are used in other regions

of the solution domain. The ‘switch’ is defined by the ratio of pressure terms as

given by (3.42). The values of the coefficients Cx, Cy and Cz are 0.3.The artificial

dissipation for the adjoint equations follows the same technique. The values if

45

the coefficients Cx, Cy and Cz for adjoint case are found to be 0.6, 0.3 and 0.3

respectively.

Sti

= Cx

∣∣∣pti+1− 2pt

i+ pt

i−1

∣∣∣pti+1

+ 2pti+ pt

i−1

(uti+1− 2ut

i+ ut

i−1)

Stj

= Cy

∣∣∣ptj+1− 2pt

j+ pt

j−1

∣∣∣ptj+1

+ 2ptj

+ ptj−1

(utj+1− 2ut

j+ ut

i−1)

Stk

= Cz

∣∣∣ptk+1− 2pt

k+ pt

k−1

∣∣∣ptk+1

+ 2ptk

+ ptk−1

(utk+1− 2ut

k+ ut

k−1)

(3.42)

At any intermediate stage ‘∗’, the expression for artificial dissipation is given by,

S∗i

= Cx

∣∣∣p∗i+1− 2p∗

i+ p∗

i−1

∣∣∣p∗i+1

+ 2p∗i

+ p∗i−1

(u∗i+1− 2u∗

i+ u∗

i−1)

S∗j

= Cy

∣∣∣p∗j+1− 2p∗

j+ p∗

j−1

∣∣∣p∗j+1

+ 2p∗j

+ p∗j−1

(u∗j+1− 2u∗

j+ u∗

i−1)

S∗k

= Cz

∣∣∣p∗k+1− 2p∗

k+ p∗

k−1

∣∣∣p∗k+1

+ 2p∗k

+ p∗k−1

(u∗k+1− 2u∗

k+ ut

k−1)

(3.43)

In the absence of a clear physical meaning of the adjoint variables, it is hard to

determine how the switch for the dissipation should work. Several experiments

with the artificial dissipation were conducted for adjoint equation stabilization.

The most stable solution is found when the switch for turning it on is based on the

adjoint density. The dissipation equations for adjoint solution are slightly different

than the direct flow equations. The ‘switch’ is dependent on the value of first

adjoint variable q1. These equations in i, j and k directions, for adjoint variable

46

q, are given by,

Sadjt

i= Cx

∣∣∣q1ti+1− 2q1

ti+ q1

ti−1

∣∣∣q1ti+1

+ 2q1ti+ q1

ti−1

(qti+1− 2qt

i+ qt

i−1)

Sadjt

j= Cy

∣∣∣q1tj+1− 2q1

tj

+ q1tj−1

∣∣∣q1tj+1

+ 2q1tj

+ q1tj−1

(qtj+1− 2qt

j+ qt

i−1)

Sadjt

k= Cz

∣∣∣q1tk+1− 2q1

tk

+ q1tk−1

∣∣∣q1tk+1

+ 2q1tk

+ q1tk−1

(qtk+1− 2qt

k+ qt

k−1)

(3.44)

At any intermediate stage ‘∗’, the expression for artificial dissipation is given by,

Sadj∗i

= Cx

∣∣∣q1∗i+1− 2q1

∗i

+ q1∗i−1

∣∣∣q1∗i+1

+ 2q1∗i

+ q1∗i−1

(q∗i+1− 2q∗

i+ q∗

i−1)

Sadj∗j

= Cy

∣∣∣q1∗j+1− 2q1

∗j

+ q1∗j−1

∣∣∣q1∗j+1

+ 2q1∗j

+ q1∗j−1

(q∗j+1− 2q∗

j+ q∗

i−1)

Sadj∗k

= Cz

∣∣∣q1∗k+1− 2q1

∗k

+ q1∗k−1

∣∣∣q1∗k+1

+ 2q1∗k

+ q1∗k−1

(q∗k+1− 2q∗

k+ q∗

k−1)

(3.45)

3.4 Summary

The numerical solutions to the direct and adjoint equations are found using several

different numerical schemes. The details of the numerical methods implemented to

achieve a stable solution are described in this chapter. Several numerical schemes

have been implemented to obtain the numerical solution of adjoint and direct prob-

lems. Local time stepping is used to increase the rate of convergence. It is shown

that a special treatment is required at the grid singularities. The grid singularities

are treated by using a characteristics based approach.

47

The next chapter describes the development of characteristics boundary conditions

for the direct and adjoint equations. The grids as shown in this chapter are multi-

block structured grids and there exist grid singularities at the block interfaces.

These grid singularities introduce numerical errors to the solution. The details of

the grid singularities and the boundary conditions to treat the interface boundaries

are given in the next chapter. Examples of flow solutions with the grids that have

grid singularities are given when treated with the proposed boundary conditions.

48

Chapter 4Direct and Adjoint CharacteristicsAnalysis

This chapter presents a methodology to deal with complex geometries when using

an adjoint-based optimization technique. In a multi-block approach complex ge-

ometries are meshed in a fashion such that the whole domain is divided in multiple

blocks and each simple block is considered independently. An issue at the block

boundaries arises when the communication of the information at the interfaces of

these blocks is considered. A method is provided here to ensure correct communi-

cation based on the direction of the propagation of information. The direction of

propagation of information and the required treatment at these interfaces is deter-

mined based on the characteristics of the flow and adjoint field. Adjoint boundary

conditions for the treatment of multi-block grids for complex geometries are devel-

oped. Two examples are given where there are grid singularities in the domain and

characteristics-based boundary conditions have been used to rectify the numerical

errors that can arise due to these grid singularities.

4.1 Grid singularities

Grid singularities are defined as singularities that arise due to mesh generation

for complex geometries. Wherever an abrupt change in the slope of the geometry

takes place, the grid metrics are not defined at such points because the right hand

and left hand limits of the grid metrics at such points do not converge to the same

49

value. These points of discontinuity are known as grid singularities. Such singular-

ities can cause numerical errors in the solution. The traditional method of dealing

with such singularities is to take an average of the right and left hand limits of

the grid metrics at these points. This treatment can act like a standing shock and

result in numerical errors (spurious oscillations) in the solution. Figure 4.1 shows

an example of a geometry where such grid singularities arise. The grid is divided

into two blocks at the line of discontinuity. These blocks are considered indepen-

dently with the treatment of the block interface as a boundary with appropriate

boundary conditions. A method for the correction of the residual or derivatives

at the block interface boundary is described in this section. Another example of

a grid singularity is shown in figure 4.2. Due to the circular cross-section of the

nozzle, there exists a centerline singularity in a polar grid. An ‘H’ shaped grid

can be used to avoid this centerline singularity. The center block is Cartesian and

is surrounded by polar blocks. The grid metrics are not continuous at the block

interfaces, hence they add grid singularities. These block interfaces are treated by

using a method of characteristics boundary conditions. The direction of the prop-

agation of information from one block to another is determined by the sign of the

characteristic speeds. If the information propagates out of a block, the derivatives

remain intact in that block but if the information enters a block, the derivatives are

replaced by those of the adjacent block. At points where multiple blocks join, an

average of the values of derivatives from all the adjacent blocks is then considered

to avoid any round-off errors.

The flow and adjoint equations are solved on the same mesh, hence the same issues

arise when solving either the direct or adjoint equations. Adjoint characteristics-

based interface boundary conditions are derived using the same approach as for

the direct flow equations. A condition for residual correction at the interfaces is

developed based on the idea that time derivatives are the same at such interfaces

because both boundaries (in both blocks) represent essentially the same physical

50

location. The residuals of the adjoint variables are calculated independently using

one sided differences in each block. These residuals are then corrected using the

residuals in the neighboring block based on the direction of the propagation of

information. Adjoint characteristics travel in the opposite direction to the direct

flow characteristics. So, where the direct flow derivative is corrected using the

neighboring block derivative, the adjoint residuals are retained at such interfaces.

When the direct flow derivatives are retained in a block, the adjoint derivatives are

corrected using the information from the neighboring block. Characteristics-based

boundary conditions allow the information to propagate smoothly at the interfaces

between blocks.

Fig. 4.1 The generation of multi-blocks showing grid singularities at the block interface.

51

Fig. 4.2 Multi-block ‘H’ type grid generation to avoid the centerline singularity.

4.2 Characteristics and Interface Boundary Con-

ditions for the Direct Flow Equations

The governing equations in conservative form are given by,

∂U

∂t+∂Fi∂xi

= 0 (4.1)

Where U is the vector of conservative variables and Fi are the fluxes in the ith

direction. This equation can be rewritten as,

∂U

∂t+∂Fi∂U

∂U

∂xi= 0 (4.2)

or,

∂U

∂t+ Ai

∂U

∂xi= 0 (4.3)

Diagonalization of Ai transforms this equation into a traveling wave form for any

specific direction. If the x1 direction is considered, only the terms in that direction

52

are retained and other terms are combined together to form a source term C, as

follows,

∂U

∂t+ A1

∂U

∂x1

+ C = 0 (4.4)

Diagonalizing A1(= SΛS−1) using the eigenvectors of A1 gives a way to identify

the direction of the propagation of information.

∂U

∂t+ SΛS−1 ∂U

∂x1

+ C = 0 (4.5)

S−1∂U

∂t+ L+ S−1C = 0 (4.6)

Where, L in the x1 direction is given byL = ΛS−1 ∂U∂x1

. Together, the eigenmatrix

(S) and the propagation terms constitute L. These terms are then either retained

or corrected using the terms from the neighboring block, based on the direction of

propagation. The expression for L is given by,

L = −S−1∂U

∂t− S−1C (4.7)

The value of L in each neighboring block is calculated using equation (4.7). Here,

the time derivative term is calculated using one-sided differences in that block.

Once the values of L are known for two neighboring blocks, they are retained in

the block from which the characteristics are traveling, and corrected in the block

into which the characteristics are pointing. The new value of L is then used to

calculate the correction to the residuals at the interface.

4.3 Adjoint Characteristics and Interface Bound-

ary Conditions

The implementation of these boundary conditions is based on the approach de-

scribed by Kim [20], Du [9], Hayashi et al.[12]. The idea is to write the adjoint

53

equations in characteristic form to identify the terms that affect the communica-

tion of the information. The characteristics are either replaced or retained in one

block based on the sign of the eigenvalues of the propagating matrix. That is, if

characteristics are pointing or outwards respectively in that block. The adjoint

equations can be written in compact form,

∂Q

∂t+∂Fi∂U

T ∂Q

∂xi= 0 (4.8)

The coefficient matrix ATi

= (∂Fi/∂U)T is the transpose of the coefficient matrix

for the direct equations. The right and left eigenvectors of Ai are written as l and

r, and satisfy,

lTAi = λlT (4.9)

and,

Air = λr (4.10)

To obtain the eigenvectors of ATi

, the transpose of equations (4.9) and (4.10) is

considered,

ATil = λl (4.11)

and,

rTATi

= λrT (4.12)

Therefore, the right eigenvectors for the direct equations become the left eigen-

vectors for the adjoint equations, and the left eigenvectors for the direct equations

become the right eigenvectors for the adjoint equations. Thus when a characteristic

wave leaves a domain for the direct flow, a corresponding characteristic wave enters

the domain for the adjoint flow. This property is used to identify those derivatives

that are corrected using the information from the interior of a particular block and

those derivatives that are copied from the neighboring block.

At a block interface, the values of flow and adjoint variables should match for both

the blocks. This condition should be satisfied for all times. Therefore, the time

54

derivatives of the flow and adjoint variables should match at the interfaces. That

is,∂Q

∂t

L

=∂Q

∂t

R

(4.13)

For a given direction, x1 for example, equation (4.8) can be written as,

∂Q

∂t+∂F1

∂U

T ∂Q

∂x1

+ C = 0 (4.14)

Multiplication by ST , where A = SΛS−1 yields,

ST∂Q

∂t+ ΛST ∂Q

∂x1

+ STC = 0 (4.15)

which can be rewritten as,

ST∂Q

∂t+ L + SC = 0 (4.16)

This equation is used for matching at the block interface, given by condition (4.13).

If the characteristics are going from R to L,

LL = LR + SCR − SC

L, (4.17)

and if the characteristics are going from L to R,

LR = LL + SCL − SC

R (4.18)

The above condition is implemented to match the residuals for the block interfaces.

Additional details can be found in Kim et al. [20], Hayashi et al.[12] and in Du [9].

4.4 Results

The goal of the present approach is to provide a method to deal with the grid

singularities such that each block in the grid can be considered independently.

This can be done using the characteristic information at the interface boundaries

55

to make a correction in the residual or time derivatives at the boundaries. The

approach is based on the fact that physically, both the boundaries represent the

same surface in three dimensions or line in two dimensions and hence the time

derivatives should match at such boundaries.

The first illustration of this approach shows the propagation of an acoustic pulse

through block interfaces inside a cube. Figure 4.3 shows the grid for this case

when there are 13 blocks in the domain. The grid is generated such that there are

block interfaces with grid singularities. The inner most block has 101x31x31 grid

points, the four blocks surrounding the inner most blocks have 101x31x11 points,

surrounded by blocks that have 101x31x41 and 101x41x41grid points. There are

no block interfaces in the third dimension (x direction) and all blocks have 101

points in the axial direction.

Fig. 4.3 The mesh and multi-block topology for propagation of a Gaussian pulse in a

cube.

56

To initialize the disturbance, a Gaussian pulse is generated at the center of the

cube by setting,

p = pa + εpaexp(−ar2)

ρ = ρ0 + ερ0exp(−ar2)(4.19)

Where p and ρ are pressure and density respectively. pa is the atmospheric pressure

and ρ0 is the mean density. The amplitude of the perturbation ε is 0.01 and the half

width of the pulse is 0.1. The radial distance from the center of the pulse (xo, yo, zo)

is given by r =√

(x− xo)2 + (y − yo)2 + (z − zo)2 . The grid is generated using

Gridgen and has a total of 101x111x111 points as shown in figure 4.3. An analytic

solution is available for the propagation of a spherical Gaussian pulse and these

results are compared with the analytic solution. The analytical solution for the

pressure at any time t is given by,

p = pa +εpa2r

((r − ct)(exp−a(r−ct)2 + (r + ct)exp−a(r+ct)2

)(4.20)

The adjoint solution depends on the direct flow solution. The direct flow and

adjoint equations are solved on the same grid with block interfaces. The treatment

method for these block interfaces was described in the preceding sections of this

chapter. The direct flow equations are solved on this grid using Euler’s equations.

