98
A350 Take-off Configuration Optimization using a Surrogate-based Steepest Descent Method Miguel Afonso Rita Thesis to obtain the Master of Science Degree in Aerospace Engineering Supervisors: Prof. Fernando José Parracho Lau Dr. Julien Delbove Examination Committee Chairperson: Prof. Filipe Szolnoky Ramos Pinto Cunha Supervisor: Prof. Fernando José Parracho Lau Member of the Committee: Dr. José Lobo do Vale November 2014

A350 Take-off Configuration Optimization using a Surrogate ... · A350 Take-off Configuration Optimization using a Surrogate-based Steepest Descent Method ... gradient steepest

Embed Size (px)

Citation preview

A350 Take-off Configuration Optimization using aSurrogate-based Steepest Descent Method

Miguel Afonso Rita

Thesis to obtain the Master of Science Degree in

Aerospace Engineering

Supervisors: Prof. Fernando José Parracho LauDr. Julien Delbove

Examination Committee

Chairperson: Prof. Filipe Szolnoky Ramos Pinto CunhaSupervisor: Prof. Fernando José Parracho LauMember of the Committee: Dr. José Lobo do Vale

November 2014

ii

For my family, unconditionally there, always.

iii

iv

Acknowledgments

First and foremost, I would like to express my gratitude to my internship tutor at Airbus, Mr. Julien Del-

bove, for his constant support, expertise and commitment. His overall guidance and trust were what

made it all possible.

I would also like to address a big thank you to Professor Fernando Lau and Professor Jose Vale for

their efforts, feedback and help in preparing this thesis for presentation here in Portugal.

Additionally, a deep thank you in general to the Instituto Superior Tecnico, Sup’Aero and Airbus, namely

to all the teachers and staff who work in these places, and who empower students to take challenges

akin to this one.

Last but certainly not least, here’s to my pals Cardeira, Brinco and Clemente. For your friendship, I

can only be grateful.

v

vi

Resumo

Actualmente, a determinacao de configuracoes de descolagem optimas de uma aeronave e feita recor-

rendo a extensas campanhas de ensaios em voo, com um elevado custo monetario associado. Este

custo limita frequentemente nao so a qualidade dos resultados obtidos, mas tambem o numero de

parametros que sao optimizados. O presente trabalho visa a implementacao de um algoritmo de

optimizacao inovador, baseado no metodo do gradiente, que permita determinar configuracoes optimas

de descolagem de uma maneira sistematica, economica e extensıvel. Para alcancar estes objectivos,

o algoritmo recorre a modelos substitutos e planos de experiencia optimos. Os resultados obtidos

mostram que e de facto possıvel atingir optimos de uma maneira muito mais economica que no pas-

sado, desde que a modelacao dos dados seja feita cuidadosamente. Uma nova funcao objectivo es-

tatıstica baseada num metodo de propagacao de incerteza foi tambem implementada, com resultados

que vem suportar o que foi feito ate entao na industria. Para concluir, o trabalho desenvolvido demon-

strou nao so a aplicabilidade do metodo, mas tambem a sua extensibilidade a problemas diferentes ou

de diferentes dimensionalidades, abrindo as portas a poupancas de tempo e ganhos monetarios muito

significativos nas futuras campanhas de configuracao optima.

Palavras-chave: Performance aviao, Configuracao de descolagem, Optimizacao baseada

em modelos, Plano de experiencia, Metodo do gradiente.

vii

viii

Abstract

At present, determination of optimal aircraft takeoff configurations is done resorting to very expensive

and time-consuming flight testing. This often limits not only the quality of the optima found but also

the number of configuration parameters that can be optimized. The present work implements a novel

gradient steepest descent optimization algorithm to tackle this problem in an economically viable way.

To achieve this, it resorts to surrogate modelling and optimal design of experiment techniques, allowing

for flexibility and optimal information gain in process, respectively. Results obtained show satisfactory

accuracy as long as great care is taken in the modelling and interpolation of data. An innovative statistical

assessment of configuration optimality solidly supports the optima obtained thus far in the industry. This

work confirms the applicability of method for real-time flight testing and, more importantly, its scalability

and adaptability to different configuration problems or problem dimensionality, opening up the door for

potentially large savings in future optimal configuration campaigns.

Keywords: Aircraft Performance, Takeoff configuration, Surrogate-based optimization, Design

of experiment, Gradient steepest descent.

ix

x

Contents

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

Resumo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvi

Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii

Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xix

1 Introduction 1

1.1 Context and motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Problem statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.3 Possible solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.4 Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Aircraft Takeoff Optimization 4

2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.2 Takeoff Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.3 Air conditioning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.4 Takeoff Speed Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.4.1 Speed Ratios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.4.2 V1/VR Ratio Influence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.4.3 V2/VS Ratio Influence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.5 Optimization Process & Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

3 Surrogate Modelling 10

3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3.2 Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3.2.1 Linear regression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3.2.2 Response Surface Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

3.2.3 Gaussian Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

3.2.4 Splines in tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

xi

4 Design of Experiment 16

4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

4.2 General DoE Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

4.3 Optimal DoE Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

5 Uncertainty Propagation 21

5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

5.2 Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

5.2.1 Taylor Moment Propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

5.2.2 Gaussian Quadrature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

5.2.3 Monte Carlo Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

5.2.4 Univariate Reduced Quadrature . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

6 Optimal Configuration Search 25

6.1 Core Tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

6.1.1 OCTOPUS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

6.1.2 OPTIMA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

6.1.3 MACROS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

6.2 Overall Architecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

6.2.1 Gradient Descent Rationale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

6.2.2 Algorithm Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

6.3 Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

6.3.1 DoE Generator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

6.3.2 FT Simulator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

6.3.3 Objective Calculator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

6.3.4 Gradient Calculator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

6.3.5 Line Designer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

6.3.6 Line Builder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

6.3.7 Line Optimizer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

7 Results 51

7.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

7.2 Single MTOW, Fixed Runway . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

7.3 Optimal Configuration, Fixed Runway . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

7.4 Different OC, Varying Runway . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

7.5 Statistically Optimal Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

8 Conclusion 72

Bibliography 78

xii

List of Tables

2.1 TO Parameters Categorization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.2 Fixed parameters for V1/VR study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

3.1 Different RSM types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

6.1 CoH steps - module correspondence. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

6.2 FT source data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

7.1 Optimal Configuration for RL 4000m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

7.2 3 Configuration Sets, Traditional and Optimal (∗) . . . . . . . . . . . . . . . . . . . . . . . 67

7.3 Different TO network coverage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

7.4 Statistical OC results and their standard counterparts. . . . . . . . . . . . . . . . . . . . . 71

xiii

xiv

List of Figures

2.1 V1/VR MTOW limitations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

4.1 Optimal and non-optimal latin hypercubes . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

6.1 CoH algorithm basic workflow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

6.2 Flap/Slat design space. The optimum configuration is represented by a light yellow star. . 30

6.3 DoE around initial guess. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

6.4 FT results are used for a MTOW TO optimization. . . . . . . . . . . . . . . . . . . . . . . . 31

6.5 Configuration overloading using OPTIMA. . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

6.6 Sampling along gradient line. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

6.7 Maximum over gradient line becomes next guess. . . . . . . . . . . . . . . . . . . . . . . 34

6.8 CoH implementation modular workflow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

6.9 CZ as function of the angle of attack α and Flap deflection. Cut at ISO median. . . . . . . 38

6.10 CX as function of α and Flaps. Cut at ISO median. . . . . . . . . . . . . . . . . . . . . . . 38

6.11 Alpha influence on output. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

6.12 Flap deflection influence on output. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

6.13 CX as function of α and Slat deflection. Cut at ISO max. . . . . . . . . . . . . . . . . . . . 40

6.14 CZ and CX as function of Slat deflection. Cut at ISO max. . . . . . . . . . . . . . . . . . . 41

6.15 CZ and CX as function of Aileron deflection. Cut at ISO mean. . . . . . . . . . . . . . . . 42

6.16 CZ and CX as function of CG (Center of Gravity) position. Cut at ISO mean. . . . . . . . 43

6.17 CZ and CX as function of the Engine setting. Cut at ISO mean. . . . . . . . . . . . . . . . 44

6.18 MTOW calculator architecture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

6.19 ACZMAX GP SM label CZMAXVS Flap and Slat deflection 3D hyper-cut, at ISO minimum. 45

6.20 CZMAXsurrogate model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

6.21 Objective Calculator architecture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

6.22 RL probability density and histogram. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

7.1 RSM and LR interpolation results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

7.2 Lift and drag (CZ and CX ) curve plots. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

7.3 Lift and drag (CZ and CX ) error analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

7.4 MTOW as function of configuration, using GP SM based on FF limit design. Note the OC,

indicated by a yellow star. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

xv

7.5 MTOW value progression for the full DS, depending on DoE sample size S and technique. 56

7.6 Absolute MTOW error for the full DS, depending on DoE sample size S and technique. . . 56

7.7 MTOW value progression for a small sampling domain, depending on DoE sample size S

and technique. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

7.8 Absolute MTOW error for a small sampling domain, depending on DoE sample size S and

technique. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

7.9 Slat deflection for a small sampling domain, depending on DoE sample size S and tech-

nique. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

7.10 Flap deflection for a small sampling domain, depending on DoE sample size S and tech-

nique. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

7.11 3D GP SM cut at ISO-max of CoH output as function of local DoE pool sample size and

number of iterations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

7.12 3D cuts at ISO-average of CoH output as function of gradient line sample size and number

of iterations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

7.13 2D cuts showing MTOW convergence with increasing number of iterations. . . . . . . . . 63