The solution for the direct flow is shown in figures 4.4a to 4.4d. It can be seen that

the pulse passes through the grid singularities smoothly and without generating

any spurious oscillations. There are both types of singularities in the grid shown in

figure 4.3. The central block would be circular having a centerline grid singularity if

a polar grid were used. This is avoided by the generation of the ‘H’-type grid. The

central block is connected to the neighboring rectangular blocks generating abrupt

changes in the slopes. As mentioned above, the analytic solution is available for the

propagation of Gaussian pressure pulse through a cube. An additional comparison

is made between the numerical and analytic solutions. Figures 4.5a to 4.5c show

57

the comparison at different times. It can be observed that the two solutions match

well.

(a) (b)

(c) (d)

Fig. 4.4 Propagation of a Gaussian pulse through block interfaces at a) t = 0s, b) t =

0.001s, c) t = 0.002s and d) t = 0.004s.

58

(a)

(b)

(c)

Fig. 4.5 Comparison of analytic (solid) and numerical (symbol) solutions for Gaussian

pulse propagation.

59

The second illustration describes the solution of the adjoint equations in a convergent-

divergent nozzle. Due to the potential singularity at the nozzle centerline, an ‘H’

shape grid is generated with 5 blocks in the axial direction. A cross section of the

nozzle is shown in figure 4.6. It can be observed that there are grid singularities

in the cross stream direction at the block interfaces. Figures 4.7a - 4.10b show the

steady state solution of the direct and adjoint equations. A similarity between the

adjoint and flow variables can be observed. But, the adjoint variables have oppo-

site trends than the direct flow variables. The inflow for the direct calculations

becomes the outflow for the adjoint variables and the outflow for the direct calcu-

lations becomes the inflow for the adjoint calculations. The regions of low pressure

become regions of high adjoint pressure. Similar traits can be observed in the

density and velocity behavior. These differences are also shown in the boundary

conditions. At the inflow, four direct boundary conditions are specified whereas

only one boundary condition is implemented for the adjoint variables. On the other

hand, at the outflow, one boundary condition is specified for the direct calculation

and four boundary conditions are specified for the adjoint solution. This reinforces

the notion that adjoint characteristics move in the opposite direction to the direct

characteristics, taking the information in the opposite direction. This is the reason

why adjoint variables can be used for the purpose of inverse optimization. They

bring feedback into the domain because of the characteristics moving in opposite

direction.

60

Fig. 4.6 The cross-section of the convergent-divergent nozzle with five blocks to avoid

centerline singularity.

(a) Direct pressure contours (b) Adjoint pressure contours

Fig. 4.7 The comparison of direct and adjoint pressure contours.

61

(a) Direct axial velocity contours (b) Adjoint axial velocity contours

Fig. 4.8 The comparison of direct and adjoint axial velocity contours.

(a) Direct cross-stream velocity contours (b) Adjoint cross-stream velocity contours

Fig. 4.9 The comparison of direct and adjoint cross-stream velocity contours.

(a) Direct density contours (b) Adjoint density contours

Fig. 4.10 The comparison of direct and adjoint density contours.

62

4.5 Conclusion

The adjoint method is a very efficient and computationally economic method to

achieve optimization when there are many design parameters in consideration. A

difficulty arises at the grid singularities when complex geometries are considered

and a multi-block grid topology is used. A method to treat the grid singularities for

complex multi-block grids has been described for the direct and adjoint equations.

The eigenvalue analysis of the adjoint equations reveals that adjoint characteristics

travel in the opposite direction to the physical characteristics. It has been shown

that this property can be used for the correction of residuals at the grid interfaces

according to the direction propagation of information.

4.6 Summary

A method to treat grid singularities for direct and adjoint calculations has been

discussed in this chapter. Grid singularities arise when there is an abrupt change

of the grid metrics. These singularities are treated by dividing the grid into multi-

ple blocks at the line of discontinuity. The line of discontinuity is considered as a

separate boundary and boundary conditions at this boundary are developed using

the characteristic form of the equations. It is shown that the method is effective

for all kinds of grid singularities with an example of the propagation of a Gaussian

pulse through the domain. The behavior of the adjoint and direct flow variables

is compared and it is shown that adjoint variables have characteristics opposite to

the direct flow variables.

The aim of this thesis to develop a method for the optimization of blowing at the

wall of a convergent-divergent nozzle to achieve a desired pressure distribution. It

will be shown that it is achievable with the use of two different approaches. The

first approach parameterizes the blowing velocity at a fixed region on the nozzle

wall and the optimum value of the parameters is found using the adjoint method.

63

The details of this approach are given in chapter 5. The second approach considers

the blowing velocity components as free design parameters, so the three blowing

velocity components at each grid point in the blowing region are considered as

independent design parameters. The sensitivity of the cost function with respect

to the blowing velocity components is then calculated using adjoint solution. The

details of this approach are given in chapter 6. Examples for both the approaches

are included in the respective chapters.

64

Chapter 5Parameterization of the Control

The general theory of the adjoint method was described in chapter 2. For a given

set of control parameters and constraints, an adjoint method is used to find the

value of the control parameters to minimize a cost or objective function. The goal of

this thesis is to demonstrate that the adjoint method can be used to find a blowing

distribution on a fixed region in the nozzle wall such that the pressure distribution

on the nozzle wall is close to a target pressure distribution. The blowing velocity

is the control parameter that is an unknown. There are several examples in the

literature of shape optimization using adjoint methods but there have been only

few uses of adjoint methods for the optimization of physical parameters such as

blowing velocity. Two approaches have been considered in this thesis. The first

approach is such that the blowing velocity is considered as a combination of basis

functions. This chapter describes the development of the adjoint equations with

the use of basis functions to determine the blowing velocity distribution. Two

examples, one with only one parameter to describe the blowing velocity and the

other with three parameters are given. The approach can be generalized for n

parameters. The blowing distribution g(x) is given by,

g(x) =N∑n=1

anfn(x) ;x ∈ Γa (5.1)

65

where an are the design parameters or control parameters, and fn(x) are the basis

functions. The basis functions are chosen such that they satisfy the constraints to

be followed by blowing velocity g(x). Γa is the region known as the ‘actuator’ where

blowing is to be implemented and N is the total number of design parameters. In

this chapter, it is shown that an adjoint method can be used to find the set of design

parameters an, such that the pressure distribution on the wall of a convergent-

divergent nozzle is close to a target distribution.

5.1 The Mathematical Development

The adjoint formulation used here is based on the boundary control theory de-

veloped by Collis [6] and Joslin [16]. Figure 5.1 shows a schematic of the nozzle

contour and the actuator and sensor regions. The sensor region Ωs is the region

where the pressure is matched (or in other words, the cost function is measured).

The actuator region Γa is the region where the blowing is implemented. Ω denotes

the full domain of computation, only half of the nozzle is considered because of

the symmetry about the nozzle centerline.

Fig. 5.1 Nozzle contour, sensor and actuator regions.

66

The present formulation is based on the invisicid Euler equations, the same ap-

proach can be used in the development for the Navier-Stokes equations and turbu-

lence models. The inviscid compressible Euler equations in conservative form are

given as,

∂U

∂t+∂F

∂x+∂G

∂y= 0 (5.2)

where,

U =

ρ

ρu

ρv

ρE

, F =

ρu

ρu2 + p

ρuv

ρuH

, G =

ρv

ρuv

ρv2 + p

ρvH

(5.3)

the system is closed with,

p = (γ − 1)ρ

(E − u2 + v2

2

)and H = E +

p

ρ(5.4)

The steady state problem is considered here and the time derivative term goes to

zero. But the time derivative term in the governing equations is retained since the

equations are marched in time to obtain the steady state solution. The adjoint

equations thus developed also include the time derivative term and artificial time

marching is used to obtain the steady state solution of adjoint equations. More

details on the steady state adjoint solution can be found in section 2.4. The

actuator region is considered as an inflow region where the blowing boundary

condition is implemented as an inflow boundary condition. More details of this

approach can be found in Collis et al. [6]. The number of boundary conditions

specified at a boundary is same as the number of incoming characteristics. In

the case of subsonic blowing, three characteristics travel into the domain whereas

one travels out of the domain. Hence three physical and one numerical boundary

condition have to be specified at the boundary. One boundary condition is that

67

the normal velocity, vTn at the lower wall is equal to the blowing velocity,

vTn = g on Γa, (5.5)

where g is the blowing velocity, v is the velocity vector and n is the normal vec-

tor. The blowing takes place at part of lower wall boundary called the actuator

boundary Γa . If the injection is considered to be isentropic and irrotational, two

more boundary conditions can be obtained in the actuator region. The boundary

condition vector is given by,

B(u,∆u, g) =

vTn− g

(S − So)min0, g2(∂u∂y− ∂v

∂x

)min0, g2

= 0 on Γa, (5.6)

Where S is the entropy and So is considered to be zero for isentropic blowing.

The third term ∂u∂y− ∂v

∂xdenotes the vorticity and is zero for irrotational blowing.

g is negative in the case of suction . Only one characteristic travels out of the

domain in the case of subsonic suction and only one physical boundary condition

is required at the wall. In such a case, the vector of boundary conditions reduces

to only one boundary condition vTn− g = 0 and other two terms in the vector B

go identically to zero.

The cost function is the quantity that is to be minimized using the adjoint method.

The cost function is formulated as a measure of the difference between the pressure

and a desired pressure, as given in equation (5.7). Here, pd is the desired pressure

and p is the pressure at any design cycle. The desired pressure pd could be the

ambient pressure or the pressure distribution corresponding to an ideally expanded

flow. Ωs is the part of the domain where the pressure is to be matched and is known

as sensor. The cost function is given by,

68

J =1

2

∫Ωs

(p− pd)2 dΩ (5.7)

It is possible to apply additional constraints on the control by adding extra terms in

the cost function. Such constraints could be to make sure that the control remains

bounded and that there are no numerical fluctuations. The following constraints

are applied on the control g,

1. The magnitude of g should be a minimum to save power input.

2. g should vary smoothly.

that is,

g =dg

dx= 0 on ∂Γa (5.8)

where ∂Γa is the boundary of the region Γa. These constraints are applied by

adding penalty terms to the cost function. The modified cost function is given by,

J =α

2

∫Ωs

(p− pd)2 dΩ +β

2

∫Γa

(g2 +

dg

dx

2)dΓ (5.9)

The first additional penalty term makes sure that minimizing the cost function J

will minimize the magnitude of the control g and the second penalty term makes

sure that the control g should go smoothly to zero at the boundary ∂Γa of control

actuator g. α and β are the weighting parameters to control the relative contribu-

tion of two terms.

A Lagrange functional using the governing equations, control equations and the

modified cost function is formulated as,

L =α

2

∫Ωs

(p− pd)2 dΩ +β

2

∫Γa

(g2 +

dg

dx

2)dΓ +∫

Ω

λT[∂F

∂x+∂G

∂y

]dΩ +

∫Γa

λbTB(u,∆u, g)dΓ

(5.10)

69

Here, λ and λb are the adjoint variables or Lagrange multipliers corresponding to

the flow equations and the boundary conditions respectively.

The optimization of the cost function with respect to the constraints is the same

as making the Lagrangian functional stationary with respect to the variables. The

adjoint equations are obtained by setting ∂L∂uu′ = 0 for all u′,where u′ is the change

in the flow variables with respect to the change in the design parameters, and the

gradient of the cost function with respect to the control is calculated by,

∂J

∂gg′ =

∂L∂gg′ (5.11)

Here g is the control (blowing velocity), J is the cost function and L is the La-

grange functional. A prime on a quantity describes a fluctuation in that quantity.

The above holds true because the constraint equations are followed by the flow

variables and therefore the third and fourth terms are zero in equation (5.10).

The vectors of adjoint variables λ and boundary conditions λb are given by,

λ =λ1 λ2 λ3 λ4

T(5.12)

λb =λb1 λb2 λb3

T(5.13)

The gradient with respect to the design variables g is given by,

∂L∂gg′ = β

∫Γa

(gg′ +

∂g

∂x

∂g′

∂x

)∂Γ

+

∫Γa

λTbBg(u,∆u, g)g′ ∂Γ

(5.14)

Using the divergence theorem, this can be rearranged as,

70

∂L∂gg′ = β

∫Γa

(g − ∂2g

∂x2

)g′ ∂Γ

+∂g

∂xg′∣∣∣∣∂Γa

+

∫Γa

−λb1g′ ∂Γ

(5.15)

Which can be further simplified as,

∂L∂gg′ =

∫Γa

(g − ∂2g

∂x2

)− λb1

)g′ ∂Γ (5.16)

Since ∂g∂x

is required to be zero on the actuator boundary, ∂Γa. The equation (5.16)

is used to calculate the value of the control parameter g. The adjoint equations

are derived by making the Lagrange functional stationary with respect to change

in the flow variables u′,

∂L

∂uu′ = α

∫Ωs

(p− pd)∂p

∂uu′ ∂Ω

+

∫Ω

λT[∂

∂x

∂F

∂uu′ +

∂y

∂G

∂uu′]∂Ω

+

∫Γa

λTb

[Bu(u,∆u, g)u′ +Buxj

(u,∆u, g)u′xj

]∂Γ

(5.17)

A simplification is made using integration by parts and use of the divergence the-

orem,

∂L

∂uu′ =

∫ t1

t0

∫Ω

rTu′ ∂Ω −∫

Ω

[∂λT

∂xA1 +

∂λT

∂yA2

]u′ ∂Ω

+

∫Γ

[(λT∂F

∂un1

)+

(λT∂G

∂un2

)]u′ ∂Γ

+

∫Γa

λTb

[Bu(u,∆u, g)u′ + Buxj

(u,∆u, g)u′xj

]∂Γ

(5.18)

71

where A1 = ∂F∂u

, A2 = ∂G∂u

and r = ∂p∂U

TKΩs. KΩs

is 1 when in Ωs and zero otherwise.

The adjoint equations as well as the boundary conditions can be summarized as,

AT1

∂λ

∂x+ AT

2

∂λ

∂y= r in Ω (5.19)

The adjoint boundary conditions are given by,

∑njB

T

uxjλb = 0 (5.20)

n1AT

1λ+ n2A

T

2λ+BT

uλb − ∂s

∑sjB

T

uxjλb = 0 (5.21)

where nj is the unit outward normal to the boundary and sj is the tangential

vector on the boundary.

5.2 Optimization with Parameterization of Blow-

ing

The goal of this chapter is to demonstrate that the above approach can be used

to find a blowing distribution on the nozzle wall such that the pressure in a fixed

region matches well with the target pressure. The blowing is dependent on a

number of design parameters. The formulation and implementation of this method

is complex and nontrivial. To simplify the problem, the blowing is incorporated on

a limited section of the nozzle wall and two-dimensional calculations are performed

for the interior of nozzle. It is shown that moving the shock towards the exit of

the nozzle is possible with the use of blowing. Two examples are given. The first

example has a Gaussian blowing distribution and the amplitude of this Gaussian

is considered as the control parameter. The optimum value of the amplitude is

found using the adjoint approach. The second example considers the blowing

distribution as a combination of three trigonometric functions. The coefficients of

72

these three functions are considered as control parameters. The optimum values of

these parameters are found using the adjoint approach as described in this chapter.

As a simple example of the design process a contoured nozzle with a simple geom-

etry is considered. The nozzle contour is given by,

y = 1.75− 3αcos[(0.2x− 1)π] for 0 ≤ x ≤ 5

y = 1.25− 3αcos[(0.2x− 1)π] for 5 ≤ x ≤ 10(5.22)

where the value of α is 0.25. A Gaussian distribution with the amplitude vb

is considered for the blowing velocity at the wall. The blowing distribution is

described as,

g(x) = vb exp

− x− (x1+x2)2

2

[(x− x1)2(x− x2)2]

for x1 ≤ x ≤ x2 (5.23)

Note that this distribution follows the condition (5.8) and the value of blowing

velocity g goes to zero at the actuator boundaries x1 and x2. A positive g means

that fluid is injected into the flow whereas a negative g means suction out of the

domain. The nozzle pressure ratio is kept constant so that the operating conditions

remain the same for all the cases. The value of nozzle pressure ratio is 1.5 for the

results presented here. The nozzle contour geometry has been kept fixed and is

given by equation (5.22). The goal is to find an appropriate blowing rate that will

remove the shock from the nozzle interior without changing the nozzle geometry

or operating conditions. The cost function is given by equation (5.7), where the

desired pressure is found for a known value of the amplitude vb. The desired

pressure distribution is first found for vb = 60 m/s. The cost function is measured

inside a fixed region of the nozzle as shown in figure 5.1. Figure 5.2 shows the

changes in the flow as the amplitude of blowing velocity is varied. The shock

location and pressure contours change as the value of vb is varied. The shock

moves towards the nozzle exit as the rate of blowing is increased. Since only the

interior of the nozzle has been considered for this calculation, complete removal of

73

the shock is not possible but the shock is pushed to the nozzle exit. The pressure

corresponding to vb = 60.0 m/s is taken as the desired pressure. The goal is to

reach to the desired value of pressure using an adjoint optimization approach. The

change in blowing velocity amplitude with design cycles is shown in figure 5.3. The

value of the amplitude changes from an initial value of zero to a value of 59.032 m/s

in 5 design cycles. The value of vb changes more rapidly in the initial design cycles

and changes at a slower rate as it approaches the desired value. It then begins to

fluctuate around this value. The decay in the cost function is shown in figure 5.4,

the cost function decays almost linearly whereas the blowing amplitude changes in

a non-linear fashion. As the pressure distribution gets close to the desired pressure,

the amplitude of blowing vb also moves towards the desired blowing. It will be seen

in the next chapter that when there is more than one control parameter, in some

cases the adjoint approach finds a new value of the control parameters, but the

pressure distribution thus found is always close to the target distribution.

74

Fig. 5.2 The variation in pressure contours and streamlines, in the divergent section of

the nozzle, as the amplitude of the blowing velocity vb is varied to 0, 10, 30 and 60 m/s.

Fig. 5.3 Variation of the amplitude of the blowing vb in m/s with design cycles.

75

Fig. 5.4 Decay of the cost function in Pa2 with design cycle.

The first example, as shown above, demonstrates that the adjoint method can be

used to find the blowing amplitude on the nozzle wall. The method can be used to

find a pressure distribution corresponding to a known blowing distribution, where

the blowing distribution is dependent only on one parameter. The second example

considered is more general example than this. A blowing distribution based on

several control parameters is considered and the optimum solution is not known.

An adjoint method is used to find that blowing distribution that removes the

shock from inside the nozzle. The blowing distribution has been considered as a

combination of trigonometric basis functions that follow the condition (5.8). The

sensitivity of the cost function with respect to these control parameters can be

found using the adjoint method. The generalized blowing distribution g is given

by,

g(x) =1

2

N∑n=1

an1− cos[(2n− 1)3π(x− x1)/2(x2 − x1)] x ∈ [x1, x2] (5.24)

76

g′(x) =1

2

N∑n=1

an(2n− 1)3π

2(x2 − x1)sin[(2n− 1)3π(x− x1)/2(x2 − x1)] (5.25)

Here an are the control parameters. The total number of design parameters, N

is taken to be 3 for simplicity but a larger value of N could be considered with-

out much difference in the computational cost. In this case the cost function is

calculated in a region near the lower wall and the desired pressure is taken to be

the ambient pressure. The goal is to achieve a pressure distribution in the sensor

region, that is as close to the ambient pressure as possible, using blowing at the

lower wall. The cost function is formulated as equation (5.7), where pd is consid-

ered to be ambient pressure pa.

The nozzle pressure ratio has been kept constant as before and the geometry is

fixed to the initial geometry as shown in figure 5.22. There is no blowing for the

first design cycle. The region of observation or sensor can be chosen as any region

in the nozzle interior domain or on the boundaries. This region is shown by the box

in figure 5.5. This region has been chosen based on the initial flow field such that

it captures the shock. The sensor region extends in the axial direction between

x/D = 8.0 to x/D = 9.1 and in the transverse direction it goes from y/D = 0.5 to

y/D = 1.0, whereD is the nozzle throat diameter. The sensor region approximately

spans 8.3% area of the total nozzle domain. The shock removal inside the nozzle

was achieved by using the blowing on the wall with a blowing distribution given by

equation (5.24). There is no blowing in the first design cycle. Figure 5.5 shows the

pressure contours after different numbers of design cycles. It takes 4 design cycles

to obtain the blowing distribution that shows no shock inside the nozzle. The

normalized cost function dropped from a value of 5.06 to 2.03. Then it increases

and then keeps fluctuating about the minimum value. The fluctuations in the cost

function are significant. But, this can be circumvented by the multiplication of a

relaxation coefficient in the updating of the control parameters.

77

Fig. 5.5 Removal of shock with adjoint design cycles. The sensor region is shown as a

box in the diverging section of the nozzle.

5.3 Conclusion

Blowing at the nozzle wall can be used to change the pressure distribution inside

the nozzle. It is possible to use an adjoint approach to find the blowing velocity

distribution such that the pressure inside the nozzle matches well with the desired

pressure. The blowing velocity can be written as a combination of multiple pa-

rameters and/or basis functions. The basis functions are chosen such that they

follow the constraints required of the blowing velocity. The coefficients of the basis

functions are considered as control parameters. An adjoint approach is developed

such that the blowing region is considered as an inflow boundary and the gradients

of the cost function with respect to the coefficients of the basis functions are found

78

using the adjoint method. These gradients are then used to calculate the new

values of the coefficients. In the first example, where the blowing velocity distribu-

tion is Gaussian in shape, the amplitude of the blowing velocity is achieved within

an error of 1.6% in 5 design cycles. In the second example, where the blowing is

described in terms of three control parameters, it is shown that moving the shock

to the nozzle exit is possible by using this approach. In the case where an external

domain is included in the computation, this approach can be used to move the

shock out of the nozzle which can help in reducing the shock associated noise.

5.4 Summary

This chapter describes the development and implementation of an adjoint method

to optimize the blowing in a fixed region on a nozzle wall. The blowing has been

considered to be a sum of the products of design parameters and basis functions.

It is shown that the adjoint method can be used to minimize the cost function

that is defined in a region inside the nozzle. For most practical cases it is not

necessarily possible to achieve a blowing distribution thus described. It also lim-

its the solution space to those blowing distributions that can be written in the

given form. These issues are addressed in the next chapter where a free form of

the blowing distribution is considered. The blowing velocity at each grid point

inside the actuator region is considered to be an independent design parameter.

All components of velocity are considered as independent variables. The blowing

distribution thus found can be used to determine the angle of injection because

all three components of the velocity are considered separately. A two-dimensional

method is first developed and is later extended to three-dimensions. More details

on how the desired pressure is calculated and the blowing distribution is found

using the adjoint method are given in the next chapter.

79

Chapter 6Adjoint Control of Nozzle Flow withSurface Blowing

6.1 Introduction

This chapter describes how the adjoint method can be used to find the values of

control parameters that describe the blowing distribution on the nozzle wall to

minimize a given cost function. In this chapter a free form of blowing has been

considered as the control. It is not necessarily possible to physically achieve a

blowing distribution that is given by a set of basis functions. So, in this chapter,

all velocity components at each grid point in the actuator region are considered

as independent parameters. The relative values of the components of the blowing

velocity can be used to determine the angle of injection at the nozzle wall. This

chapter provides the tools to optimize the blowing on the nozzle wall.

The mathematical development of the adjoint equations when the control param-

eters are the blowing velocity components, differs from that given in the previous

chapter, when the control parameters are the coefficients of given basis functions of

the normal blowing velocity. A brief introduction of the mathematical development

is given first in this chapter. Because of the complexity of the adjoint equations,

a two-dimensional case is first considered. Two examples for the two dimensional

case are given. The first example demonstrates that the adjoint method can be

80

used to find an arbitrary blowing distribution such that the pressure at the nozzle

wall matches with the pressure corresponding to a known blowing distribution.

The second example shows how to achieve ideally-expanded flow from a nozzle

that is operating at off-design conditions. Then three-dimensional calculations are

performed for both the nozzle internal and external flows. First, internal flow re-

sults for the blowing optimization on the nozzle wall are given when the pressure is

to be matched with a desired pressure distribution on the nozzle wall. The second

example also includes the nozzle external flow and the optimization technique for

blowing when the external flow is included.

6.2 The Mathematical Development with Wall

Blowing Control

This section describes the development of the adjoint equations and boundary

conditions and the determination of the sensitivity of the cost function to the

design parameters. The present analysis and computations are based on the Euler

equations. The same general approach can be used for the Navier-Stokes equations

and for turbulence model equations. The reason for choosing the inviscid model

is the simplicity of development and implementation that is offers. This would

result in a loss of flow physics in the boundary layer region but this is neglected

for the present analysis since the purpose of this analysis is to demonstrate the

methodology of adjoint and not to capture the fine details of flow behavior. A

more sophisticated analysis provides the scope of the future work.

The three-dimensional, compressible, inviscid governing equations in conservative

form can be written as,

∂U

∂t+∂Fi∂xi

= 0, (6.1)

Here a repeated index implies a summation. The flux vector Fi is given by,

81

Fi = mi,miu1 + pδi1,miu2 + pδi2,miu3 + pδi3,miHT , (6.2)

And the vector of conserved variables, U, is given by,

U = ρ,m1,m2,m3, eT , (6.3)

where,

m1 = ρu1 , m2 = ρu2, m3 = ρu3, and e = ρEn (6.4)

The Kronecker delta δij is 1 when i = j and 0 otherwise. u1, u2 and u3 are the

velocity components in the x1, x2 and x3 directions respectively, En is the internal

energy, H is the enthalpy, ρ is the density and p is the static pressure. The system

is closed by the following equations for a calorically perfect gas,

H = (e+ p)/ρ, and p = (γ − 1)

[e− (mimi)

]. (6.5)

The boundary conditions at the nozzle inflow, outflow, far field and wall are given

by,

p = pa for subsonic outflow on Γo

Po = constant and To = constant on Γi

uN = 0 on Γw

ui = ubqi on Γa

(6.6)

where, Γo is the outflow boundary, Γi is the inflow boundary, Γw is the wall bound-

ary and Γa is the actuator boundary where blowing is applied. Po and To are

respectively the total pressure and total temperature at the nozzle inlet, uN is the

wall normal velocity, pa is the ambient pressure and ubqi is the blowing velocity at

82

the qth node in the ith direction.

Together, the governing equations can be written as,

R = 0 (6.7)

The changes in R due to changes in the flow variables U are given by,

δR =∂(δU)

∂t+∂(δFi)

∂xi, (6.8)

where,

δFi =∂Fi∂U

δU (6.9)

and,

δU =∂U

∂UδU = PδU (6.10)

where, U is the vector of primitive variables given by,

U = p, u1, u1, u3, TT , (6.11)

The cost function to be minimized depends on the difference between the pressure

at any design cycle and a desired pressure on a sensor where the sensor is the

region where the cost function is measured. The sensor could either be a section of

the nozzle wall or a region in the flow domain. In the former case the cost function

is given by,

J =α

2

∫t

∫Γs

(p− pd)2 dΓdt, (6.12)

where Γs is the sensor region.The time derivative terms have been retained in the

formulation as the steady state direct and adjoint solutions are obtained using time

marching.