7.14 CoH full parameterization and run results. . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

7.15 CoH Configuration and Objective convergence. . . . . . . . . . . . . . . . . . . . . . . . . 66

7.16 Traditional and CoH TO complete network coverage for the three different configurations. 68

7.17 Traditional and CoH configuration 1+F TO network detail. . . . . . . . . . . . . . . . . . . 68

7.18 Traditional and CoH configuration 2 TO network detail. . . . . . . . . . . . . . . . . . . . . 69

7.19 Traditional and CoH configuration 3 TO network detail. . . . . . . . . . . . . . . . . . . . . 69

7.20 Optimal Configuration evolution with Runway Length. . . . . . . . . . . . . . . . . . . . . . 70

xvi

Nomenclature

CX Drag coefficient

CZ Lift coefficient

CZMAXMaximum lift coefficient

dinput Number of model inputs

doutput Number of model outputs

g1 Domain border parameter

GS Set of gradient line points

∇G Normalized masked gradient

K1 Local DoE hyperbox size parameter

lk Design space k-th dimension lower bound

mk Local design of experiment k-th dimension lower bound

Mk Local design of experiment k-th dimension upper bound

Pu,v Multi-dimensional box with delimiters u, v

S Training set matrix

Uk Design space k-th dimension upper bound

X Input (design of experiment) part of the training set

Y Output (response) part of the training set

Y Output surrogate estimate

α Angle of attack

α Linear regression coefficients estimate

αi, βi,j RSM coefficients

γX Distribution skewness

ΓX Distribution kurtosis

µf Random variable function mean

µX Distribution mean

ψ RSM design

ρ Minimax interpoint distance

σf Random variable function standard deviation

σX Distribution standard deviation

xvii

xviii

Abbreviations

AFM Aircraft Flight Manual

APU Auxiliary Power Unit

ASD Accelerate-Stop Distance

DD Design Domain

DoE Design of Experiment

DS Design Space

GP Gaussian Process

GQ Gaussian Quadrature

GWN Gaussian White Noise

LHS Latin Hypercube Sampling

LR Linear Regression

MM Method of Moments

MTOW Maximum TakeOff Weight

OF Objective Function

OLHS Optimized Latin Hypercube Sampling

RDO Robust Design Optimization

RRE Ridge Regression Estimate

RS Random Sampling

RSM Response Surface Model

SBO Surrogate Based Optimization

SP Sigma Point

SPLT Splines with Tension

TO TakeOff

TOD TakeOff Distance

TOW TakeOff Weight

UP Uncertainty Propagation

URQ Univariate Reduced Quadrature

xix

xx

Chapter 1

Introduction

The present thesis tackles the issue of developing and implementing a methodology to find optimal

aircraft TakeOff (TO) configurations according to different optimality criteria and budgetary constraints.

In this brief introduction, the context and motivation behind the project are first explained, followed by

a clear problem statement. Possible solution outlines are then advanced, and finally the thesis basic

structure is detailed.

1.1 Context and motivation

Since the advent and recognition of engineering as a science, the search for ways to model and predict

physical phenomena has always been at the heart of the practice. In order to do so, in the past as

well as in the present, empirical experiments and tests are carried out. Nowadays, modern computer

science complements those designs and analyses via computer models and simulations. In certain

cases, these allow us to bypass the elevated monetary costs associated with building concrete, tangible

experiments. However, the complexity and detail of today’s scientific problems, coupled with an ever-

increasing demand for accuracy, accounting for second-order effects, finer discretizations, among other

reasons [1], can potentially render computer simulations very time-consuming. This is particularly true

in a complex, multidisciplinary domain like aerospace engineering [2]. More specifically, looking into

the field of aircraft performance, time-consuming optimization loops in simulators (TO optimization, for

instance) are a reality, as well as the heavy reliance on real flight test campaigns. This exacerbates the

problem, as flight tests are both extremely expensive and very time-consuming to carry out. On top of

this, the dream of eventually developing computer models accurate enough to forgo true flight testing is

not even theoretically possible, since certification bodies in this field are ruthlessly unyielding concerning

the subject. Thus, if there is to be any hope of undertaking any complex problem in this domain, namely

finding and defining optimal TO configurations in a systematic and economic fashion, we will forcefully

have to resort to ingenious and ground breaking approaches. A possible solution is detailed later in this

chapter.

1

1.2 Problem statement

In essence, the aim of this project is to define and implement an efficient, reliable and economical way

to find the optimal TO configuration for an aircraft. This main ultimate goal, however, is rather generic.

Four other more specific key question/objective couples can be derived from our main goal:

1. What does optimal configuration mean? - Define and quantify the optimality criteria.

2. Given a well-defined objective function, how to carry out the actual optimization process? - Imple-

ment an optimization algorithm to find the configuration.

3. After finding possible solutions, how fast/cheap can we be without compromising accuracy? - Take

algorithm implementation and improve its architecture, analyze and discuss results.

4. Is our method compatible or adaptable to existing flight test procedures? - Devise a module to

interface with the flight testing team.

The approach we come up with must be able to deal satisfactorily with all four aforementioned ques-

tions.

1.3 Possible solutions

One increasingly popular and promising method to deal with expensive optimization problems is Surrogate-

Based Optimization (SBO) [3, 4, 5], or something derived from it. Consider the following generic mini-

mization problem:

x∗ = arg minx

f(x) (1.1)

where

• x∈ Rn is our design vector

• x∗ denotes the optimal design vector

• f(x) is our expensive resource or function

Using traditional optimization algorithms, with our simulator integrated on the loop, the number of ”calls”

to this expensive resource f(x) would normally be too expensive to bear. Surrogate models attempt to

circumvent this in a simple away: they are nothing more than ”cheap” approximations of our ”expensive”

function. They offer a relatively high accuracy for a fraction of the cost, since to create the surrogate we

just have to evaluate the expensive function at a few carefully selected points of the domain. This careful

choice of points (also known as the training set) is made resorting to DoE techniques [6]. True SBO

consists on applying traditional optimization techniques to surrogate models, and iteratively refining and

updating those models with new information until ending criteria are met. Another option would consist

on performing just one single iteration of the process, i.e., choosing a training set, building the surrogate

2

and finding its optimum. Yet another possibility, detailed in [7], would be to build our approximation model

itself based on approximated simulations.

In the end, the final algorithm implemented could best be described as a heavily surrogate-dependent

steepest descent method, with an objective function evaluation based on uncertainty propagation tech-

niques. It models the aircraft’s aerodynamic behaviour using surrogate models, whilst at the same time

using a custom real-time gradient-based optimization algorithm that is compatible with the operational

procedures of a real flight test. It has an extensive and complex implementation, the devising of which

constitutes the main goal of the present work. This is thoroughly detailed in Chapter 6. Additionally, SBO

techniques were used to validate, compare with, and complement the main gradient-based approach.

1.4 Thesis outline

This thesis is structured in 8 main chapters, and a brief summary of their contents reads as follows:

Chapter 1 : Introduction

Where the context and motivation behind this project are explained, the problem is stated and a

solution approach is proposed.

Chapter 2 : Aircraft Takeoff Optimization

Presents the concepts and basic ideas behind the TO optimization process, which is at the core of

the solutions later developed.

Chapter 3 : Surrogate Modelling

Exposes the different surrogate modelling techniques used throughout this work, detailing their

limitations and strong points, as well as underlying principles.

Chapter 4 : Design of Experiment

Where different Design of Experiment techniques are presented and discussed, since they are

extensively used in the algorithm implemented.

Chapter 5 : Uncertainty Propagation

Defining a statistically optimal configuration later in the project required using state-of-the-art un-

certainty propagation techniques. These are briefly synthesized here.

Chapter 6 : Optimal Configuration Search

The heart of the subject, extensively describing the algorithm implemented, from both a functional

and modular point of view.

Chapter 7 : Results

Presentation, analysis and validation of the results obtained.

Chapter 8 : Conclusion

Summarizes all work done, with emphasis on key ideas and insights obtained as well as on future

work proposals.

3

Chapter 2

Aircraft Takeoff Optimization

2.1 Overview

In the present project, we are essentially concerned with aircraft TO optimization. The aim of this section

is then to explain all the different elements that are involved in a traditional TO optimization process.

Only a few aircraft performance terms will be explicitly defined here. For a comprehensive list of all

maximum/minimum speeds, runways, weights and other performance elements, please refer to [8].

The TO optimization traditional objective is to obtain the highest possible performance-limited takeoff

weight, whilst at the same time fulfilling all airworthiness requirements. In order to do that, it is neces-

sary to first determine what parameters influence TO performance, and then determine which of those

parameters can/cannot be controlled (free/sustained parameters respectively). Table 2.1 synthesizes

this.

Sustained parameters Free parameters

RunwayClearway Takeoff configurationStopwayElevation

Slope Air conditioningObstacles

TemperaturePressure V1

WindRunway condition

Anti-ice V2Aircraft status (MEL/CDL)

Table 2.1: TO Parameters Categorization

In the following sections an analysis of each of the free parameters is carried out, since those are

the only ones controllable and therefore relevant to the optimization process.

4

2.2 Takeoff Configurations

Traditionally, TO can be accomplished using one of three TO configurations: Conf 1+F, Conf 2 or Conf

3. Each of these configurations is associated with a set of certified performances, and as a result

it is always possible to determine a Maximum TakeOff Weight (MTOW) for each TO configuration. The

optimum/best configuration from among the set of 3 is the one that allows for the highest MTOW. Again,

it should be emphasized that traditionally optimum configuration is spoken about in the sense of best

choice from among the available configurations. In the present project, this definition of optimality

will acquire a different meaning.

For longer runways, a better climb gradient is searched, whereas for shorter runways a shorter

TakeOff Distance (TOD) is wanted. The Conf 1+F, having the best ’finesse’, is thus better suited for long

runways. The Conf 3 sacrifices ’finesse’ for brute lift, and as such is suited for shorter runways. The

Conf 2 represents a natural compromise between climb and runway performance, and may sometimes

be the optimum choice for TO.

2.3 Air conditioning

Air conditioning, when switched on during takeoff, decreases the available power and thus degrades the

takeoff performance. It is then advisable to switch it off during TO, but this is not always possible as

some constraints exist (high air temperature in the cabin or/and company policy), unless Auxiliary Power

Unit (APU) bleed is used.

2.4 Takeoff Speed Optimization

TO speeds play a major role in achieving the maximum TakeOff Weight (TOW) gain. The following

subsections describe how this gain is obtained via speed ratio optimization (V1/VR and V2/VS).

2.4.1 Speed Ratios

Before identifying and defining the speed ratios used in the optimization process, a quick recap of the

definitions of the speeds involved is in due order [9]:

V1 Maximum speed at which the crew can decide to reject the takeoff, and is ensured to stop the aircraft

within the limits of the runway.

VR Rotation speed, the speed at which the pilot initiates the rotation, at the appropriate rate of about

3◦ per second.

V2 Minimum climb speed that must be reached at a height of 35 feet above the runway surface, in case

of an engine failure.

5

VS Conventional stall speed, the speed at which the lift suddenly collapses. At that moment, the load

factor is always less than one.

The first ratio used in a TO optimization is the V1/VR. The rationale behind its usage is detailed here.

The decision speed V1 must always be less than the rotation speed VR, which in turn depends on weight.