83

Using the cost function and the flow equations a Lagrangian functional L can be

constructed as,

L = J +

∫t

∫Ω

QTR dΩdt, (6.13)

where Ω is the computational domain surrounded by the boundary Γ. Q is the

vector of adjoint variables and is given by,

Q = q1, q2, q3, q4, q5T . (6.14)

The perturbation of the Lagrangian functional with respect to the design param-

eters is given by,

δL = δJ +

∫t

∫Ω

QT δRdΩdt. (6.15)

δR is given by equation (6.8), so that,

δL = δJ +

∫t

∫Ω

QT

[∂(δU)

∂t+∂(δFi)

∂xi

]dΩdt (6.16)

From equation (6.10),

δL = δJ +

∫t

∫Ω

QT

∂(δU)

∂t+∂ [(∂Fi/∂U) δU]

∂xi

dΩdt (6.17)

Integration by parts and use of the divergence theorem yields,

δL = δJ +

[∫Ω

QT δU

]t2t1

+

∫t

∫Γ

QT ∂δFi∂U

niδUdΓdt

−∫t

∫Ω

(∂QT

∂t+∂QT

∂xi

∂Fi∂U

)δUdΩdt

(6.18)

Equation (6.18) has three terms that have separate meanings. The concept behind

the adjoint approach is to find the gradients of the cost function with respect to the

design variables, without having to calculate the change in the flow variables δU.

84

Hence the integrand in the last term of equation (6.18) is set to zero everywhere in

the flow domain. This results in a set of equations that are known as the “adjoint

equations”. The first and third terms on the right hand side of equation (6.18)

represent boundary terms and are discussed in detail in the next section. The

second term is the time condition term and the adjoint variables are set to zero at

the initial adjoint time, t2, that is,

Qt=t2 = 0. (6.19)

This condition is used as the initial condition for the adjoint solution.

The adjoint equations are developed such that the sensitivity of the cost function

with respect to the control parameters is independent of the sensitivity of the

flow variables with respect to the design variables. In order to achieve this, the

integrands in equation (6.18) are set to zero in the volume integrals. So, the adjoint

equations are given by,

∂Q

∂t+∂Fi∂U

T ∂Q

∂xi= 0 (6.20)

or,∂Q

∂t+ Ai

T ∂Q

∂xi= 0 (6.21)

with Ai = ∂Fi∂U

. After further simplification, the adjoint equations are found to be

given by,

∂q1

∂t= −(γ − 1)q2(

∂q2

∂x1

+∂q3

∂x2

∂q4

∂x3

) + u1ui∂q2

∂xi

+ u2ui∂q3

∂xi+ u3ui

∂q4

∂xi

−(− c2

γ − 1+ (γ − 2)q2

)ui∂q5

∂xi,

(6.22)

85

∂q2

∂t= − ∂q1

∂x1

− (3− γ)u1

∂q2

∂x1

− u2

(∂q2

∂x2

+∂q3

∂x1

)− u3

(∂q2

∂x3

+∂q4

∂x1

)+ (γ − 1)u1

(∂q3

∂x2

+∂q4

∂x3

+ ui∂q5

∂xi

)− H

∂q5

∂x1

,

(6.23)

∂q3

∂t= − ∂q1

∂x2

− (3− γ)u2

∂q3

∂x2

− u3

(∂q3

∂x3

+∂q4

∂x2

)− u1

(∂q3

∂x1

+∂q2

∂x2

)+ (γ − 1)u2

(∂q4

∂x3

+∂q2

∂x1

+ ui∂q5

∂xi

)−H ∂q5

∂x2

,

(6.24)

∂q4

∂t= − ∂q1

∂x3

− (3− γ)u3

∂q4

∂x3

− u1

(∂q4

∂x1

+∂q2

∂x3

)− u2

(∂q4

∂x2

+∂q3

∂x3

)+ (γ − 1)u3

(∂q2

∂x1

+∂q3

∂x2

+ ui∂q5

∂xi

)−H ∂q5

∂x3

,

(6.25)

and,∂q5

∂t= −(γ − 1)

(∂q2

∂x1

+∂q3

∂x2

+∂q4

∂x3

)− γui

∂q5

∂xi. (6.26)

Here, q2 = uiui/2 and c is the speed of sound, given by, c2 = (γp)/ρ. H is the

enthalpy, given by q2 + c2

γ−1. The time derivatives have been retained for the arti-

ficial time marching.

Equations (6.22)– (6.26) are the adjoint equations. The adjoint equations are

homogeneous linear equations. The boundary conditions on the adjoint variables

depend on the direct boundary conditions. The boundary conditions for the adjoint

variables are developed in the following section.

86

6.3 Boundary Conditions

The adjoint boundary conditions play a crucial role in the accuracy and stability

of the adjoint solution as in any other problem in numerical fluid mechanics. The

adjoint boundary conditions are related to the direct flow boundary conditions and

are developed in conjunction with the direct flow boundary condition.

The direct flow boundary conditions given by equation (6.6) are implemented ei-

ther in terms of the derivatives of the flow variables (Neumann conditions) or the

values of the flow variables (Dirichlet conditions). In both cases, the information

about the flow variables is obtained at the boundaries. This information can be

transformed to provide the change in the flow variables, δU at the boundaries. The

value of δU at the boundaries is then considered in the adjoint boundary terms

to provide information about the adjoint variables at the respective boundaries.

Thus, the adjoint boundary conditions are developed based on the change in direct

flow variables δU at the boundaries. The details are given in the following section.

The Simplification of the Adjoint Boundary Integrand

The boundary integral terms in the expression for δL are given by,

δL = δJ +

∫t

∫Γ

QT ∂Fi∂U

niδUdΓdt (6.27)

The cost function J is given by equation (6.12). Taking the perturbation of this

equation yields,

δJ = α

∫t

∫Γs

(p− pd)δp dΓdt, (6.28)

The integrand in the boundary terms can be written as,

87

QT ∂Fi∂U

niδU = QT ∂Fi∂U

ni∂U

∂UδU (6.29)

where U and U are the vectors of conserved and primitive flow variables respec-

tively given by,

U = ρ, ρu1, ρu2, ρu3, ρEnT ,

U = p, u1, u2, u3, TT .(6.30)

The transformation matrix of conversion between the primitive and conservative

variables, P is defined by,

P =∂U

∂U. (6.31)

Then the integrand of the boundary terms on the right hand side of the equation

(6.27) can be written as,

QT ∂Fi∂U

PniδU =(P δp+ Uiδui + T δT

). (6.32)

The integrand of the boundary term in the new expression for δL is given by,

Boundary Integrand =(α(p− pd)k + P

)δp+ Uiδui + T δT (6.33)

Here k is introduced to generalize the form of boundary conditions and is given

by,

k = 0 on Γ ∩ Γs,

k = 1 on Γs.(6.34)

The terms with an overbar represent the coefficients of the respective flow variable

perturbations. The coefficients in equation (6.33) can be simplified using the flow

equations and are given by,

88

P =uNRT

(u · q + q1 +Hq5) + qN , (6.35)

T =−pRT 2

uN (u · q + q1 +Hq5) +p

T

γ

γ − 1q5uN , (6.36)

U1 =p

RT(q1n1 + u1uNq5 +Hq5n1 + 2u1q2n1 + u2(q2n2 + q3n1) + u3(q2n3 + q4n1)) .

(6.37)

Similarly,

U2 =p

RT(q1n2 + u2uNq5 +Hq5n2 + 2u2q3n2 + u3(q3n3 + q4n2) + u1(q3n1 + q2n2)) ,

(6.38)

and,

U3 =p

RT(q1n3 + u3uNq5 +Hq5n3 + 2u3q4n3 + u1(q4n1 + q2n3) + u2(q4n2 + q3n3)) .

(6.39)

Here uN and qN are the wall normal direct and adjoint velocities given by,

uN = u1n1 + u2n2 + u3n3, (6.40)

and,

qN = q2n1 + q3n2 + q4n3. (6.41)

where n1, n2 and n3 are the components of the outward normal to the wall.

The adjoint equations are developed such that the requirement to calculate the

change in the flow variables δU is circumvented by making the coefficients of δU

vanish whenever possible. The adjoint boundary conditions are developed using

the same approach. The boundary integrals can be eliminated by setting the

89

boundary terms equal to zero when their coefficient δU is arbitrary. In the subse-

quent sections, the adjoint boundary conditions have been developed for the inflow,

outflow, slip wall, sensor and actuator boundaries.

6.3.1 Adjoint Outflow Boundary Condition

6.3.1.1 Subsonic Outflow

The direct flow boundary condition for the outflow is given by equation (6.6). At

a subsonic outflow, the static pressure is prescribed to be the ambient pressure.

So, for the direct flow problem, one physical boundary condition is applied at the

subsonic outflow and four numerical boundary conditions are applied to determine

the value of four remaining variables.

For a prescribed pressure p = pa, the change in pressure δp will be zero.

p = pa =⇒ δp = 0 (6.42)

The remaining terms, δui and δT , in equation (6.33) are non zero at the subsonic

outflow boundary. After substitution in to equation (6.33),

Boundary Integrand =(α(p− pd)k + P

)

0δp+ Ui δui︸︷︷︸

6=0

+T δT︸︷︷︸6=0

, (6.43)

So, four terms remain that need to be eliminated. That is at the outflow boundary,

U1 = U2 = U3 = T = 0. (6.44)

To simplify, it is assumed that the outflow is directed in the positive x1 direction.

That is, n1 = 1 and n2 = n3 = 0 at the outflow boundary. Using the coefficients

given by equations (6.36)- (6.39), the following boundary conditions are obtained

at a subsonic outflow,

q3 = −u2q5 (6.45)

90

q4 = −u3q5 (6.46)

q2 = − 1

u1

(u2

1+

c2

γ − 1

)q5 (6.47)

q1 =

(q2 +

c2

γ − 1

)q5 (6.48)

Equations (6.45)- (6.48) are the boundary conditions for the adjoint outflow. At

the subsonic outflow one adjoint variable is extrapolated from the interior of the

domain and the four boundary conditions given by (6.45) to (6.48) are applied to

calculate the values of the remaining four adjoint variables at the outflow.

6.3.1.2 Supersonic outflow

At a supersonic outlet, none of the flow quantities is specified for the direct flow

and all of the variables are computed from the interior of the domain. Thus, at

the outflow,

δp 6= 0, δui 6= 0, δT 6= 0. (6.49)

So the coefficients of δui, δp and δT in equation (6.33) must vanish at the outlet.

That is,

Boundary Integrand =(α(p− pd)k + P

)δp︸︷︷︸6=0

+Ui δui︸︷︷︸6=0

+T δT︸︷︷︸6=0

, (6.50)

So, at a supersonic outflow,

U1 = U2 = U3 = T = P = 0. (6.51)

Five boundary conditions are obtained from these relationships. On simplification,

the equations reduce to,

q3 = −u2q5 (6.52)

q4 = −u3q5 (6.53)

91

q2 = − 1

u1

(u2

1+

c2

γ − 1

)q5 (6.54)

q1 =

(q2 +

c2

γ − 1

)q5 (6.55)

and

q1 + q2u1 + q3u2 + q4u3 + q5

uiui2

= 0. (6.56)

Equations (6.52) - (6.56) can be used to solve for q1, q2, q3, q4, q5T at a supersonic

outflow.

6.3.2 Adjoint Inflow Boundary Condition

The boundary condition at the nozzle inlet for the direct problem is given by

equation (6.6). The total pressure and total temperature are specified at the

nozzle inlet. The flow direction at the nozzle inlet is also specified. For simplicity,

it is assumed that the nozzle inlet surface is normal to the x1 direction. That is,

n1 = −1 and n2 = n3 = 0 at the inlet. An inlet in any general direction can be

considered in the same way.

For the direct flow,

Po = constant ;To = constant, u2 = u3 = 0, (6.57)

at the nozzle inflow. Then, at the inflow,

δPo = 0 ; δTo = 0 ; δu2 = δu3 = 0 (6.58)

The isentropic flow relations are given by,

To = T

(1 +

γ − 1

2

u21

γRT

)and Po = p

(1 +

γ − 1

2

u21

γRT

) γγ−1

. (6.59)

92

Finding the change in these equations and using equations (6.58), (6.59) the rela-

tions between δT , δp and δu1 can be found. These relations are given by,

δT =γ − 1

γRu1δu1. (6.60)

Similarly,

δp = −pu1

RTδu1. (6.61)

Substituting these into the boundary integrand (6.33) yields,

Boundary Integrand =(α(p− pd)k + P

)δp︸︷︷︸

f(δu1)

+U1 δu1︸︷︷︸6=0

+U2>

0δu2 + U3

>0

δu3 + T δT︸︷︷︸f(δu1)

,

(6.62)

using equations (6.60) and (6.61), the coefficients of δT and δp can be combined to

make one coefficient of δu1. u1 is arbitrary at the inflow boundary, so the coefficient

of δu1 must vanish at the inflow. This results in one adjoint boundary condition

at the inflow, given by,

q2 = − 1

u1

[q1 +

(q2 +

c2

γ − 1

)q5

](6.63)

At the adjoint inflow, four numerical boundary conditions are applied to the ad-

joint variables q1, q3, q4, q5 and the values are extrapolated from the interior of the

domain. So, one adjoint boundary condition, given by equation (6.63), is applied

to calculate q2 at the inflow boundary.

6.3.3 Adjoint Slip Wall Boundary Condition

At a slip wall, for direct flow, there is no flow penetration (except at the actuator

boundary). The wall boundary condition for the direct flow problem is given by

zero wall normal velocity,

uN = 0. (6.64)

93

This can be written as,

δu1n1 + δu2n2 + δu3n3 = 0. (6.65)

For an adiabatic wall, there is no specification of p or T on the wall, therefore δp

and δT are arbitrary on the wall. Considering the adjoint boundary integrand :

Boundary Integrand =(α(p− pd)k + P

)δp︸︷︷︸6=0

+U1 δu1︸︷︷︸6=0

+U2 δu2︸︷︷︸6=0

+U3 δu3︸︷︷︸6=0

+T δT︸︷︷︸6=0

,

(6.66)

The coefficients of δT and δp are required to vanish on a slip wall because no

information is given about δT and δp. That is,

T = P = 0 (6.67)

The coefficients of δui are not individually equated to zero because of the known

condition given by (6.65). Using equation (6.35) and (6.65), a single adjoint

boundary condition is obtained at the slip wall,

q2n1 + q3n2 + q4n3 = qN = 0, (6.68)

where qN is the wall normal adjoint velocity. Using the wall normal adjoint and

direct velocities equal to zero, equation (6.64) and (6.68), the equation for T and

P (6.67) is satisfied automatically. Thus equation (6.68) provides the only bound-

ary condition required at a slip wall.

6.3.4 Sensor Boundary Condition

In the present formulation, the sensor is a region on the wall boundary. The wall

boundary condition must be modified to account for the sensor region. In gen-

eral, the sensor could be any part of the domain. If the sensor region is inside

the computational domain, the cost function term appears as a source term in the

94

equations. If the sensor were at any other boundary, the cost function term would

appear in the respective boundary condition.

The general expression for the wall boundary conditions is still applicable at the

part of the wall which is defined as sensor. The parameter k, as defined earlier by

equation (6.34), is non-zero at the sensor and adds an extra term to the boundary

condition. At the sensor,

P + α(p− pd) = 0 (6.69)

Simplification of this equation yields,

qN + α(p− pd) = 0 (6.70)

The wall normal adjoint velocity in the sensor region is non-zero and is given by

the difference between the pressure at any design cycle and the desired pressure.

6.3.5 Actuator Boundary Condition

The actuator is the part of boundary where control (in this case, the injection or

blowing) is implemented. The boundary condition on the actuator for the direct

flow solution is different from than that on a solid wall. So the boundary condition

for the adjoint calculation on the actuator is also different from that of the solid

wall. The control, or actuator boundary, consists of blowing velocities at different

nodes that are each considered as independent control variables. Each grid point in

the actuator region is considered as one node. In general, all three components of

the blowing velocities at these nodes are considered as separate control parameters.

So, if the number of nodes in the actuator region is q, the total number of control or

design parameters N for a three-dimensional problem will be 3q. Since all of these

control parameters can change independently, the sensitivity of the cost function

95

with respect to each of these control parameters must be calculated. Therefore a

total of 3q sensitivities have to be calculated.

The boundary conditions for the direct flow calculation on the actuator are given

by,

u1 = ub1 , u2 = ub2 , u3 = ub3 . (6.71)

Where the subscript ‘b′ expresses a blowing velocity. The boundary conditions on

node ‘q’ can be written as,

u1 = ubq1 , u2 = ubq2 , u3 = ubq3 . (6.72)

The perturbation in these quantities with respect to the blowing velocities is either

1 or 0 depending on which component of the blowing velocity is considered. That

is, at the actuator node ‘q’, from Equation (6.72),

δuqiδubqj

= δij, (6.73)

where δij is Kronecker delta and has been defined earlier. Consider the boundary

term in the expression for δL, given in equation (6.33),

Boundary Integrand =(α(p− pd)k + P

)δp+ Uiδui + T δT. (6.74)

The coefficients of δui in the above expression remain non-zero because the δui

are non-zero in the actuator region. Hence, the boundary conditions in this re-

gion are obtained from the first and last terms (the coefficients of δT and δp), in

equation (6.74). They are given by,

P = T = 0, (6.75)

96

After simplification, the boundary conditions on the actuator are then given by,

q5 = − γ

(γ − 1)

qNuN

(6.76)

q1 = −(q2ubq1 + q3ubq2 + q4ubq3) +γ − 1

γ

ubqiubqi2

qNuN

(6.77)

Here, uN and qN are the direct and adjoint wall normal velocities, and are given

by,

uN = uini (6.78)

qN = q2n1 + q3n2 + q4n3 (6.79)

The remaining non zero terms on the actuator add to the gradient of the cost

function with respect to the actuator velocity components. These terms are given

by,

δLδubqi

= Ui for i = 1, 2, 3, (6.80)

where ui is the velocity component in the ith direction.

Equation (6.80) is used to calculate the sensitivity of the cost function to the

blowing velocities at the qth node.

6.4 The Optimization Procedure

The gradients of the cost function are calculated as described above. The details

of the adjoint optimization cycle are given in section 2.2. A gradient based search

method can then be incorporated to find the local optimum. A Newton’s search

method is used here to find the next set of design variables. The new vector of

blowing velocities is given by,

unewbi

= uoldbi− Jold

δL/δubi(6.81)

97

6.5 Results and Discussion

The method used to develop the adjoint equations and adjoint variables to calcu-

late the sensitivities of a defined cost function with respect to the blowing velocities

was described in the previous section. In this section, examples of the use of this

approach are given. Two and three-dimensional calculations are performed for

several cases. The first case is a two-dimensional example that recovers a pressure

distribution on the nozzle wall that is calculated using a known blowing distri-

bution. The second example is set such that two-dimensional ideally expanded

flow is achieved inside a nozzle when the nozzle is operating at off design con-

dition. The blowing velocities at the nozzle wall are used as control parameters

and the adjoint approach is used to optimize their values. Three-dimensional cal-

culations are then performed for both nozzle internal and external flows. First a

three-dimensional example demonstrates that the adjoint method can be used to

optimize the blowing velocity on fixed actuator regions on the nozzle wall. The

second three-dimensional example shows how the adjoint method can be used to

achieve a pressure distribution close to the ideally expanded flow when the noz-

zle is working at an off-design condition. The overall aim of these examples is to

demonstrate how the adjoint design cycles work and that the adjoint approach can

be used to optimize the blowing on a nozzle wall to meet the various objectives.

6.6 Two-dimensional Optimization of Blowing

This section describes two examples of the application of the adjoint optimization

method. The first case is set up to validate the method. In this case a blowing

distribution is first prescribed. The pressure distribution corresponding to this

blowing distribution is considered as the ‘desired’ pressure distribution. The ad-

joint method is then used to find the blowing distribution that achieves this desired

pressure distribution based only on the knowledge of desired pressure distribution.

98

The second case involves a situation where the desired blowing distribution is un-

known but the desired pressure distribution is prescribed. The desired pressure

distribution is chosen to be that corresponding to an ideally expanded flow. The

adjoint method is used to determine the blowing distribution needed to obtain a

pressure distribution close to this desired pressure distribution.

6.6.1 Example 1

A subsonic case is considered in this example. The nozzle has an exit to throat

area ratio of 1.18 for a design Mach number Md = 1.5. The nozzle pressure ratio

(NPR) is fixed at 1.1. The flow is subsonic throughout the nozzle for this nozzle

pressure ratio. The nozzle grid is generated using Gridgen and consists of 102

by 51 grid points in axial and transverse directions respectively. The number of

blowing ports is chosen to be 11 and the blowing is implemented only in the trans-

verse direction. Hence, the total number of design variables is 11, corresponding

to the transverse blowing velocity at each of the 11 ports. The actuator region is

kept fixed on the upper wall of the divergent section of the nozzle. The throat is

located at x/D= 2.9, where D is the throat diameter. The divergent section has

a non-dimensional length of 0.5, 80% of which is occupied by the actuator region

such that the actuator region starts at the throat.

The cost function is based on the difference between the pressure distribution at

any design cycle and the desired pressure distribution on the sensor as given by

Equation (6.12). The blowing distribution that generates the desired pressure

distribution is taken to be a Gaussian-like blowing distribution. This is shown as

the black line in Figure 6.1 and is given by,

vb = vboexp−ar2 (6.82)

where, vbo is the amplitude of the blowing velocity, r is the distance from the center

of the actuator region and is based on the actuator region length such that the

99

half length of the Gaussian blowing distribution is the same as the half length of

the actuator region so that the velocity goes to zero at the actuator boundaries.

The amplitude of the desired blowing, vbo is taken to be 30 m/s.

Fig. 6.1 Variation of the blowing distribution with design cycles.

A dual step MacCormack scheme is used for the discretization of flow and the

adjoint equations. Local time stepping with CFL equal to 0.5 is used to obtain a

faster convergence of the direct flow equations. The adjoint equations are solved

using the same grid and same numerical scheme as the direct flow equations. The

details of the numerical method are given in chapter 3. The time step required

for the stability of the adjoint calculations has been found to be one order of

magnitude smaller than the time step possible for the direct flow equations. The

convergence is accelerated if the boundary conditions are implemented after both

100

steps of the MacCormack scheme. Artificial dissipation with the same coefficients

for the direct and adjoint solutions has been added to both calculations. The cost

function is observed to decrease by 92 percent in three design cycles. The change

of the cost function with design cycles is summarized in the Table 1.

Table 6.1 The decay of the cost function with design cycles for two-dimensional calcula-

tions.

Design Cycle Cost Function(Pa2)

1 170713

2 122781

3 2864

Figure 6.1 shows the change in the blowing velocity distribution with design cycle.

The solid black line shows the Gaussian distribution that is used to calculate

the desired pressure. The green dotted line and blue dashed line show the blowing

velocity found after the second and third design cycles. The distribution of blowing

velocity found by the adjoint method (blue dashed line) is different than the original

blowing distribution but the area under the curve is similar for the two blowing

distributions. This implies that the mean flow rate is an important parameter in

the use of blowing for control. Figure 6.2 shows the desired pressure distribution

and the pressure distribution for the initial and third (final) design cycles. The

final and desired pressure distributions are very close even though the blowing

distributions are different. In this case, the adjoint method has found another

blowing distribution that gives the desired pressure distribution.

101

Fig. 6.2 Variation of the pressure distribution on the upper nozzle wall with design

cycles.

6.6.2 Example 2

It was shown in the previous example that the adjoint method can be used to

achieve a desired pressure distribution on the nozzle wall. The goal in the second

example is to achieve an ideally expanded flow, so the desired pressure distribution

corresponds to that of ideally expanded flow. The adjoint equations are solved to

find a blowing distribution that provides this desired pressure distribution. The

nozzle operating conditions are chosen to be such that there is a shock in the initial

flow and the adjoint method is used to generate a shock free flow field. A nozzle

domain and the boundaries are shown in figure 6.3. The nozzle geometry with Md

= 1.5 is considered. The nozzle pressure ratio (NPR) is kept fixed at 1.5. The

nozzle grid is generated using Gridgen and consists of 102 by 51 grid points in

102

the axial and transverse directions respectively. The number of blowing ports is

chosen to be 8 and the blowing is implemented only in the transverse direction.

The throat is located at x/D = 3.9. The divergent section has a non-dimensional

length of 0.6. The actuator region is distributed over 4.1 ≤ x/D ≤ 4.436, which

represents 56 percent of the divergent section of the nozzle. The cost function is

based on the difference between the pressure distribution at any design cycle and

the desired pressure distribution, as given in Equation (6.12). Figure 6.4 shows

the placement of the blowing ports and the sensor region in the diverging section

of the nozzle. Each grid point in the actuator region acts as a blowing port and

these locations are fixed for each design cycle. The sensor region is located on the

upper nozzle wall.

Γi Γo

Γw

Γw

Fig. 6.3 Two-dimensional nozzle domain and boundary conditions.

103

Fig. 6.4 The actuator and sensor regions on the divergent section of nozzle.

For the given nozzle pressure ratio of 1.5, the geometry of the nozzle, and the static

pressure imposed at the nozzle exit, the flow is over-expanded and there is a shock

in the diverging section of the nozzle. This pressure distribution in the nozzle is

shown in Figure 6.5. For a fixed area ratio and fixed pressure ratio, it is now desired

to find a blowing distribution to eliminate the shock in the nozzle. To obtain the

desired pressure distribution pd, the area ratio corresponding to subsonic isentropic

flow is considered while the pressure ratio is kept constant at 1.5. This area ratio

is 1.04, for a nozzle pressure ratio of 1.5. The flow inside the nozzle remains

subsonic in this case. A new geometry with this area ratio is generated and the

pressure distribution corresponding this new geometry is calculated. This pressure

distribution on the sensor is then taken to be the ‘desired’ pressure distribution.

This is shown in Figure 6.6. The initial geometry is now considered again and

the adjoint method is used to find the blowing distribution that gives the target

pressure distribution on the sensor. The cost function decreases from 512 Pa2 to

104

22 Pa2 in the first design cycle and in the second cycle it reduces slightly to 20

Pa2. The corresponding flow distribution is shown in Figure 6.7. A comparison

of Figures 6.6 and 6.7 shows that the two flow distributions are very similar even

though the area ratios are different for the two geometries. Figure 6.8 shows the

pressure distribution on the nozzle centerline for the initial (shocked), desired and

final design cycles. It can be seen that the final and desired pressure distributions

match well.

105

Fig. 6.5 Pressure distibution for initial shocked flow. Area ratio = 1.12, NPR = 1.5, no

blowing.

Fig. 6.6 Ideally expanded ‘desired’ pressure distribution. NPR = 1.5, area ratio = 1.04,

no blowing.

106

Fig. 6.7 Nozzle pressure distribution after two design cycles. NPR = 1.5, area ratio =

1.12, with blowing.

Fig. 6.8 Initial, first and desired pressure distributions on the nozzle centerline.

107

6.6.3 Three-dimensional Optimization of Blowing

Two-dimensional examples of the optimization of blowing on a nozzle wall to

achieve a desired pressure distribution were given in the previous section. Three-

dimensional cases are considered in this section. Nozzle flows are three-dimensional

and it is important to consider full three-dimensional calculations to see the ef-

fects of blowing. Effects such as the generation of stream wise vorticity can not

be studied by performing two-dimensional calculations. The blowing velocity in

the previous sections was considered to be normal to the wall. In this section the

three components of the blowing velocity are treated independently.

The goal of the present approach is to provide a method to achieve a desired noz-

zle pressure distribution, such as a nearly ideally expanded flow field, even when

the nozzle is operating at an off-design condition. In addition, the generation of

stream-wise vorticity can promote jet mixing and reduce the large scale mixing

noise. Both these effects can be achieved with the use of blowing in the diverging

section of the nozzle.

The overall geometry of the convergent - divergent nozzle under consideration is

shown in figure 6.9 by the surface mesh. The area ratio of this nozzle is 1.22.

The design Mach number and ideally-expanded nozzle pressure ratio for this noz-

zle are 1.56 and 4.0 respectively. The non-dimensional location of the throat is

x/D = 6.01, where D is the nozzle throat diameter. The divergent section takes

0.15% of the total length of the interior of the nozzle. The nozzle pressure ratio

has been kept constant at 1.5.