As such, the maximum value of V1 is not fixed, but the maximum V1/VR ratio is fixed and is equal to one

(regulatory value). On the minimum side, aircraft manufacturers have shown [8] that V1 speeds less

than 84% of the VR render TO distances too long, and don’t therefore present any TO performance

advantages. The minimum V1/VR ratio is then equal to 0.84 (manufacturer value). This is the reason

why the V1/VR ratio is used as optimization variable, since its range is well-identified:

0.84 ≤ V1VR≤ 1 (2.1)

It should be noted that any change in the V1/VR ratio will qualitatively have the same effect on TO

performance as a corresponding change in the V1 speed.

The second speed ratio used in the TO Optimization process is the V2/VS ratio. The minimum V2

speed is defined by regulations (its value differs depending on the aircraft, see [9]), and VS depends on

weight. In analogy with before, while the minimum V2 is not a fixed value, the minimum ratio is fixed a

priori (for a given aircraft type). And also like before, if the V2 speed is too high, it will lead to long TO

distances and thus poor TO performance. Depending on the aircraft, the manufacturer will place a limit

on the V2/VS ratio. The range of this ratio is now well-defined:

(V2VS

)min

≤ V2VS≤(V2VS

)max

(2.2)

Like with V1/VR, any change in the V2/VS ratio will affect TO performance in the way a similar V2

change would.

2.4.2 V1/VR Ratio Influence

The influence of V1/VR variations on the TO optimization process will now be discussed, considering a

fixed value of V2/VS . A number of other parameters will also be considered fixed, namely:

For a description of all performance related terminology used hereafter, namely abbreviations, please

refer to the Abbreviations section at the beginning of this document.

Looking at the runway limitations, an increase in V1/VR leads to (where the Accelerate-Stop Distance

is denoted by ASD):

• An increase in MTOW limited by TODN−1 and TORN−1,

• A decrease in MTOW limited by ASDNorN−1,

• Not influencing the MTOW limited by TODN and TORN

Let us now consider the climb and obstacle limitations. For the former, the V1 speed has no influence

on climb gradients (first, second and final takeoff segments). For the latter, an obstacle-limited weight

6

Fixed parameters

ElevationRunway data Runway

ClearwayStopway

SlopeObstacles

QNHOutside conditions Outside Air Temperature

Wind component

Flaps/SlatsAircraft data Air conditioning

Anti-ice

Table 2.2: Fixed parameters for V1/VR study

is improved with a higher V1, as TO distance is reduced. Therefore, the start of the takeoff flight path is

obtained at a shorter distance, requiring a lower gradient to clear the obstacles.

For these limitations, an increase in V1/VR leads to:

• An increase in MTOW limited by obstacles,

• Not influencing the MTOW limited by the first, second and final TO segments.

Finally, looking into the brake energy and tire speed limitations, we quickly conclude that a V1/VR

increase leads to:

• A decrease in MTOW limited by brake energy,

• Not influencing the MTOW limited by the tire speed.

Figure 2.1 represents all the aforementioned limitations put together. Clearly an optimum MTOW is

always achievable, most often at the intersection of two limitation curves. The result of this optimization

process is, for a given V2/VS ratio, an optimum MTOW and an associated optimum V1/VR ratio.

2.4.3 V2/VS Ratio Influence

For a given V1/VR ratio, the influence of the V2/VS ratio on the TO optimization process will now be

detailed.

Considering the runway limitations, and as a general rule, for a given V1/VR ratio, any increase in

the V2/VS ratio leads to an increase in the one-engine-out and the all-engine TO distances. Intuitively,

in order to achieve a higher V2 at 35 feet, more energy needs to be acquired on the runway, leading to a

longer acceleration phase. It would seem that the V2 speed has no direct impact on the ASD. However,

a higher V2 results in a higer VR and, therefore, for a given V1/VR ratio, in a higher V1 speed. Hence, the

effect on ASD. A V2/VS increases leads thus to:

• A decrease in MTOW limited by TODN−1, TODN , TORN−1, TORN , ASDN−1 and ASDN .

7

Figure 2.1: V1/VR MTOW limitations.

Source: [8]

For climb and obstacle limitations, increasing V2/VS provides better climb gradients and consequently

a better climb-limited MTOW (first and second segments, and obstacles). Since the final TO segment is

flown at green dot speed, it is not affected by a change in V2. Taking into account these limitations, an

increase in V2/VS will lead to:

• An increase in MTOW limited by the first and second segments, as well as by any obstacles,

• Not influencing the MTOW limited by the final TO segment.

Lastly a look into brake energy and tire speed limitations is due. The V2 indirectly affects the impact

brake energy limitation, since a V2 increase implies a VR increase and therefore a V1 increase (at fixed

V1/VR). Regarding the tire speed limitation, since the VLOF is limited by the tire speed, it limits the V2 to

a maximum value. An increase in V2/VS can then be considered equivalent to a VS reduction (V2 being

fixed), and thus the tire speed-limited MTOW is reduced. In essence, a V2/VS increase will lead to:

• A decrease in MTOW limited by brake energy and tire speed.

2.5 Optimization Process & Results

As discussed in sections 2.4.2 and 2.4.3, given a V2/VS ratio, an optimal MTOW and corresponding

V1/VR ratio can be found. To carry out the optimization all that is needed is to perform this MTOW and

V1/VR optimum determination for each value of V2/VS in the range given by equation 2.2. In the end,

the results of this optimization process are an optimal MTOW and both optimal V1/VR and V2/VS ratios.

8

Once the optimal speed ratios and MTOW are obtained, using the Aircraft Flight Manual (AFM) and

the aforementioned MTOW the VS can be obtained, which yields in turn V2 from the optimal speed ratios.

With this V2 and referring to the AFM the VR can be derived, which will immediately yield V1 via the speed

ratios.

9

Chapter 3

Surrogate Modelling

3.1 Overview

Surrogate modelling [10] refers to the construction of approximations (surrogates) that fit and explain

user-provided data (training set). In this section, a general terminology concerning surrogate modelling is

first presented. Then, a theoretical overview of the techniques used throughout this project is explained,

as well as the strengths and weaknesses of each technique. This section closes with considerations on

the accuracy evaluation of the different techniques used.

Consider our data, a training set S, to be a collection of vectors representing an unknown response

function f :

Y = f(X) (3.1)

where:

• X is a dinput-dimensional vector

• Y is a doutput-dimensional vector

A single element of the training set is denoted by (Xk, Yk), where Yk = f(Xk). We represent the number

of elements in the training set, i.e. its size, by |S|. When referring to the input parts of the training set,

{Xk}|S|k=1, the term Design of Experiment (DoE) will be used. Numerically speaking, S is represented by

|S|, a (dinput × doutput) matrix, henceforth denoted (XY )training. This matrix (XY )training is naturally

divided into two submatrices, Xtraining and Ytraining, corresponding to the DoE and output components

of S respectively.

Given a training set S, the goal of surrogate modelling is to construct a function f ,

f : Rdinput −→ Rdoutput , (3.2)

that approximates our unknown response function f (Eq. 3.1). To achieve this, a range of different

techniques may be employed.

10

3.2 Techniques

For each technique used throughout this project, its principles are herein presented. A brief summary of

the technique’s strengths and weaknesses follows each description.

3.2.1 Linear regression

In implementing a Linear Regression (LR) approximation, we start by assuming that our training set was

generated by the following linear model:

Y = Xα+ ε (3.3)

where:

• α is a dinput-dimensional vector containing the unknown model parameters

• ε ∈ R|S| is a vector generated by a white noise process

The coefficients of α may be estimated using, for instance, a Ridge Regression Estimate (RRE) [11],

as follows:

α = (XTX + λI)−1XTY (3.4)

where:

• I ∈ Rdinput×dinput ,

• λ ≥ 0 is estimated by leave-one-out cross-validation.

Once α is estimated, output prediction for an input X is given by:

Y = Xα (3.5)

A linear regression is a very crude and basic model, albeit a highly universal and simple one. It is fast

to create, even for large sample sizes and problem dimensionality, and is practically insensitive to noise.

In some cases, particularly if the size of the training sample is comparable to the input dimension of the

problem, a LR may suffice (to analyze our response function sensitivity with respect to the independent

variables, for example). It is worth noting that a LR model cannot normally be significantly improved be

adding new training data.

11

3.2.2 Response Surface Model

Historically one of the most popular surrogate techniques [12], a Response Surface Model (RSM) gen-

eralizes the LR technique presented before. The RSM model is defined as follows:

f(X) = α0 +

dinput∑i=1

αi xi +

dinput∑i,j=1

βij xi xj (3.6)

where:

• X ∈ Rdinput

• αi, βij are the unknown model parameters

There are several types of RSM (see table 3.1 below), depending on what coefficients in equation

3.6 we set to zero[6]:

RSM Type Short description

Linear No second-degree terms. All βij = 0.

Linear w/interactions Besides linear terms, the products of pairsof distinct variables are considered. βij = 0, ∀i = j.

Quadratic no interactions Quadratic model that ignores variableinteractions, i.e., βij = 0, ∀i 6= j.

Quadratic Model containing all terms.

Table 3.1: Different RSM types

The RSM model 3.6 can be written as f(X) = φ(X) c, with c = (α, β) being our vector of unknown

model parameters. Let Ψ = ψ(X) be our design matrix. A number of different ways exist for estimating

our coefficients c, such as:

Least Squares [13]

c = (ΨTΨ)−1ΨTY

Ridge Regression (see 3.2.1)

c = (ΨTΨ + λI)−1ΨTY , same as for the LR method.

Multiple ridge [14]

c = (ΨTΨ + Λ)−1ΨTY , where Λ = diag(λ1, λ2, ..., λdinput) is a diagonal matrix of regularization

parameters, estimated sucessively by cross-validation.

Numerous other techniques such as [15] exist, but these suffice for the present work. For a more

exhaustive and complete treatment of the subject please see [16].

The RSM being a generalization of the LR, shares its traits of robustness and unsensivity to noise,

as well as high construction speed. It can handle large training sets and high dimensionality easily. The

number of regression terms used, however, increases rapidly with an increasing number of dimensions.

RSM’s drawbacks are also similar to the LR’s, albeit slightly attenuated: it is still a crude approximation,

and adding more samples to the model will normally not improve its accuracy.

12

3.2.3 Gaussian Processes

Gaussian Process (GP) based modelling, also known as Kriging, is another very popular surrogate

method that has already been widely documented (see [17, 18], for example). In essence, it is a spatial

optimal linear prediction, where the unknown random-process mean is estimated with a linear unbiased

estimator. A GP is thus fully determined by its mean function m(X) = E[f(X)] and covariance function

cov(f(X), f(X ′)) = k(X,X ′) = E[(f(X)−m(X))(f(X ′)−m(X ′))].

Firstly, it is assumed that the training data set S = (X,Y ) was generated by a GP f(X):

Yi = Y (Xi) = f(Xi) + εi, , i = 1, 2, ..., |S| (3.7)

Where εi is a Gaussian White Noise (GWN) with zero mean and variance σ2. A zero mean function

m(X) = E[f(X)] = 0 is also assumed, as well as a covariance function k(X,X ′) belonging to a para-

metric class of covariance functions k(X,X ′|a) (with a being a vector of unknown parameters). For the

present project, two classes of covariance functions are considered:

Squared exponential covariance function [19]:

k(X,X ′|a) = σ2 exp

− dinput∑i=1

θ2i (xi − x′i)s, s ∈ [1, 2] (3.8)

Where a = {σ, θ, i = 1, ..., dinput} is the vector of parameters.

Mahalanobis covariance function [20]:

k(X,X ′|a) = σ2 exp (−(X −X ′)TA (X −X ′)) (3.9)

Where A ∈ Rdinput×dinput is a positive definite matrix and a = {σ,A}.

Under these assumptions, the training set S is modeled by a GP with the following covariance func-

tion:

cov(Y (X), Y (X ′)) = k(X,X ′) + σ2δ(X −X ′) (3.10)

Where δ(X) is a delta function. Thus, the mean value of the process for a test point X∗ is given by

(incorporating our training data):

f(X∗) = k∗(K + σ2 I)−1Y (3.11)

where:

• I ∈ R|S|×|S| is an identity matrix,

• k∗ = k(X∗,X) = [k(X∗, Xj), j = 1, ..., N ], which implies

k(X,X) = [k(Xi, Xj), i, j = 1, ..., N ]

This mean value is the value used for prediction. To measure the accuracy of the prediction at any given

13

point, the covariance function based on the training set can be used:

V[f(X∗)] = k(X∗, X∗) + σ2 − k∗(K + σ2I)−1(k∗)T (3.12)

The sole missing piece now is the unknown vector of parameters a of the covariance function.

The values of a are estimated based on the training sample by maximizing the logarithm of corre-

sponding likelihood [21]:

maxa,σ

log p(Y |X, a, σ) = −1

2Y T (K + σ2I)−1Y − 1

2log |K + σ2I| − |S|

2log 2π, (3.13)

with |K + σ2I| being the determinant of K + σ2I.

GP is a surrogate method that demonstrates very accurate behaviour, provided |S| is of small/moderate

size. This is a method perfectly suited for modelling spatially homogeneous functions, i.e., functions with-

out discontinuities, as well as high dimensionality problems. It is however a resource-intensive method

in terms of memory capacity, and thus does not cope well with large training sets.

3.2.4 Splines in tension

Splines in tension is a shape-preserving spline method for approximation of 1-D functions (dinput = 1

and doutput ≥ 1). A tension parameter σi, i = 1, ..., n will be associated to each abscissa interval

[Xi, Xi+1] of our function. Varying this tension parameter from zero to infinity will alter the fitting curve

from a cubic polynomial to a linear function. A method, presented in [22], can be used to select tension

factors in a way such that concavity and monotonicity are preserved, producing a smooth curve that

avoids oscillations in case of discontinuity in the underlying function. Only the one-dimensional case

output is relevant for the present project, and thus only this particular case will be presented here (the

generalized algorithm can be found in [23]).

For our one-dimensional interpolation problem, consider a sequence of values of abscissas X1 <

X2... < X|S|, and the corresponding function values Yi, i = 1, ..., |S|. The interpolation problem is then

to find the function f(X):

f(Xi) = Yi, i = 1, ..., |S|, (3.14)

f ∈ Cm[X1, X|S|], m = 1 or 2

The generalized definition of a 1-D interpolating tension spline is the following [23]:

f(X) = arg ming(X)

∫ X|S|

X1

[g′′(X)]2dX +

|S|−1∑k=1

σ2k

∆X2k+1,k

∫ Xk+1

Xk

[g′(X)]2dX]

, (3.15)

where:

• g(Xi) = Yi

• ∆Xk+1,k = Xk+1 −Xk

14

• σi is the tension parameter in the interval [Xi, Xi+1], i = 1, ..., |S| − 1.

The tension spline procedure used chooses the minimum tension factors that satisfy constraints

related to the smoothness of the derivatives. When the σi are fixed, the solution of 3.15 is known. For

convenience let us write an explicit expression for f(X), focusing only on the interval [X1, X2]. For this,

let Y1, Y2 and Y ′1 , Y′2 denote the data values and derivatives, respectively, associated with X1, X2, and

define:

h = X2 −X1, b =X2 −X1

h, s =

Y2 − Y1h

, d1 = s− Y ′1 , d2 = Y ′2 − s.

A convenient set of basis functions for the interpolant is obtained from the modified hyperbolic functions:

sinhm(Z) = sinh(Z)− Z and coshm(Z) = cosh(Z)− 1

Further defining:

E = σ · sinh(σ)− 2 coshm(σ)

α1 = σ · coshm(σ) d2 − sinhm(σ)(d1 + d2)

α2 = σ · sinh(σ) d2 − coshm(σ)(d1 + d2)

Thus, for σ > 0, the interpolant is given by:

f(X) = Y2 − Y ′2 h b+h

σE[α1coshm(σb)− α2sinhm(σb)] (3.16)

which, when σ = 0, simplifies to:

f(X) = Y2 − h[Y ′2b+ (d1 − 2d2)b2 + (d2 − d1)b3] (3.17)

For a detailed description of the σ estimation heuristic see [22]. With σ known, evaluation of our predic-

tion at X is trivial using 3.16.

This SPLines in Tension (SPLT) technique is computationally cheap, and as such can be used with

extremely big training sets. Being a combination of linear and cubic splines, it offers both good ro-

bustness and smoothness, being an interpolating technique. It is, however, inadequate for very noisy

problems, and is restricted to 1-D input models (dinput = 1).

15

Chapter 4

Design of Experiment

4.1 Overview

Simply put, DoE [6] is a strategy of experimentation that maximizes learning using a minimum of re-

sources. DoE strategies are of prime importance when each individual experiment run is costly and/or

time-consuming. In this section, some general DoE terminology and concepts are first presented, fol-

lowed by some possible examples of general DoE techniques currently in use. The section closes with

an exposition of optimal DoE techniques applied to RSM models, something that is central to this project.

Let us start by defining some terminology:

Design variables Parameters or quantities to be varied during the experiment. Here each design vari-

able shall be noted xk, an element of the design variable vector x, with k = 1, ..., d, and d being

the number of design variables.

Design space The d-dimensional space defined by the lower and upper bounds of each design vari-

able.

Design vector/Design point A concrete instance of x, where all values in the vector x fall within the

bounds of the design space.

Response Concrete measure or evaluation on a specific design point.

DoE A subset of the design space, denoted X = {xi}Ni=1, where N is the size of the DoE, i.e. number

of design vectors in it.

A DoE technique is then a procedure for choosing efficient DoE in the design space, with the goal of

maximizing information gained from a limited number of design vectors and their responses.

To measure the quality of a given DoE, one of the most important properties that should be consid-

ered is the uniformity of said DoE in the design space. A number of metrics exist in the literature that

attempt to quantify and measure this uniformity. Some of them are presented below. Let our design

space be the normalized hypercube [0, 1]d.

16

Discrepancy [24]

Denoting by Pu,v the d-dimensional box

Pu,v =

d⊗k=1

[uk, vk[,

where 0 ≤ uk < vk < 1 ∀k = 1, ..., d, discrepancy of our set X is defined as:

D(X) = sup0≤uk<vk<1

∣∣∣∣#(X ∩ Pu,v)N

− |Pu,v|∣∣∣∣ , (4.1)

where:

1. #(·) denotes the number of points,

2. |Pu,v| =∏dk=1(vk − uk) is the volume of Pu,v.

Intuitively it can be said that the discrepancy attempts to measure ’how well’ do the points fill the

design space at ’all’ scales.

Minimax Interpoint Distance ρ

A self-explanatory metric, defined as

ρ(X) = maxi

minj:j 6=i

‖xi − xj‖, (4.2)

with ‖ · ‖ being the euclidean norm in Rd. It simply looks at the ’worst-case’ in the DoE, that is,

what is the point of the set for which the longest distance must be bridged to reach its nearest

neighbour.

φ Metric [25]

Defined for p ≥ 1 as

φp(X) =

N∑i<j

‖xi − xj‖−p 1

p

. (4.3)

This is a sort of ’potential energy’-type measure to characterise a DoE.

Evidently, it should be noted that a more uniform DoE corresponds to lower values of the aforementioned

metrics.

4.2 General DoE Techniques

A short description of some currently available DoE techniques will now be presented, coupled with

their strenghts and weaknesses. This is but a short preview of the vast literature on the subject. For a

presentation of more advanced, sequential DoE techniques, please refer to [26, 27, 28].

Random sampling (RS)

RS is arguably the simplest approach possible, consisting in the uniform generation of random

17

(a) Non-optimal LHS design. (b) Optimal LHS design

Figure 4.1: OLHS iterates over various LHS designs to maximize uniformity.

Source: [33]

points in a hypercube. It is thus a low uniformity technique, especially in low dimensions. However,

its a very flexible and universal approach, meaning an existing design can be easily improved by

adding more points.

Latin Hypercube Sampling (LHS)

A very popular and widespread technique [29, 30, 31], LHS is a technique based on the preser-

vation of the uniformity of marginal distributions. It is done by dividing the range of each design

variable into N equal intervals, and placing one point in each. Statistically speaking, a multidi-

mensional grid containing sample positions is a Latin hypercube if (and only if) there is only one

sample in each such interval. N sample points are placed to satisfy the Latin hypercube require-

ments. This forces the number of intervals, N , to be the same for each variable.

One of the main advantages of this LHS scheme is that it does not require more samples for more

dimensions, the existing samples just ’accommodate’ themselves in higher dimensions. It also

has good low-dimensional projection properties, and is extremely fast to generate. Its drawbacks

are mainly twofold: It has a non-zero probability of filling the design space unevenly (although this

probability declines rapidly with increasing N ), and it is not sequential, that is, not extensible by

adding new points without breaking LHS properties.