Figure 6.10 shows the mesh for the nozzle described above. It is a multi-block

structured orthogonal grid with 11x31x101 points in the outer blocks in the radial,

azimuthal and axial directions respectively and 31x31x101 points in the inner block

in the radial, azimuthal and axial directions respectively. The total number of

108

grid points is approximately 2.5 million with more clustering near the throat and

outflow. The smallest mesh size is ∆x/D = 0.0036 and ∆y/D = ∆z/D = 0.0021.

Figure 6.11 shows the placement of the actuators in the diverging section of the

nozzle. In this example problem there are two actuators placed opposite to each

other. The actuators occupy a non-dimensional area of approximately 4.7% of the

nozzle wall in the diverging section. Each grid point inside the actuator region

works as an independent control parameter. There are 2x77 such grid points

and the blowing velocity can vary independently at these points. The blowing is

implemented in such a way that there is no swirl in the injected fluid so that the

nozzle axis and the actuator axis are in the same plane. The angle of flow injection

relative to the wall is unconstrained.

Fig. 6.9 Nozzle showing surface mesh.

109

Fig. 6.10 Nozzle multi-block grid structure.

Fig. 6.11 The location of actuators (the blowing ports).

Fluidic inserts are known to reduce the broadband shock associated noise as well

as the mixing noise by changing the effective nozzle area ratio and introducing

streamwise vortices, as discussed by Morris et al. [30]. A method to find the

110

injection rate and distribution to achieve a desired flow behavior has been discussed

here. Figure 6.12 shows the pressure distribution inside the nozzle when there is

no blowing and the nozzle pressure ratio is fixed (NPR = 1.5). The jet is over-

expanded and a normal shock exists in the diverging section of the nozzle. The

contours represent the pressure on an azimuthal slice that passes through the

center of the two actuators. Figure 6.13 shows the pressure distribution on the

same plane when a prescribed blowing distribution has been implemented at the

actuators. The prescribed blowing distribution is given by a radial velocity,

ub = ubo exp

(−−(x− xo)2 − (y − yo)2 − (z − zo)2

2a2

)(6.83)

The blowing is implemented such that ubo is −150m/s, a is chosen to be 0.1 and

(xo, yo, zo) is the center for the blowing distribution, located at the center of the

actuator region. The positive normal at the wall is considered in the outwards di-

rection, hence the sign of the blowing speed is negative. For the two actuators, the

centers are located at a non dimensional axial distance xo/D = 0.39 downstream of

the throat, the other two coordinates are yo/D = zo/D = ±0.355. The injection of

the fluid changes the effective area ratio and even though the nozzle pressure ratio

is unchanged, there is no shock in the solution. Figure 6.14 shows the development

of counter rotating vortices caused by the blowing at the actuators. These vortices,

which are also generated in nozzles with hard-walled corrugations or fluidic inserts,

promote mixing in the jet plume, which can lead to mixing noise reductions. In the

practical situation, the blowing distribution that deflects the fluid is unknown, but

the desired pressure distribution is known to be either the ideally-expanded pres-

sure or some other prescribed pressure distribution. In the example given here, the

desired pressure distribution is considered to be the pressure distribution shown

in figure 6.13. The blowing that gives this pressure distribution is considered un-

known and the adjoint method is used to find a blowing distribution that achieves

this desired pressure distribution.

111

Fig. 6.12 The pressure distribution when there is no blowing in the divergent section.

Fig. 6.13 The pressure distribution with blowing in the divergent section.

112

Fig. 6.14 Introduction of counter rotating stream-wise vortices due to blowing.

The cost function to be minimized is given by equation (6.12), where the desired

pressure is shown in figure 6.13. The surface area in the integral is taken to be

the entire interior wall of the nozzle and supply pipe. The adjoint and direct

flow equations are solved on the same mesh as shown in figure 6.10. The spatial

terms are discretized using second order finite differences and a modified 4 stage

compact Runge-Kutta scheme is used to artificially march in time to a steady

state solution. The time step is determined by the Courant-Friedrichs-Lewy (CFL)

condition with a CFL number of 0.5. Artificial dissipation has been added to all

the equations for both calculations. The description of artificial dissipation is given

in section 3.3. The artificial dissipation coefficients in the x, y and z directions are

113

different for the direct and adjoint calculations and are chosen to be (1.3, 1.3, 1.3)

and (0.3, 0.3, 0.3) respectively. More details on the numerical method are given in

chapter 3. Figure 6.15a shows the initial difference between the desired and initial

pressures with no blowing. Figures 6.15a - 6.15d show the difference between

the current pressure distribution and the desired distribution with design cycles.

The contour scales are the same for figures 6.15a and 6.15d. The initial pressure

difference on the nozzle wall is of the order of 50,000 Pa and the final difference

goes down to the order of 5,000 Pa. The change in the cost function with design

cycle is given in table 6.2.

Table 6.2 Three-dimensional calculations: The decay of the cost function with design

cycles.

Design Cycle Normalized Cost Function Percentage decay

1 2.565x10−1 0

2 1.833x10−1 28

3 7.821x10−2 69

4 3.155x10−2 87

The decay of the cost function is almost linear with the design cycles. This is

shown in figure 6.16.

114

(a) Initial difference between desired and initial pressure on nozzle wall.

(b) Difference between desired and current pressure on nozzle wall after second

design cycle.

(c) Difference between desired and current pressure on nozzle wall after third

design cycle.

(d) Difference between desired and current pressure on nozzle wall after the fourth

design cycle.

Fig. 6.15 Variation of the pressure difference on the nozzle wall between the current

and desired pressure distribution with design cycle.

115

Fig. 6.16 The decay of cost function with design cycles.

Figures 6.17 and 6.18 show the pressure and Mach number distributions on the

nozzle centerline. The initial (grey) and desired (blue) pressure distributions are far

apart. After four design cycles, a pressure distribution close to the desired pressure

distribution is achieved using the adjoint method. Even though the desired blowing

distribution is known in the present case, it is not required by the adjoint solution

and the initial blowing velocities are set to zero. A relaxation factor of 10 has

been used for the first iteration to increase the convergence rate. After the first

iteration, this factor is no longer used in subsequent iterations. This process is

repeated until the cost function decreases to a value less than a specified tolerance.

From table 6.1 it is seen that the cost function reduces by 87% in just the first

four design cycles. The computational cost associated with each design cycle is

equal to the cost of solving the flow equations and adjoint equations once. This is

independent of the number of control parameters. There are 154 control parameters

in this example but the gradients of the cost function with respect to each of these

control parameters are calculated with no change in computational cost. This is

the major advantage of using the adjoint method.

116

Fig. 6.17 Pressure distribution on nozzle centerline with design cycles.

Fig. 6.18 Mach number distribution on nozzle centerline with design cycles.

Figure 6.19 shows the blowing that has been implemented to obtain the desired

or target pressure distribution. This blowing is applied normal to the wall in

the actuator region. Then the initial blowing normal to the wall is set to zero.

Figure 6.20 shows the blowing velocities normal to the wall after four design cycles

obtained using the adjoint method. The blowing distribution achieved by the

117

adjoint method is not same as, or even similar to, the blowing that corresponds

to the desired pressure, but the pressure distribution obtained with this blowing

is close to the desired pressure distribution as shown in figure 6.17 and by the

decrease in the cost function.

Fig. 6.19 Blowing velocity normal to the

wall used to generate the desired pressure

distribution.

Fig. 6.20 Blowing velocity normal to the

wall as given by fourth design cycle.

118

6.7 External Flow Calculations

The goal of the present research is to develop a tool to modify the shock cell

structure inside a nozzle and in the jet plume to achieve a reduction in broad-

band shock associated noise. In the previous section, it was shown that an adjoint

method can be used to recover a target pressure on the nozzle wall using surface

blowing. The present section is developed such that the external flow is included

in the calculation. It is shown that adjoint method can be used to modify the

shock cell structure using surface blowing on the nozzle wall.

The nozzle under consideration is the same as the nozzle described in section 6.6.3

and shown in figure 6.9. The external domain and shroud are added to the nozzle

using multiple additional blocks as shown in figure 6.21. The mesh is generated

using Gridgen and is shown in figure 6.22. The grid consists of 14 blocks with 5

blocks internal of the nozzle, 4 blocks in the domain external to the shroud and 5

blocks in the external domain upstream of the nozzle exit, constituting a total of

5.7 x 106 points. Figure 6.23 shows the placement of the actuators on the nozzle

wall. Three actuators are placed in the diverging section of the nozzle domain with

25 grid points in each actuator. The actuators are placed symmetrically on the

nozzle diverging section.

The spatial terms are discretized using second order finite differences and a modi-

fied 4 stage compact Runge-Kutta scheme is used to artificially march in time to a

steady state solution. The time step is determined by the Courant-Friedrichs-Lewy

(CFL) condition with a CFL number of 0.5. Artificial dissipation has been added

to all the equations for both calculations. The artificial dissipation coefficients in

the x, y and z directions are different for the direct and adjoint calculations and

are chosen to be (1.3, 1.3, 1.3) and (0.3, 0.3, 0.3) respectively. More details on the

numerical method are given in chapter 3.

119

Fig. 6.21 The placement of blocks when the external domain is included.

120

Fig. 6.22 The mesh including the external domain.

Fig. 6.23 The location of symmetrically placed actuators in the nozzle diverging section.

The approach based on the adjoint method is same as the previous section. The

details of the approach are omitted to avoid repetition. The exit to throat area

121

ratio of the nozzle is 1.176. The nozzle pressure ratio for ideal expansion and

the design Mach number for the nozzle under consideration are 3.6768 and 1.5

respectively. The operating conditions are such that the pressure ratio is 4.58

so the nozzle is under-expanded for this pressure ratio and there are expansion

waves in the nozzle jet plume. The Mach number contours are shown in figure

6.24. An adjoint method is incorporated to get the flow field close to the desired

flow as shown in figure 6.25. Three design cycles are run where the normalized

cost function drops from a value of 66.07 to 3.58. The cost function is normalized

with respect to the total pressure and unit length. The Mach number contours

corresponding to design cycles 1, 2 and 3 are shown in figures 6.26, 6.27 and 6.28

respectively. A comparison of the final (figure 6.28) and desired solution (figure

6.25) shows that the two solution are close to each other. This is also reflected by

the small value of cost function (3.5).

Fig. 6.24 The Mach number contours for the nozzle operating with 4.58 nozzle pressure

ratio and no blowing.

122

Fig. 6.25 The Mach number contours for the nozzle operating with 4.58 nozzle pressure

ratio and the desired blowing.

Fig. 6.26 The Mach number contours after the first design cycle for the nozzle operating

with 4.58 nozzle pressure ratio.

123

Fig. 6.27 The Mach number contours after the second design cycle for the nozzle oper-

ating with 4.58 nozzle pressure ratio.

Fig. 6.28 The Mach number contours after third design cycle for the nozzle operating

with 4.58 nozzle pressure ratio.

124

6.8 Conclusion

Two and three-dimensional calculations have been performed to achieve a desired

pressure distribution on the nozzle wall. The adjoint design method has been im-

plemented to determine the blowing distribution on a fixed region of the nozzle

wall known as the actuator. The blowing distributions as described have no con-

straints applied, except that the injection locations are fixed. This can be modified

by adding additional constraints on the blowing distribution. These constraints

could be applied to make sure that the blowing starts smoothly at the ends of

the actuator region. In addition the blowing velocity and its derivative could be

required to be zero at the end locations. These constraints can be incorporated

into the definition of the Lagrange functional. In both two and three-dimensional

cases, the adjoint method takes 3 - 4 design cycles to achieve a reduction of the

order of 60% to 80% in the cost function. When the desired pressure is calculated

using a known blowing distribution, the adjoint method is found to find a different

blowing distribution that gives a pressure distribution close to desired pressure dis-

tribution. This implies that there is more than one local minimum in the solution

space and the adjoint method finds one of these minima.

6.9 Summary

In this chapter, an adjoint approach to determine the required blowing on a nozzle

wall has been described. The blowing velocity components are considered to be

independent design parameters and the optimum value of the control parameters

is found using an adjoint approach. The advantage of using the adjoint approach

is that the computational cost of calculating the sensitivities of the cost function

to change in the control parameters is almost independent of the number of design

parameters. The adjoint approach is very effective in cases where there are sev-

eral design parameters to be considered. Several examples of the optimization of

125

blowing are given in this chapter. In the first example the desired pressure distri-

bution was calculated for a giving blowing distribution. In the second, the blowing

distribution required to achieve ideal expansion in the nozzle was obtained. First,

two-dimensional calculations were performed and the approach was then extended

to three-dimensional calculations. It has been shown that an adjoint design method

takes less than 10 design cycles to reach a pressure distribution close to the desired

pressure distribution in the cases considered.

Conclusions from the described approaches and results are given in the next chap-

ter. There are several ideas for future development and a lot of scope to the use of

adjoint equations. The adjoint approach can be used to optimize other quantities

such as vorticity generation. The adjoint approach can be used to optimize the

shape of the fluidic injectors or the nozzle exit seals. These ideas are discussed in

the final chapter.

126

Chapter 7Conclusions and Future Work

The use of adjoint methods for aerodynamic optimization has been increasing. The

two primary reasons for the use of adjoint methods are the flexibility they offer

in terms of choosing a cost function or control parameters and the computational

cost taken by adjoint methods is almost independent of the number of design pa-

rameters. Several design parameters can be considered without a great increase in

the computational cost that it takes to compute the sensitivities of a cost function

with respect to these design parameters. Motivated by the development of “fluidic

inserts,” an adjoint method to optimize the rate of injection such that a desired

pressure distribution is achieved inside the nozzle has been developed. Fluidic

inserts are devices that are used to inject by-pass air into the nozzle core flow to

achieve jet noise reduction. Two approaches to consider the blowing velocity on

the nozzle wall as the control parameters have been given.

Jet noise is generated due to the turbulent mixing of the jet shear layer with the

atmosphere (mixing noise) and due to the existence of shock cells in the jet plume

(broadband shock associated noise). The shock cell structure in the jet plume

exists when the nozzle is operating such that the exit pressure does not match

the ambient pressure. The pressure then adjusts to the ambient pressure through

shocks and expansions. It has been shown that the incorporation of flow injection

on the nozzle walls can help in achieving a reduction in both noise components [28].