Optimized Latin Hypercube Sampling (OLHS)

OLHS [32] is an improved version of LHS. One of the drawbacks of LHS is the sometimes uneven

filling of the design space. OLHS corrects this (Figs.4.1a, 4.1b), by iteratively generating LHS

designs and choosing the best one according to one or more uniformity metrics (see section 4.1).

This OLHS is thus more reliable than its counterpart LHS, while keeping the same good low-

dimensional projection properties. However, the optimization process can be quite slow, and the

resulting DoE still cannot be easily extended by adding new points without breaking LHS proper-

ties.

18

Full Factorial (FF)

The FF approach can be viewed as a ”brute-force” DoE, that is, it generates a DoE consisting in

all possible combinations of design variables. For discrete design variables, this is straightforward,

whilst for continuous design variables we need to select a set l of discrete levels - by partitioning

the design variable ranges. Consider the case of d design variables with l levels each. The DoE

will have ld points, rapidly rendering the FF computationally impossible for problems of significant

size.

FF DoE design’s main strengths are its very fast generation and extremely low value of uniformity

metrics (section 4.1). However, the two drawbacks directly related to those advantages are the

very fast growth of ld with d for any reasonable l and the fixed number of points restriction (they

must number exactly ld).

4.3 Optimal DoE Techniques

The purpose of this section is to describe a specific class of DoE techniques, named optimal DoE

designs, applied to Response Surface Models (RSM). An optimal design is constructed by optimizing

some criterion that results in minimizing either the generalized variance of the parameter estimates, or

the variance of the prediction, for a pre-specified RSM model structure (see Sec. 3.2.2). Then, given

a budget of experimental runs, the optimization procedure chooses the optimal set of design points

from a candidate set of possible design points. Optimal designs seek not only to place points in the

design space uniformly, but also to achieve more robust and consistent estimates of our (RSM) model

parameters or its predictions, effectively using a very limited number of points (experiment runs) in the

process.

Chapter 3.2.2 of the present work already treats the subject of RSM models, and as such it will not

be repeated here. Notation used there will, however, be used in what follows. Additionally, let:

• D denote the design space,

• X = (xi : xi ∈ D)Ni=1 be the design matrix,

• ψ(X), as mentioned before, refer to a full RSM design.

A vast number of optimality criteria exist and have been extensively studied (see, for example, [34]).

Here, focus is put into two widespread [35, 36] optimality criteria :

D-optimality

Provides a design that minimizes the determinant of the design inverse covariance matrix, i.e.

finds a design X solution to the problem:

arg minX

det[(ψ(X)Tψ(X))

−1]] (4.4)

This approach allows for a minimizing of the variance of the parameter’s estimates. ”D” stands for

Determinant optimality.

19

IV-optimality

A DoE that minimizes the integrated variance of the prediction throughout the design space, i.e.

finds a DoE X solution of:

arg minX

∫Dψ(x)(ψ(X)Tψ(X))−1ψ(x)Tdx (4.5)

”IV” standing for Integrated Variance optimality.

Solving optimization problems 4.4 and 4.5 can be done using the algorithm detailed in [37]. In essence,

it consists of 3 main steps:

1. Generate a full factorial (FF, see Sec. 4.2) design, called the candidate set for optimal design.

2. Take the required (pre-defined) number of points from the candidate set at random and calculate

the optimality criterion for these points.

3. Iterate by inserting/removing points originating from the candidate set until criterion convergence.

Provided the RSM-like assumption for the system being analysed holds, as well as the specific

RSM type assumed (see Table 3.1 in Sec. 3.2.2), optimal DoE techniques achieve the same level of

accuracy as general techniques with much less experiment points. In the field of aircraft performance,

especially when it comes to flight testing where each design point corresponds to a very expensive and

time-consuming flight of an aircraft, this advantage is of paramount importance.

20

Chapter 5

Uncertainty Propagation

5.1 Overview

Uncertainty Propagation (UP) refers to the study of the effect uncertainty in the input of system has

on its output. At present, it is widely used [38] in the field of Robust Design Optimization (RDO). In

context of this project, UP techniques will be employed in a way that differs from the traditional RDO

framework [39], as UP will be applied in the definition of custom objective functions. This different

approach will be described in Sections 6.3.3 and 7.5. The aim of this section is to provide the essential

background needed for comprehension of the UP techniques used later, as well as detail their strengths

and weaknesses.

Simply put, consider a function f(X), with X ∈ Rn. For continuous variables, let the mean µf and

variance σ2f of f be defined as:

µf = E[f(X)] =

∫ +∞

−∞f(X)pX(X) dX (5.1)

σ2f = E[(f(X)− µf )2] =

∫ +∞

−∞(f(X)− µf )2pX(X) dX (5.2)

A closed-form solution for Equations 5.1 and 5.2 is not available, in general, for problems of practical

interest. The numerical approximation of these statistical moments is the subject of UP, and numerous

techniques exist for carrying this out, offering different compromises between cost and accuracy. A brief

review of these methods will be presented next.

5.2 Techniques

In the present section, let µX , σX , γX ,ΓX represent the mean, standard deviation, skewness and kur-

tosis respectively of the statistical distribution assumed as input X ∈ Rn of f(X). A more detailed

description of the techniques subsequently presented can be found in [40, 41].

21

5.2.1 Taylor Moment Propagation

Provided f(X), our system response, is differentiable a sufficient number of times regarding X (the un-

certain input), µf and σ2f can be approximated by a Taylor expansion around µX . This method is known

as Method of Moments (MM), and can be applied using differently-truncated expansions depending on

the precision/performance desired. Expressions for the expansions of µf and σ2f as well as a thorough

mathematical review of this method is availble at [42, 43].

This is a popular non-linear method, offering a freedom in the choice of the order of the expansion.

A first order MM can yield practical results at a low computational cost [44]. That being said, if f is

non-linear, the accuracy of a first-order MM in terms of µf can be compromised. Retaining higher-order

terms fixes this issue, at a higher computational cost. It is worth noting, however, that in order to improve

accuracy for the σ2f estimate (compared to a 1st order MM) a minimum of a 3rd order MM should be

considered.

5.2.2 Gaussian Quadrature

Gaussian Quadrature (GQ) is method of calculation of the integral of a function (f ) by a properly

weighted sum of particular values f(Xi), i = 1, 2, ..., N where Xi are N selected points in the func-

tion domain. The one-dimensional case is straightforward. For a multivariate integration, the following

generalizing formulas are used:

µf =

N∑i1

Wi1

{N∑i2

Wi2

[· · ·

N∑in

Winf(Xi1,i2,...,in)

]}(5.3)

σ2f =

N∑i1

Wi1

{N∑i2

Wi2

[· · ·

N∑in

Win[f(Xi1,i2,...,in)− µf ]2

]}(5.4)

Where n denotes the dimension of the input and each dimension k has N sampling points corre-

sponding to a weight set Wik. A detailed description of this method is done in [45].

A disadvantage of this method is the generally high number of evaluations of f for multidimensional

integration. That being said, sophisticated techniques have been successfully developed that attempt

to counter this, such as [46]. In low dimensions, GQ are very cost-effective, knowing at present a

widespread use.

5.2.3 Monte Carlo Methods

Monte Carlo (MC) methods, originating from the statistical interpretation of integrals in Equations 5.1

and 5.2, are another option for UP. Probability distributions over the outputs of a process induced by the

probability distributions of over the inputs are obtained by performing m repetions of said process, each

of which corresponds to a sampling point Xi ∈ Input Space. Considering a random sampling, unbiased

estimators for integrals 5.1 and 5.2 are given by:

22

µf =1

m

m∑i=1

f(Xi) (5.5)

σ2f =

1

m− 1

m∑i=1

[f(Xi)− µf ]2 (5.6)

Both 5.5 and 5.6 converge with a normalized error magnitude of O(m−12 ), and thus MC methods

normally require a significant number of evaluations of f . This drawback is compensated by the ease of

implementation of a MC method.

5.2.4 Univariate Reduced Quadrature

The Univariate Reduced Quadrature (URQ) [41] is a quadrature method inspired on Sigma-Point (SP)

methods [47], which aims at obtaining a cheap and accurate univariate integration method for a generic,

non-symmetric distribution.

Using this method, the following expressions for mean and variance of our function f are obtained

[40]:

µf = W0f(µX) +

n∑p=1

Wp

[f(X+

p )

h+p−f(X−p )

h−p

](5.7)

σ2f =

n∑p=1

{W+p

[f(X+

p )− f(µX)

h+p

]2+W+

p

[f(X−p )− f(µX)

h−p

]2

+W±p[f(X+

p )− f(µX)][f(X−p )− f(µX)]

h+p h−p

} (5.8)

where:

• X±p are the sampling points, defined as: X±p = µX + h±p σXpIp,

• Ip is the p-th vector of the identity matrix of size n,

• h±p are given by h±p =γXp

2 ±√

ΓXp−

3γ2Xp

4 .

The weights W must be chosen as follows:

• W0 = 1 +∑np=1

1h+p h−p

,

• Wp = 1h+p −h−p

,

• W+p =

(h+p )2−h+

p h−p −1

(h+p −h−p )2

,

• W−p =(h−p )2−h+

p h−p −1

(h+p −h−p )2

,

• W±p = 2(h+

p −h−p )2.

This URQ method requires 2n + 1 evaluations of f . It has thus a cost similar to that of a lineariza-

tion method (i.e. a MM method where the first-order derivatives are determined by finite differences).

23

However, the accuracy of the URQ is much higher than that of a linear method [40], making it a cheap

and relatively accurate method, suited for application in the context of this flight performance project. It

provides deterministic estimates of µf and σ2f , suitable for use with deterministic optimization algorithms

(such as the one first developed in the present work). It is worth noting that the extra accuracy of the

URQ derives from its use of all first four moments from the input distribution, whereas a MM method

would only use the first two. It is then a key requirement for this method to work that the aforementioned

moments be available.

24

Chapter 6

Optimal Configuration Search

The goal of the present Chapter is to expose in detail the algorithm (henceforth called CoH, for ConfOpt

Hunter) implemented for Optimal Configuration Search, as well as the hypothesis and architectural op-

tions that were made during development. The Chapter’s structure is as follows. Section 6.1 presents

the different Airbus in-house engineering tools that were used to help create CoH. Section 6.2 sum-

marizes CoH’s structure, via a high-level overview. Section 6.3 closes the Chapter with an exhaustive

description of the implementation details on each different CoH module.

6.1 Core Tools

6.1.1 OCTOPUS

OCTOPUS (Operational and Certified TakeOff and landing Performance Universal Software) is a family

of low speed performance tools developed in-house by Airbus. It provides the following main functional-

ities:

• Aircraft Flight Manual (AFM) - A performance calculation kernel,

• TO weight optimization,

• Landing weight optimization,

• Operational Flight Paths (OFP).

It is a comprehensive low-speed performance tool. In the context of this project, however, focus shall

be placed on the TO weight optimization functionality.

Before starting a TO optimization, we need to select an Aircraft Definition File (ADF) to model aircraft

physics. Then, after entering a group of user-defined settings and input conditions (such as temperature

data, pressure altitude, runway information, ...) OCTOPUS performs a TO optimization (see Section

2) and calculates the MTOW for the conditions specified. This single calculation, done for one of the

pre-defined TO configurations, is at the heart of the objective function definition in the CoH optimizer.

25

Internally, OCTOPUS represents the aircraft’s characteristics by a group of Labels, each one being

essentially a set of data. These labels can be accessed and altered by OPTIMA, described next in

Section 6.1.2.

For the present project, only Labels concerning the Lift Coefficient CZ , Drag Coefficient CX and

Maximum Lift Coefficient CZMAXare of interest. Each coefficient is computed in the following fashion:

Lift Coefficient CZ (or CL)

See Confidential Appendix, Data 1. (6.1)

Drag Coefficient CX

See Confidential Appendix, Data 1. (6.2)

Maximum Lift Coefficient CZMAX(or CLMAX

)

See Confidential Appendix, Data 1. (6.3)

From amongst the panoply of terms involved in Eqs. 6.1 to 6.3, only the following are relevant for the

present study (and their respective labels, indicated between [· · · ]):

CL free air [ACZALPHA] Lift coefficient in free air and reference conditions.

CD polar [ACXCZ2] Drag coefficient issued from balanced symmetrical drag polar in free air and ref-

erence conditions.

CLMAXBASIC [ACZMAX] Maximum lift coefficient at reference CG with gear up, including Mach num-

ber effect.

It is by altering (also known as overloading) via OPTIMA the aforementioned labels, and nullifying all

other terms that the three coefficients CZ , CX and CZMAXare manipulated, effectively allowing complete

control over the aircraft’s performance.

6.1.2 OPTIMA

OPTIMA is a project currently in development at Airbus that aims to provide next-generation aircraft

performance modelling and analysis capabilities. From amongst its already available capabilities, the

relevant one for the project at hand is the Python wrapping it provides for OCTOPUS. A Python wrapping

means that OCTOPUS’ computational functionalities can be called and accessed via a Python script.

This is of paramount importance for embedding large performance evaluation black-boxes into larger

computational workflows as well as one of the reasons coding of the CoH algorithm was done in Python.

Other reasons for preferring Python have to do with its versatility, the ample array of 3rd-party scientific

library support and scripting nature of the language itself which facilitates quick prototyping.

26

6.1.3 MACROS

MACROS, developed by the Airbus Group in partnership with DATADVANCE, is a set of software tools

for process integration, predictive modeling, data mining and multidisciplinary optimization. In the scope

of this project, MACROS was used namely in three different areas: DoE, Surrogate Modelling and

Optimization.

Its tools are accessed via Python scripting, thus facilitating integration with OPTIMA and, by exten-

sion, with OCTOPUS.

6.2 Overall Architecture

The final CoH algorithm draws inspiration from the steepest gradient descent method. A brief justification

for this choice is given on Section 6.2.1, before delving into the description of the CoH on Section 6.2.2.

6.2.1 Gradient Descent Rationale

Gradient optimization methods are a standard, widely used and well documented family of optimization

algorithms [48, 49]. The steepest gradient descent is a first-order optimization algorithm, whose basic

principle is given below.

Consider f(X) ∈ C1 to be our multivariate objective function to be minimized, and note that, for a

small enough step γ:

b = a− γ∇f(a) =⇒ f(b) < f(a), (6.4)

for a, b ∈ Df .

Starting with an initial guess X0, the algorithm builds the sequence X0, X1, X2, ..., Xn such that:

Xk+1 = Xk − γk∇f(Xk), k ≥ 0 (6.5)

If the function f is well-behaved [50], local convergence can be guaranteed.

Now, there are several reasons as to why the steepest descent was chosen as a starting point for

the CoH algorithm development, namely:

1. Simplicity of implementation. Being one of simplest methods available, custom tailoring of the

algorithm is made easier. It should be noted that the final CoH differs quite considerably from the

standard algorithm described above.

2. Expected smoothness and good behaviour of the objective functions. Discontinuous or abrupt

performance changes from one configuration to the next are neither common nor expected.

3. Expected accuracy of the initial guess. When optimizing a flight configuration, the initial guess is

based on pre-flight models as well as on the expertise and know-how of performance engineers. It

is thus normally a reasonable guess, which means a more complex heuristic full space-searching

algorithm is not necessary.

27

4. Confirmation/certification nature of the search. Should it happen that the initial guess is accurate

enough, the simple observation of a nearly null gradient is enough to tell that the optimum was

attained, and save on the flight tests.

6.2.2 Algorithm Description

What follows is an overview of one iteration of the CoH algorithm (Fig. 6.1). The steps are presented in

order, starting with an initial guess and up to the stop criteria, where it is decided if another iteration is

needed. For each step only the main idea is explained. Implementation and hypothesis description is

done in Sec. 6.3.

Design Space Description

An optimal TO configuration is defined as a set of control surface deflection angles that maximize some

optimality criteria. Traditionally, this set of surfaces comprises only the Slat and Flap angles, for two

reasons:

• The Slat and Flap deflection explain most of the TO performance, as they play a major aerodynamic

role (Slats increase CZMAX, Flaps increase CZ0

);

• Adding more parameters exponentially increases the number of tests to be done, thus increas-

ing the costs. Examples of such additional surfaces could be the Aileron angles, or making the

distinction between Inboard and Outboard Flaps.

As such, and for the remainder of this Chapter, let us consider a two-dimensional Design Space (DS)

of Slat and Flap angles, defined as:

DS = {(s, f) ∈ R2 : s ∈ [0◦, 27◦], f ∈ [0◦, 35◦]} (6.6)

Where the limitations placed on the Slat and Flap angles (s and f respectively), are geometrical

ones. See Fig. 6.2.

It should be noted that this choice of DS implies absolutely no loss of generality for the process

implemented, i.e., all the tools developed were conceived to work with and be easily extensible to any

given number of design variables.

Initial guess

Before proceeding, the definition of a Flight Point (FP), in the context of flight tests, will be introduced.

As the name suggests, a FP corresponds to a set of speed measurements taken while the aircraft is

airborne at a fixed configuration (or point).

Until now, within Airbus, the optimal configuration was determined in a non-systematic fashion, that

is, by performing a set of flight tests that covered a pre-defined list of FP around the pre-flight config-

28

Figure 6.1: CoH algorithm basic workflow.

uration. The data was then analyzed and a choice was made in terms of which should be the optimal

configuration.

That being said, for the CoH, relatively precise initial guesses can be made for example by analogy

with previous aircraft models, by resorting to wind tunnel testing or CFD pre-flight computations, by

looking at the optimal configurations that were defined for them, or even by asking for the opinion of

Airbus experts.

Initially, for a straightforward MTOW objective function, the chosen initial guess will depend mainly on

29

Figure 6.2: Flap/Slat design space.The optimum configuration is represented by a light yellow star.

the runway length available. This single factor, the runway, is of paramount importance when it comes to

deciding between a ’cleaner’ configuration (higher L/D ratio) or a ’high-lift’ one, as explained in Section

2.2. Later, when defining statistically optimal objectives, providing an initial guess is normally not an

easy task.

Optimal DoE around guess

Having picked an initial guess in the DS, the next step is to calculate the local gradient, to determine

in which direction to continue the exploration. However, in order to calculate a gradient, a local linear

approximation of the objective function must be built in the vicinity of the DS point currently being consid-

ered. To build this approximation, a sampling of points is needed (for the 2-dimensional DS, a minimum

of 3 points, see Fig. 6.3).

To perform that sampling (which is actually a DoE), in the most cost-effective manner possible a linear

RSM-optimal DoE technique was used. Described thoroughly in Section 4.3, this is a DoE procedure

that maximizes information gain at a minimum budget cost, provided the linear RSM assumption holds.

Not only that, applying such a DoE ensures scalability (in case more FP need to be added for

robustness) and generality (in the event more design variables are considered).

First breakpoint: DoE flight test results

This is the most central and significant step in the CoH algorithm. The reason for that is simple: to

evaluate the objective function at any given FP, several flight tests must be conducted (at least at 3

different speeds). At this stage, algorithm execution must stop (hence the designation ’breakpoint’)

and wait for the inflight measurements. The real-time nature of the CoH is due to these necessary

breakpoints, since no a priori knowledge of the next FP locations (besides the initial group) is assumed.

30

Figure 6.3: DoE around initial guess.

Obviously, in order to continue development of the CoH, a solution had to be adopted. A simulator

for flight testing was then conceived, using surrogate modelling techniques to model the aircraft’s flight

physics. A very important word of clarification is due: at this stage of an aircraft performance certification

campaign, when the Optimal Configuration (OC) is not yet fixed, only pre-flight calculations can be made

using OCTOPUS, since no flight tests have been carried out. This means all the aircraft physics comes

from theoretical models. At the time of development of the present project, however, the A350 OC

campaign had already been terminated, meaning flight test (FT) data was available. It was from this

A350 OC campaign FT data that the aforementioned flight test simulator was built, in an attempt to

emulate reality and truly validate the algorithm being developed. Granted, the applicability of the CoH

algorithm for the A350 is limited at best. Nevertheless, it is a potent way to demonstrate the applicability

of the method for future aircraft OC determination.

The FT results for a FP are a number of pairs of aerodynamic coefficients CZ and CX , measured for

different angles of attack α (i.e., different speeds).

Gradient Calculation

At this stage, using OPTIMA (Sec.6.1.2), FT results are injected into OCTOPUS (Sec.6.1.1), and a TO

optimization (Sec.2) is performed at each FP (Fig.6.4).

Figure 6.4: FT results are used for a MTOW TO optimization.