127

The injection of the fluid works in two ways to reduce both noise components: the

broadband shock associated noise and the mixing noise. The injection of fluid on

the wall moves the flow away from the wall, actively “morphing” the nozzle inner

contour while keeping the geometry of the nozzle fixed. This active morphing of

the inner contour changes the effective exit to throat area ratio of the nozzle which

can help in achieving a flow field similar to the ideally expanded condition, even

when the nozzle is operating at off-design conditions. The injection of fluid also

introduces stream-wise vortices into the jet plume. These vortices help to increase

the turbulent mixing and result in a decrease in the mixing noise. Finding the

optimum rate of injection is a challenge. In this thesis an adjoint method has been

used to optimize the nozzle wall blowing as a first step in the optimization of the

fluid inserts.

Two adjoint approaches have been developed. The first approach is such that the

blowing velocity is written as a combination of a set of basis functions multiplied

by the appropriate coefficients. These coefficients of the basis functions are consid-

ered as the control parameters. An adjoint method is then developed to find the

optimum value of these design parameters. The nozzle geometry and operating

conditions are kept fixed. The location of the actuators (blowing ports) is also

kept fixed. Two examples of optimization are given. The first example considers

a Gaussian blowing distribution dependent on only one design parameter. The

desired or target pressure is found using a known blowing distribution. The ad-

joint method is then used to find the value of the design parameter: that is, the

amplitude of the blowing distribution. The amplitude of the blowing is recovered

within 1.6% error in 3 design cycles. The second example is set up to obtain a

nearly uniform pressure distribution in the nozzle divergent section. In this case

the blowing velocity depends on three design parameters. The optimum value of

these parameters is found using the adjoint method.

128

A second adjoint approach for the optimization of blowing has been developed to

offer more flexibility. The approach allows for the use of any kind of design/control

parameters. For practical purposes, the design parameters are chosen as the blow-

ing velocities at a number of fixed nozzle wall points. Two and three-dimensional

calculations are performed and it is demonstrated that with the use of the adjoint

method the optimum blowing distribution can be found in 4-5 design cycles. The

optimum blowing distribution is the blowing distribution that gives a pressure dis-

tribution close to the target distribution. This target pressure distribution can be

found in two ways. It can be found using a known blowing distribution, or for

more practical case, it can correspond to nearly ideally expanded flow. The blow-

ing to achieve a pressure distribution close to the target pressure is found using

the adjoint method. It is shown that with the use of the adjoint method, a blowing

distribution on the nozzle wall can be found such that the pressure distribution is

close to that for ideal expansion irrespective of the nozzle operating conditions.

A multi-block structured mesh is used for the three-dimensional calculations. Multi-

block grids are generated for relatively complex geometries. However, this can

result in grid singularities. The whole domain is divided into multiple Cartesian

blocks. The block interfaces are considered as separate boundaries. The required

block interface conditions for the direct and adjoint problems have been derived.

These boundary conditions are based on the method of characteristics. The resid-

uals (or time derivatives) in a block are either replaced or retained based on the

direction of propagation of information. This direction is determined by the char-

acteristics of the flow or adjoint equations. It is shown that adjoint characteristics

travel in the opposite direction to the direct flow characteristics.

129

Recommendations for Future Work

The scope for the future usage of adjoint methods in the field of noise reduction is

very wide. The development of the method presented in this thesis is generic and

can be modified to incorporate several changes and additions. The calculations

performed here are presented for the inviscid Euler equations. The same approach

can be used to add viscous effects and a turbulence model. An important next step

would be to perform unsteady calculations to be able to perform noise predictions.

The introduction of blowing on the nozzle wall introduces counter rotating vortices

in the flow field. These vortices help in reducing the mixing noise by increasing

the turbulent mixing. The strength of these vortices can be optimized using the

adjoint approach. A measure of the magnitude of the stream-wise vorticity can be

added to the cost function while continuing to require a targeted pressure distri-

bution.

The blowing distributions as described here have no constraints applied, except

that the injector locations are fixed. This can be modified by adding additional

constraints on the blowing distribution. These constraints could be applied to

make sure that the blowing starts smoothly at the edges of the actuator region.

For example, the blowing velocity and its derivative could be required to be zero

at the edge locations. These constraints can be incorporated into the definition of

the Lagrange functional. The approach described in this thesis can be extended to

implement these changes. However, the locations of the injectors can be limited

to the experimental locations such as in the experiments of Morris et al. [28].

Due to the presence of a shock there is a discontinuity in the cost function. In this

case the cost function can be redefined such that it is smooth even at the location

of the shock. Also, a term quantifying the pressure gradient can be added to the

cost function to reduce the shock and expansion strengths with the aim of reducing

130

the broadband shock associated noise.

The use of adjoint methods for shape optimization has been the subject of much

prior research. In the context of the present application, the adjoint method could

be used to optimize the shape of the injectors. The shape of the injectors can be

parameterized and an adjoint approach can be used to obtain an optimum shape

to achieve the greatest noise reduction with the use of fluidic inserts.

The cost functions described in the thesis have involved properties integrated over

a sensor surface or surfaces. Volume integrals can also be considered. For example,

the cost function could involve an integral of the pressure variations, or some other

parameter such as the turbulent kinetic energy, in the entire jet plume. In this

case the adjoint equations will contain source terms and these will drive the adjoint

solution, as opposed to the boundary condition forcing described in this thesis.

131

Appendix AThe Solver Development

A solver to find the numerical solution to the direct and adjoint equations and

for the calculation of sensitivities has been developed for the research presented in

this thesis. An optimizing technique has been implemented in the solver to find

the new set of design parameters. The details on the structure of the code, input

and output file formats and other details of the solver development are given in

this chapter.

A.1 Code Structure

All subroutines have been written in the FORTRAN90 language and the solver is

portable to the linux clusters. Figures A.1 and A.2 show the basic structure of the

code. The first step of the process is to calculate the desired flow conditions. The

direct solver is first called to solve for the desired flow based on the specific method

to calculate the desired flow as explained in chapters 5 and 6. This output is stored

in a file and then it is read by the direct-adjoint design solver. The direct-adjoint

design solver starts by calling the direct solver first to calculate a steady state

solution of the direct equations for a given set of design parameters. This solution

is then used by the adjoint solver that solves for the adjoint equations. Both adjoint

and direct solutions are then fed into the sensitivity calculation subroutine. The

sensitivity calculation subroutine first calculates the cost function and then finds

the sensitivities of the cost function with respect to the design parameters. These

132

sensitivities are used in an optimizing technique to calculate the design parameters

for the next design cycle.

133

Initialize

Allocate stack memory

Read input Grid and boundary condition files

Input parameters

Allocate heap (dynamic) memory

Read desired flow conditions Initialize flow variables

and fluxes

Direct flow solver

Start time loop

End time loop

Calculate flow fluxes

Calculate time derivatives Add dissipation

Flow variables at the next time step

Residual correction boundary condition

Boundary conditions on flow variables

Stages of time marching scheme

Overwrite new flow values in the old ones

Repeat based on the time marching scheme

Write intermediate restart files

Exit if convergence is achieved

Calculate time step

Print solution files

Fig. A.1 The code structure to find direct and adjoint solutions and optimization of the

cost function - continued to next figure

134

Adjoint flow solver

Initialize adjoint variables

Start time loop

End time loop

Calculate adjoint vectors

Calculate time derivatives Add dissipation

Adjoint variables at the next time step

Residual correction boundary condition

Boundary conditions on flow variables

Stages of time marching scheme

Overwrite new adjoint values in the old ones

Repeat based on the time marching scheme

Write intermediate restart files

Exit if convergence is achieved

Print solution files

Optimizer

Calculate sensitivities

Calculate new values of design parameters

Go to direct flow solution

Fig. A.2 The code structure to find direct and adjoint solutions and optimization of the

cost function

135

A.2 Input requirements

Input parameters

The following inputs are given to the solver,

1. Flow parameters

2. Grid file

3. Boundary condition file

Table A.1 Input Parameters: the values of fluid properties, convergence criteria, inlet

and ambient conditions, initial conditions and numerical parameters.

variable parameter

NPR Pressure ratio

P0 total pressure

pa ambient pressure

T0 total temperature

gama ratio of specific heats

mu coefficient of viscosity

gasconst gas constant

CFL Courant number

Cx, Cy, Cz Coefficients of dissipation

ki tmp intermediate solution printing frequency

tn Initial condition indicator

if 0; uniform initialization

if 1; gaussian pulse initialization

if 2; initial condition read from a file

136

Input files

There are four input files that are required by the solver. The boundary condition

file and grid file generated by gridgen are the input files for the solver. The other

two files are the gas properties file and the input parameters file.

Boundary conditions file

The boundary condition files are generated by gridgen when the solver is set to

generic. The boundary conditions are indicated by the regions of blocks and the

initial and final indices in all three directions. The index that defines the region

has the same value for initial and final index. The boundary condition is specified

by a number. The specific boundary conditions are noted in the table A.2.

Table A.2 Boundary Conditions: the number that specify the type of boundary condition

at a given boundary

Indicator Boundary Condition

2 Slip wall

3 Symmetry

5 Inflow

6 Outflow

8 Two dimensional

9 Atmospheric inflow

11 Radiation bc

51 Reimann variant

-1 Interface

The grid file is read in the PLOT3D file format. The grid metrics are calculated

inside the code using the values of (x, y, z) coordinates and a second order finite

137

difference scheme. A Fortran code fragment for a Gridgen style volume grid file

follows:

c.....number of grid points in block integer ni(nmax), nj(nmax), nk(nmax)

c.....grid point coordinates

real x(imax,jmax,kmax), y(imax,jmax,kmax), z(imax,jmax,kmax)

write(1) nblocks

do mb = 1, nblocks

write(1) ni(mb), nj(mb), nk(mb)

write(1) (((x(i,j,k),i=1,ni(mb)),j=1,nj(mb)),k=1,nk(mb)),

(((y(i,j,k),i=1,ni(mb)),j=1,nj(mb)),k=1,nk(mb)),

(((z(i,j,k),i=1,ni(mb)),j=1,nj(mb)),k=1,nk(mb))

end do

The boundary condition file is in the following format,

write flow solver id

write number of blocks

138

do number of blocks

write block imax, block jmax,

(and block kmax if 3D), block name

c........a region is a boundary condition or interblock connection write num-

ber of regions on block

c........a region is a boundary condition or interblock connection write number of regions on block

c..............For an interblock connection, there are three

c..............pairs of indices that must be matched (three on

c..............the source face and three on the target face).

c..............The index that’s constant on both the source and

c..............target faces is easy to determine simply by

c..............comparing the min and max indices and finding the

c..............pair in which min and max are the same.

c..............On each of the source and target faces, Gridgen

c..............writes one of the two remaining min-max index

c..............combinations as negative numbers. The negative

c..............min-max pair on each face correspond. Once you

c..............determine the corresponding negative indices

c..............just convert them to positive numbers.

c..............That leaves one pair of indices that align.

c..............Keep in mind that the max index may actually

c..............be less than the min index. You can either use

c..............a negative increment to step through the index

c..............or you can swap the min and max indices on BOTH

c..............source and the matching target indices.

139

write target region imin, target region imax, target region jmin, target region jmax,

(and target region kmin, target region kmax if target block number

endif

end do number of regions on block

end do number of blocks

A.3 Output files

The output files are solution files at a specified time interval. The output files can

be written in two file formats, namely, tecplot file format and plot3D file format.

140

Appendix BGoverning and Adjoint Equations inCurvilinear Coordinates

This sections lists the governing equations and adjoint used to solve the flow-field.

The Euler equations and adjoint equations as developed in chapter 6 are written

in curvilinear coordinates to solve the equations on a multi-block structured grid.

The grid is transformed using a grid transformation matrix from the physical do-

main (x, y, z) to the computational domain (ξ, η, ζ) and the governing equations

are re-written to comply with the computational domain coordinates.

The governing equations are presented here in three dimensions. The equations

for two dimensional or one dimensional calculations can be derived by setting the

derivatives in the ζ, or η and ζ directions to zero respectively. The generalized

coordinates (ξ, η, ζ) are a function of physical coordinates (x, y, z) and can be

written as equation B.1:

ξ = ξ(x, y, z)

η = η(x, y, z)

ζ = ζ(x, y, z)

(B.1)

141

and the Jacobian transformation matrix used to relate the computational and

physical coordinates is defined as given in equation B.2:

J =

∣∣∣∣∂(ξ, η, ζ)

∂(x, y, z)

∣∣∣∣ =

∣∣∣∣∣∣∣∣∣ξx ξy ξz

ηx ηy ηz

ζx ζy ζz

∣∣∣∣∣∣∣∣∣ =

∣∣∣∣∣∣∣∣∣xξ xη xζ

yξ yη yζ

zξ zη zζ

∣∣∣∣∣∣∣∣∣−1

(B.2)

where ∂ denotes the partial derivative, det denotes the determinant and subscripts

represent partial derivatives. The unit normal vector at a grid point along a

constant ξ value, denoted by ~lξ, is defined as:

~lξ =

ξx|5ξ| ,

ξy|5ξ| ,

ξz|5ξ|

(B.3)

where, | 5 ξ| =√ξ2x

+ ξ2y

+ ξ2z. Similar forms can be used to derive unit normal

vectors in the η and ζ directions. These values are typically used in boundary

condition implementation: for example, to calculate the flow direction (inflow or

outflow) on a computational face. The transformation from physical to computa-

tional domain is applied to the dimensional direct and adjoint Euler equations and

the transformed equations are obtained as shown in equation B.4:

∂Q

∂t+∂E

∂ξ+∂F

∂η+∂G

∂ζ= 0 (B.4)

Variables with hat, for example Q, are in generalized curvilinear coordinates. They

are a function of flow variables in Cartesian coordinates and are related to them

as shown in equations B.5 - B.8:

Q =Q

J=

1

J

ρ

ρu

ρv

ρw

ρe

(B.5)

142

E =ξxE + ξyF + ξzG

J=

1

J

ρU

ρUu+ ξxp

ρUv + ξxp

ρUw + ξxp

(ρe+ p)U

(B.6)

F =ηxE + ηyF + ηzG

J=

1

J

ρV

ρV u+ ηxp

ρV v + ηxp

ρV w + ηxp

(ρe+ p)V

(B.7)

G =ζxE + ζyF + ζzG

J=

1

J

ρW

ρWu+ ζxp

ρWv + ζxp

ρWw + ζxp

(ρe+ p)W

(B.8)

In the equations B.6 - B.8, U , V and W are the contravariant velocities in the

three generalized coordinate directions. The variables ρ, p and e are the density,

pressure and total energy respectively. u, v and w are the velocity components in

the three directions. The contravariant velocities can be calculated as:

U = ξxu+ ξyv + ξzw (B.9)

V = ηxu+ ηyv + ηzw (B.10)

W = ζxu+ ζyv + ζzw (B.11)

143

The derivatives such as ∂/∂x can be calculated from the generalized coordinates

as,

∂x= ξx

∂ξ+ ηx

∂ξ+ ζx

∂ξ(B.12)

∂y= ξy

∂ξ+ ηy

∂ξ+ ζy

∂ξ(B.13)

∂z= ξz

∂ξ+ ηz

∂ξ+ ζz

∂ξ(B.14)

The partial derivatives can be calculated by using spatial discretization schemes

such as MacCormach scheme. The grid transformation matrices are calculated

using the second order central differencing schemes. Since the form of adjoint

equations are exactly the same as the direct equation as given by (B.4), exactly

the same approach is used to transform the adjoint equations from physical space

to the computational domain.