31

OCTOPUS then outputs a simple MTOW for a given custom configuration. This is not a default

functionality of OCTOPUS. Indeed, all performance calculations in OCTOPUS are limited to using one

of the three default available Takeoff configurations. Using OPTIMA, however, it is possible to overload

the flight physics inside OCTOPUS, namely the drag polar and lift curves, and effectively ’turn’ a default

configuration into a custom one (see Fig.6.5). This overloading is done using surrogate models (Sec.3),

with the help of MACROS (Sec.6.1.3). It is explained in detail in Section 6.3.3.

Figure 6.5: Configuration overloading using OPTIMA.

Initially, the objective function to be maximized was precisely the MTOW, for a given runway length.

Later, a statistical objective function was developed, which takes into account the worldwide airport

runway length distribution to optimize the TO configuration. In any case, the output value of the objective

function shall henceforth be referred to as objective.

Having calculated the objective for the selected DoE points around the initial guess, a LR surface

(Sec.3.2.1) is built to approximate the objective function. From this surface, the gradient is analytically

calculated.

Second breakpoint: Gradient line points

With gradient information now available, the direction of increasing objective value can be determined.

However, unlike in a traditional steepest ascent/descent algorithm (Sec.6.2.1), no step is taken in that

direction. Instead, the design space is sampled at N locations over the direction of increasing objective

defined by the gradient (Fig.6.6).

It is worth noting that an accurate gradient estimation greatly facilitates this approach, as sampling

will be done in a really ’relevant’ gradient direction.

This design choice was made mainly for operational reasons, as a step-by-step real-time in-flight

measurements would be impractical. Performing measurements by chunks avails itself as a more prac-

tical and economical approach.

A very important and challenging aspect not yet mentioned is the handling of optima located in the

borders of the domain. This is often the case in the present project, as will be shown later. A number

of measures were put into place to handle gradient calculation and gradient line sampling in the frontier

of the DS, to counter convergence problems (namely slow convergence, which is simply not affordable

when each objective calculation corresponds to a real FP measurement.

A mathematical and exhaustive description of the implementation is done in Sec.6.3.

32

Figure 6.6: Sampling along gradient line.

Maximum over gradient line

Once the set of sampling points to be evaluated over the gradient line is defined, FT can be performed in

each of those FP. Again, in the context of this project, this implies resorting once more to the previously

mentioned flight test simulator.

After obtaining the FT results, objective values are calculated for each FP in the set, using the objec-

tive function under consideration. Interpolation using a spline interpolation (SPLT) technique (Sec.3.2.4)

is then carried out. It is possible to build such a spline since the gradient line is a one-dimensional

manifold regardless of the input dimensionality. That being said, a parametrization scheme was devised

to achieve this.

The SPLT technique yields an analytical expression for the interpolant spline, making it trivial to find

its optimum. Once the point of maximum objective is calculated, it becomes the initial FP for the next

iteration (Fig.6.7).

Flight testing being a heavily budget-constrained process, the stopping criteria for the CoH algorithm

will necessarily be the exhaustion of said budget. Thus, once the pre-allotted flight time or number of FP

evaluations run out, termination of the CoH ensues. In case there is some budget left, another iteration

begins. However, a convergence criteria based on the gradient norm could also be put in place to spare

as much as possible the budget in case the initial guess is good. It should be noted that the CoH can

stop in the midst of an iteration.

33

Figure 6.7: Maximum over gradient line becomes next guess.

34

6.3 Implementation

The division proposed for the CoH algorithm in Section 6.2.2 is the most suited for an explanation of the

idea behind the algorithm. It is not, however, a very practical division when it comes to its implemen-

tation. Although care was taken to identify the two breakpoints in the CoH flow (the two points where

the algorithm waits for the FT results), FT simulations and objective calculations are still done twice per

iteration. This redundancy prompted a different functional division for the coding of CoH. Seven modules

were created, each encapsulating a different functionality. It not only eliminates code redundancy, but

also facilitates the creation of an interface for the FT team. The modules and their correspondence with

the CoH algorithm steps detailed previously are presented in Table 6.1.

Algorithm step Modules usedOptimal DoE around guess DoE Generator

First breakpoint: DoE FT results Flight Test Simulator

Gradient calculation Objective CalculatorGradient Calculator

Second breakpoint: gradient line results Line DesignerFlight Test Simulator

Maximum over gradient lineObjective Calculator

Line BuilderLine Optimizer

Table 6.1: CoH steps - module correspondence.

These modules interact in the fashion depicted on Fig. 6.8. There is no direct data sharing between

two modules. Instead, all information pertaining to the current execution of the CoH is kept stored and

up-to-date on a separate entity. Information flow in CoH must be managed by the high-level script that

is using the modules.

6.3.1 DoE Generator

The purpose of this module is to generate a DoE around a given point of the design space DS. This

point point will be called the design point P = (p1, p2, ..., pn) ∈ DS, with n being the dimensionality of

DS. Given this goal, the procedure followed to achieve it is described next.

First, a rigorous definition of the zone ’around’ P is needed, since this will be the domain for the DoE.

For that purpose, a parameter K1 ∈ [0, 1] was introduced. Let:

• Uk denote the upper bound for the k-th DS dimension,

• lk denote the lower bound for the k-th DS dimension.

The DoE domain (henceforth denoted DD) is then defined as follows:

DD = {(x1, x2, ..., xn) ∈ Rn : ∀k ∈ [1, 2, ..., n], mk ≤ xk ≤Mk}, (6.7)

35

Figure 6.8: CoH implementation modular workflow.

where:

• pk denotes the k-th component of the current FP P .

• mk = max [lk, pk −K1 · (Uk − lk)],

• Mk = min [Uk, pk +K1 · (Uk − lk)].

This definition of the DD is simple and scalable, also handling well design near DS borders.

Once the DD is set, a RSM-optimal DoE (see Section 4.3) is done via MACROS, using either a D or

IV optimality criteria. Depending on the value of K1, a linear or quadratic RSM (see Table 3.1) is used.

For smaller K1, that is, for a smaller hyperbox around P , a linear RSM is a more accurate aproximation

of the local response function shape. Conversely, for bigger K1, a quadratic RSM will perform better (in

which case more FP are necessary).

An important entry parameter for the DoE Generator is the number of points of the DoE, np. For any

meaningful RSM surface to be built, np ≥ n + 1 must be enforced. A higher value of np will necessary

lead to a more robust DoE, at the expense of longer flight testing. An analysis of the influence of np on

CoH performance is to be found in Section 7.3.

The final product of the DoE Generator is thus a set of np points ∈ DD that maximize information

gain, assuming the surface in the vicinity of P is either quadratic or linear.

6.3.2 FT Simulator

The FT Simulator is an essential module for the CoH, even though it is only meant to be a temporary

replacement for real flight testing. It embodies the greatest hypothesis made in the development of this

algorithm, which is: results from the FT Simulator are meant to emulate real FT results. To achieve this,

as mentioned in Section 6.2.2, a complex surrogate model (SM) was built using FT data from the A350

Optimal Configuration campaign.

36

From the FT databanks, only the data represented on Table 6.2 was used for SM construction.

Inputs Outputs

Slats Flaps Ailerons Alpha Engine CG CZ CX

......

......

......

......

Table 6.2: FT source data

Following terminology adopted in Section 3.1, the Input data columns stacked together make up the

Xtraining matrix, and the two output columns constitute the Ytraining matrix.

Regarding the inputs, two additional hypothesis were made:

• The Flap values are taken as the average of the Inboard and Outboard Flap values;

• No distinction was made between left or right engine. Most FT are carried out in a 1-Engine

Inoperative setting, and thus the operating engine during a flight may vary.

A GP SM technique was used used to build the FT model. Simultaneously, RSM and LR SM were

also built, albeit for different reasons. They were used only for an easier sensitivity analysis of the

response to a given input.

The FT SM being a f : R6 7→ R2 function, lower dimension hyperplane cuts were employed for

visualization and model analysis. When representing these hyper-cuts, two issues must be addressed:

• Representation of the data points: All data points are represented by their projections on the

dimension being displayed, meaning some points appearing to be near the SM in the hyper-cut

may be far in reality;

• Values of the non-represented dimensions, i.e., where to cut: When displaying a hyper-cut, the

input dimensions not being represented must be provided with a value. This will be called the ISO

value, and will be indicated for each cut. E.g., a median ISO-value means all non-represented

inputs are kept at the median values of their respective data point values, an average/max/min

ISO-value at their average/max/min values, etc.

Without further ado, the FT model is presented next.

The adequacy of the FT model can be ascertained by looking at Fig. 6.9, where the data points

(in dark blue) are overlapped by the GP surface. The increase in CZ0 with flap deflection is accurately

represented, as well as the linear lift curves CZ vs α.

In terms of drag coefficient CX , Fig. 6.10 once again showcases the accurate fitting of the GP model

with the data. An increasing flap deflection predictably raises drag, and considering the base FT were

performed in the linear CZ vs α zone, far from the stall zone, the quasi-linear CX vs α relation was

expected. A closer look into the outputs CZ and CX is obtained via 2D cuts (Fig. 6.11), where two

additional LR and RSM surfaces SM were built to complement analysis.

Figs. 6.11a and 6.11b represent 2D cuts with other parameters set at their minimum and maximum

values, respectively. The proper model fitting by the GP, RSM and LR SM can be clearly seen in both

37

Figure 6.9: CZ as function of the angle of attack α and Flap deflection. Cut at ISO median.

Figure 6.10: CX as function of α and Flaps. Cut at ISO median.

situations. An analogous observation can be made concerning the flap 2D cuts illustrated in Fig. 6.12,

where the RSM and LR surfaces play a qualitative role. That is, they confirm the expected output trends

of increasing drag and lift with flap deflection, independently of α and other parameters. Note that all

the data points are represented (via their projections) in Fig. 6.12. Also, the fact that there are not many

38

(a) CZ and CX versus α 2D cuts at ISO minimum. (b) CZ and CX versus α 2D cuts at ISO maximum.

Figure 6.11: Two dimensional visualization of the α input parameter influence on output, using 3 differentSM: LR, RSM and GP.

points that match the curves has to do with the way flight tests were performed, meaning that during

FT emphasis was put into varying α, not the configuration, for example. Still, the curves do interpolate

perfectly the few corresponding points.

When it comes to Slat input of the FT Simulator, these are not expected to influence CZ , as the Slats

only increase the critical α (delay the stall). CX , however, is expected to increase with increasing Slat

deflection. Both observations are confirmed in Figs. 6.14 and 6.13. It is by this worsening of finesse

(Lift-to-drag-ratio) that high slat deflections are detrimental to long runway TO performance.