144

Bibliography

[1] M. B. Alkislar, A. Krothapalli, and G. W. Butler. The effect of streamwise vorticeson the aeroacoustics of a Mach 0.9 jet. Journal of Fluid Mechanics, 578:139–169,2007.

[2] T. R. Bewley, P. Moin, and R. Temam. DNS-based predictive control of turbu-lence: an optimal benchmark for feedback algorithms. Journal of Fluid Mechanics,447:179–225, 2001.

[3] K. S. Brentner and F. Farassat. Analytical comparison of the acoustics analogy andkirchhoff formulation for moving surfaces. AIAA Journal, 36(8:1379–1386, 1998.

[4] J. Bridges and C. A. Brown. Parametric testing of chevrons on single flow hot jets,2004. AIAA Paper 2004-2824.

[5] J. Bridges and C. A. Brown. Parametric testing of chevrons on single flow hotnozzles, 2004. AIAA Paper No. 2004-2824.

[6] S. S. Collis, K. Ghayour, and M. Heinkenschloss. Optimal transpiration boundarycontrol for aeroacoustics. AIAA Journal, 41:1257–1270, 2003.

[7] A. W. Correia, J. L. Peters, J. I. Levy, S. Melly, and F. Dominici. Residentialexposure to aircraft noise and hospital admissions for cardiovascular diseases: multi-airport retrospective study. British Medical Journal, 3:347:f5561, 201.

[8] M. J. Doty and D. K. McLaughlin. Acoustic and mean flow measurements of high-speed, helium-air mixture jets. International Journal of Aeroacousticsl, 50:293–334,2003.

[9] Y. Du. Supersonic jet noise prediction and noise source investigation for realisticbaseline and chevron nozzles based on hybrid RANS/LES simulations. PhD thesis,Pennsylvania State University, 2011.

[10] J. Freund. Adjoint-based optimization for understanding and suppressing jet noise.Procedia Engineering, 6:54–63, 2010.

[11] M. B. Giles and N. A. Pierce. An introduction to the adjoint approach to design.Flow, Turbulence and Combustion, 65(3-4):393–415, 2000.

145

[12] M. Hayashi, M. Ceze, and E. Volpe. Characteristics-based boundary conditions forthe euler adjoint problem: 2-d formulation. In 28th AIAA Applied AerodynamicsConference, Chicago, Illinois, 2010. AIAA Paper No. 2010-4677.

[13] B. Henderson and J. Bridges. An MDOE investigation of chevrons for supersonicjet noise reduction, 2010. AIAA Paper No. 2010-3926.

[14] A. Jameson. On optimum design in fluid mechanics. Journal of Scientific Comput-ing, 3:233–260, 1988.

[15] A. Jameson. Optimum aerodynamic design using CFD and control theory, 1995.AIAA Paper 95-1729.

[16] R. D. Joslin, M. D. Gunzburger, and R. A. Nicolaides. Self-contained automatedmethodology for optimal flow control. AIAA Journal, 35(5):816–824, 1997.

[17] B.-J. Jung, H. M. Kim, X. Z. T. Aulign, X. Zhang, , and X.-Y. Huang. Adjoint-derived observation impact using wrf in the western north pacific. American Mete-orological Society Journal, 141:4080–4097, November 2013.

[18] J. Kim, D. J. Bodony, and J. B. Freund. Aeroacoustics control of a turbulent mach1.3 jet using adjoint-based optimization, 2011. AIAA Paper 2011-2832.

[19] J.-H. Kim, Kearney-Fischer, M. M., Samimy, and S. Gogineni. Far-field noise controlin supersonic jets using conical and contoured nozzles. ASME Journal of Engineer-ing for Gas Turbine and Power, 133:081201–1 to 081201–9, August 2011.

[20] J. W. Kim and D. J. Lee. Characteristic interface conditions for multiblock high-order computation on singular structured grid. AIAA Journal, 41:2341–2348, 2003.

[21] K. Kinzie, B. Henderson, J. Whitmire, and A. Abeysinghe. Fluidic chevrons forjet noise reduction, 2004. ACTIVE 04: 2004 International Symposium on ActiveControl of Sound and Vibration; 20-22 Sep.; Williamsburg, VA; United States.

[22] A. Kohl and Willerband. An adjoint method for the assimilation of statisticalcharacteristics into eddy-resolving ocean models, 2002. Tellus A, 54: 406-425. doi:10.1034/j.1600-0870.2002.01294.x.

[23] A. Krothapalli, L. Ventkatakrishnan, L. Lourenco, B. Greska, and R. Elavarasan.Turbulence and noise suppression of a high-speed jet by water injection. Journal ofFluid Mechanics, 491:131–159, 2003.

[24] C.-W. Kuo, P. J. Morris, and M. D. K. Noise reduction in supersonic jets by fluidicnozzle inserts, 2012. AIAA Paper 2012-2210.

[25] B. J. Lee and M.-S. Liou. Unsteady adjoint approach for design optimization offlapping airfoils. AIAA Journal, 50:2460–2475, 2012.

[26] J. L. Lions. Optimal Control of Systems Governed by Partial Differential Equations.Springer-Verlag, Berlin, 1971.

146

[27] D. K. McLaughlin, J. Bridges, and C.-W. Kuo. On the scaling of small, heat sim-ulated jet noise measurements to moderate size exhaust jets. International Journalof Aeroacoustics, 9:627–654, 2010.

[28] P. J. Morris and D. K. McLaughlin. Methods and apparatus for providing fluidicinserts into an exhaust stream to reduce jet noise from a nozzle. USA ProvisionalPatent Application 61/648,441, May 2012.

[29] P. J. Morris and D. K. McLaughlin. High speed jet noise reduction: past progressand future possibilities, 2013. AIAA Paper No. 2013-0006.

[30] P. J. Morris, D. K. McLaughlin, and C.-W. Kuo. Noise reduction in supersonic jetsby nozzle fluidic inserts. Journal of Sound and Vibration, 332:3992–4003, 2013.

[31] P. J. Morris, N. Sikarwar, D. K. McLaughlin, C. W. Kuo, and M. Lurie. Adjointdesign methods for nozzles for reduced noise in high-speed jets, 2012. AIAA PaperNo. 2012-3821.

[32] D. Munday, N. Heeb, E. G. Markus, O. Burak, and L.-E. Eriksson. Supersonicjet noise from a conical c-d nozzle with forward flight effects. AIAA Journal,50(11):2336–2350, Nov 2012.

[33] S. Nadarajah and A. Jameson. A comparison of the continuous and discrete adjointapproach to automatic aerodynamic optimization, 2000. AIAA 38th AerospaceSciences Meeting and Exhibit, AIAA-2000-0667.

[34] C. Nelson. An overview of the nparc alliance’s wind-us solver, 2010. AIAA Paper2010-0027l.

[35] M. Nemec, M. J. Aftosmis, and M. Wintzer. Adjoint-based adaptive mesh refine-ment for complex geometries, 2008. 46th AIAA Aerospace Sciences Meeting.

[36] B. P. Petitjean, D. K. McLaughlin, and K. Viswanathan. Acoustic pressure wave-forms measured in high speed jet noise experiencing nonlinear propagation. Inter-national Journal of Aeroacoustics, 5:193–215, 2006.

[37] O. Pironneau. On optimum design in fluid mechanics. Journal of Fluid Mechanics,64:97–110, 1974.

[38] R. W. Powers, D. K. McLaughlin, P. J. Morris, and C. W. Kuo. Supersonic jetnoise reduction by nozzle fluidic inserts with simulated forward flight. In 20thAIAA/CEAS Aeroacoustics Conference, Atlanta, GA, 16 - 20 June 2014.

[39] M. Samimy, J.-H. Kim, J. Kastner, I. Adamovich, and Y. Utkin. Active control ofhigh speed and high reynolds number jets using plasma actuators. Journal of FluidMechanics, 578:305–330, May 2007.

[40] J. M. Seiner. Advances in high speed jet aeroacoustics, 1984. AIAA Paper 84-2275.

[41] J. M. Seiner. Method and device for reducing engine noise, January 2009. USAPatent US 7,475,550 B2.

147

[42] J. M. Seiner, L. S. Ukeiley, and B. Jansen. Aero-performance efficient noise reductionfor the F404-400 engine, 2005. AIAA Paper No. 2005-3048.

[43] N. Sikarwar. The use of adjoint methods in jet nozzle design, 2009. Masters Thesis,Pennsylvania State University.

[44] N. Sikarwar and P. J. Morris. Optimization of blowing in a convergent-divergentnozzle for noise reduction. In 20th AIAA/CEAS Aeroacoustics Conference, 16 - 20June 2014. AIAA Paper No. 2014-2473.

[45] N. Sikarwar and P. J. Morris. The use of adjoint method for optimization of blowingin a convergenct-divergent nozzle. International Journal of Aeroacoustics, 14:327–351, 2015.

[46] D. A. Venditti and D. L. Darmofal. Adjoint-derived observation impact using wrfin the western north pacific. Journal of Computational Physics, 164:204–227, 2000.

[47] K. Viswanathan, M. Shur, M. Strelets, and P. R. Spalart. Numerical prediction ofnoise from round and beveled nozzles, July 18-20, 2005. Euromech Colloquium 467:Turbulent Flow and Noise Generation.

[48] Q. Wang, P. Moin, and G. Iaccarino. Minimal repetition dynamic checkpointingalgorithm for unsteady adjoint calculation. SIAM Journal on Scientific Computing,31:25492567, 2009.

[49] M. Wei and J. B. Freund. A noise-controlled free shear layer. Journal of FluidMechanics, 546:123–152, 2006.

NIDHI [email protected]

412-638-9887

Education

• Pennsylvania State UniversityPh.D. in Aerospace Engineering, defended December 2014 GPA: 3.93/4.0

– Ph.D. minor in Computational Science GPA: 4.0/4.0

• Pennsylvania State UniversityM. S. in Aerospace Engineering, graduated December 2009 GPA: 3.96/4.0

• Indian Institute of Technology, Kanpur, IndiaB. Tech. in Aerospace Engineering, graduated June 2007 GPA: 8.40/10

Work Experience

• The Use of Adjoint Method in Optimization of Blowing in a C-D Nozzle to Achieve Noise ReductionPennsylvania State University, Graduate Research Assistant, Advisor: Dr. Philip Morris

– Developed and implemented adjoint method for optimization of active flow control.

– Developed a multi-block, three-dimensional, parallel RANS solver for the nozzle flows.

– Developed a multi-block, three-dimensional, parallel adjoint solver.

– Developed and implemented a method of characteristics block interface boundary conditions for adjointequations.

– Generated various 3D multi-block grids using CAD designs and gridgen to run the above solvers.

• LES Simulation of Supersonic Jet FlowPennsylvania State University, Graduate Research Assistant, Advisor: Dr. Philip Morris

– Gained experience in working with large, multi-modular, parallel code.

– Provided support to Pratt & Whitney for setting up, running and troubleshooting the code.

• Development of Adjoint Method to Design the Nozzle Contour in Order to Optimize the NoisePennsylvania State University, Graduate Research Assistant, Advisor: Dr. Philip Morris

– Developed and implemented Adjoint method for automatic optimization of the nozzle geometry of a jetaircraft.

Publications

• Nidhi Sikarwar, Philip Morris: “The Use of an Adjoint Method for Optimization of Blowing in aConvergent-Divergent Nozzle”, International Journal of Aeroacoustics, vol 14, number 1 & 2, pp 327 - 351,2015.

• Nidhi Sikarwar, Philip Morris: “Characteristics Boundary Conditions to Treat Adjoint Block Interfaces”,IEEE Aerospace Conference 2015, MAR 7 - 15, 2015, Big Sky, MT.

• Nidhi Sikarwar, Philip Morris: “Optimization of Blowing in a Convergent-Divergent Nozzle for NoiseReduction”, AIAA Paper No. 2014,-2473, 2014.

• Philip Morris, Nidhi Sikarwar, D. K. McLaughlin, C. W. Kuo and M. Lurie: “Use of AdjointDesign Methods for Nozzles for Reduced Noise in High-speed Jets”, AIAA Paper No. 2012-3821, 2012.

• Nidhi Sikarwar, Satya Prakash, Biju Uthup: “Understanding Viscous Flow Field on a Cropped DoubleDelta Wing at High Angles of Attack”, 9th annual CFD symposium, The Aeronautical Society of India andNational Aeronautical Laboratories, Bangalore, August 2006.

• Nidhi Sikarwar, Ashutosh Chauhan: “Use of Machine Learning to Optimize Fluidic Inserts”, acceptedby : AVIAION 2015.