The following three entry parameters: Ailerons, Center of Gravity (CG) position and Engine thrust

are presented next, in Figures 6.15, 6.16 and 6.17 respectively.

See Confidential Appendix, Data 2.

6.3.3 Objective Calculator

Definition of the most relevant Objective Function (OF) for OC optimization was one of the goals of the

present project. This implied a progressive approach to the problem was needed, starting with simple

solutions.

The simplest possible OF corresponds to calculating the MTOW for a given Runway Length (RL),

then maximizing said MTOW. This MTOW calculation process is illustrated in Figure 6.18.

The entry parameters for the calculation are:

39

(a) CZ , CX versus Flaps 2D cuts at ISO minimum. (b) CZ , CX versus Flaps 2D cuts at ISO maximum.

Figure 6.12: Two dimensional visualization of the Flaps input parameter influence on output, using 3different SM: LR, RSM and GP.

Figure 6.13: CX as function of α and Slat deflection. Cut at ISO max.

FT Results Results from flight testing performed at the given FP and RL. These will be used to emulate

the aircraft’s lift and polar curves, namely the ACZALPHA and ACXCZ2 labels.

Runway Length Runway length, a key parameter in TO optimization intimately tied to TO configura-

tions, as described in Section 2.2.

Flight Point Current point (that is, configuration) being considered, belonging to the DS. It is used for

ACZMAX calculation, since for other labels configuration information is implicity contained in the

40

Figure 6.14: CZ and CX as function of Slat deflection. Cut at ISO max.

FT Results.

Outside Conditions Self-explanatory, refers to the myriad of sustained parameters (other than RL,

such as Temperature, Pressure, etc) needing explicit definition (see Section 2.1).

Other settings Other CoH, OPTIMA and OCTOPUS settings.

In the label builder phase, SM are built from the FT results, using LR or RSM surfaces. The ACZAL-

PHA curve is built from the FT input α and FT output CZ , whereas the ACXCZ2 curve is built from two

FT outputs, CZ and CX . This means the FT model isn’t able to produce ACXCZ2 curves, but could

theoretically be directly used for ACZALPHA emulation. Note that even though traditionally these labels

explicitly contained information on the configuration, that is no longer the case. Instead, configuration

information is implicitly contained in the FT results, with the created labels reflecting precisely that.

Having obtained the label models, these are passed on to OCTOSUP via OPTIMA, along with an

ACZMAX model loaded from memory. This model was built using a GP SM based on existing pre-flight

CZMAXdata in OCTOSUP (see Figs. 6.19, 6.20a and 6.20b).

The reason for using pre-flight data to overload the ACZMAX label is simple: the FT Data used to

build the FT Simulator contain no stall information. This represents thus another important hypothesis

made for the CoH, which is: ’The CZMAXemployed from pre-flight is assumed to be the ’real’ CZMAX

of

41

Figure 6.15: CZ and CX as function of Aileron deflection. Cut at ISO mean.

the aircraft’.

The output of the MTOW calculation is given by OCTOSUP, once the TO optimization is finished.

One problem in using the MTOW as OF has to do with the choice of RL. For a given RL, the OC found

will be optimal for that and only that RL value. Traditionally, aircraft manufacturers (namely AIRBUS)

solved this problem by introducing 3 TO configurations, 1+F, 2, and 3, for long, medium and short

runways respectively. One immediate solution is thus to set 3 different RL values then determine the 3

configurations in the traditional sense. This is demonstrated in Section 7.4. Yet the question of which 3

42

Figure 6.16: CZ and CX as function of CG (Center of Gravity) position. Cut at ISO mean.

RL to pick remains. At this stage, engineering experts can come up with educated guesses, wrapping

up the issue for this Objective Function (OF).

However, this OF question begs to be pushed further. Using the OF previously described, the final

aircraft will have 3 TO configurations optimized for 3 different RLs: R1 < R2 < R3. For simplicity, let us

assume R1 = 2000m, R2 = 3000m and R3 = 4000m. Consider now these two scenarios:

• In real life usage, the aircraft will rarely TO from runways measuring exactly R1,R2 or R3 meters.

If, as an example, the aircraft has to serve three airports with 3 RLs R1 = 2500m, R2 = 3500m

43

Figure 6.17: CZ and CX as function of the Engine setting. Cut at ISO mean.

and R3 = 4500m, all its configurations will be non-optimal.

• Suppose the aircraft is to serve 20 airports with RL 4000m and 100 airports with RL 4200m. In

this case, the OC for R1 should really have been defined at a RL closer to 4200m.

These scenarios hint at the need for a more general approach to the OF definition problem. A pos-

sible answer unexpectedly arises from Uncertainty Propagation (UP) methods (see 5), and is detailed

next.

Consider the RL of a given airport served by our aircraft to be a continuous random variable X of

44

Figure 6.18: MTOW calculator architecture.

Figure 6.19: ACZMAX GP SM label CZMAXVS Flap and Slat deflection 3D hyper-cut, at ISO minimum.

unknown distribution, representing the probability of, in case an airport is picked randomly, its RL is near

X. In that case, the MTOW for the hypothetical airport in question will also be a RV, of unknown and

possibly unusual distribution. This distribution is of interest, but before it can be used two questions need

45

(a) CZMAX VS Mach at ISO minimum. (b) CZMAX VS Mach at ISO maximum.

Figure 6.20: Two dimensional visualization of the Mach parameter influence on the ACZMAX label.

to be addressed:

1. How to economically and accurately find the MTOW Distribution (MDist)? It should be reminded

that a TO optimization takes time, and long simulations (using numerous RL values) are unpractical

(read expensive) for a real-time procedure like the CoH.

2. How to exploit this MDist?

Answer to 1 is given by the UP method described in Section 5.2.4. Using only three RL values,

accurate µMDist and σMDist values can be obtained.

Answer to 2 calls for a brief ponderation. The only information about MDist cheaply obtainable are the

µMDist and σMDist moments. Given that nothing else is known about MDist, maximizing or minimizing

σMDist would not be of use. As a measure of dispersion, a bigger σMDist would guarantee an occasional

greater MTOW at the expense of punctual smaller MTOWs. Since aircraft are sold with strict minimum

TOW guarantees for different airports, this maximization could constitute a problem. A minimization

would bring no significant benefits either. Maximizing µMDist, however, is both a simple and relevant

target. It corresponds to finding a OC that would, in average, yield the greatest MTOW at any given

airport. This is the criteria that defined the final choice of OF for the CoH.

Figure 6.21 illustrates the final Objective Calculator architecture that was put into place.

The entries to the module are analogous to those depicted in Fig. 6.18, with the exception of the

Runway range. This is an interval of the form RR = [RLmin, RLmax]. Only airports where RL∈ RR are

considered for uncertainty propagation. It is the input for control of the OF.

In step 1 of UP, the 4 statistical moments µ, σ, γ,Γ (mean, standard deviation, skewness and kurtosis)

of the MDist are computed, taking care to only consider a subset of airport limited by RR. It should be

noted that the Airport Runway Data used to compute these statistical moments comprises a list of only

the airports operated by the aircraft, their RL, as well as the number of flights operated there by time-

period. As a result, the weight or importance attributed to a given runway value X for the distribution is

given by:

Value of X =

ak∑ai=a1

Number of flights operated at ai (6.8)

46

Figure 6.21: Objective Calculator architecture.

Where {a1, a2, ..., ak} is a list of airports with runway length near X. A visualization of the distribution

as well as a histogram are present in Fig.6.22. The distribution was estimated using a Gaussian Kernel

smoothing technique.

Figure 6.22: RL probability density and histogram.

Having this information, the 3 RL values where the TO optimization is to be conducted are calculated

next, using the MTOW calculator described previously.

47

Once all MTOWs are calculated, step 2 of UP deals with the calculation of µMDist and σMDist. For

all practical purposes, µMDist is the output of the Objective Calculator module.

Before closing the module description, a physical interpretation of this statistical objective is due. It

can be said that a statistically optimal configuration for the runway range RR yields the biggest MTOW,

in average, throughout aircraft operation in the airports falling inside the RR range.

6.3.4 Gradient Calculator

In this module, an approximation for the gradient of the OF is calculated at the DoE center P.

First, a LR or RSM-quadratic SM is built using the set of objective values calculated beforehand.

Note that the choice between LR or RSM will depend on the optimal DoE RSM assumption made

before, which itself depends in turn on the chosen value for K1 (Section 6.3.1).

Once the SM surface is constructed, the gradient is trivially determined analytically. For example, for

the LR case (which is the most frequently used case) it follows (Section 3.2.1, LR Model):

Y = Xα+ ε =⇒ ∇Y = α, (6.9)

where α are the coefficients of the LR model.

Now, cases where gradient calculation takes place near the DS border must be carefully considered

and treated. In order to do so, a masking procedure was created, and is described below. Let:

• g1∈ [0, 1] be the near-border parameter. This defines a margin for P to be considered ’near the

border’ of the DS domain. More precisely, P is considered to be bordering the high limit of the k-th

(k ∈ n) dimension of the DS if

|pk − Uk| < g1

(Uk − lk

2

), (6.10)

where Uk, lk, pk are defined as in Section 6.3.1. Similarly, P borders the lower limit of the DS k-th

dimension if

|pk − lk| < g1

(Uk − lk

2

). (6.11)

• P r1,r2,...,roq1,q2,...,qo′denote that P is near the higher limit of DS dimensions r1, r2, ..., ro and near the lower

limit of DS dimensions q1, q2, ..., qo′ , with rs 6= qs ∀s ∈ n (P never near two borders of the domain at

the same time).

• Fk be the gradient flux through the DS k-th dimension at point P. It is defined as Fk = ∇yk|P , with

yk being the k-th component of vector Y .

With these definitions in mind, a masked gradient is defined as ∇Ym = ∇Y · m, where m =

[a1a2 · · · an], ak ∈ {0, 1}. The mask m is calculated as follows:

1. PRQ is calculated, with R = {r1, r2, ...} and Q = {q1, q2, ...} being the sets of dimensions P is

bordering;

2. For each d ∈ (R ∪Q), the d-th component md of the mask is set to zero iff:

48

• d ∈ R and Fd > 0, or

• d ∈ Q and Fd < 0.

If this is not the case, md is set to 1.

This masking of the gradient is necessary for the Line Design process done along the gradient line

as explained in Section 6.3.5.

Finally, the output of this module is then a masked normalized gradient at point P, or ∇Ym(X)

‖∇Ym(X)‖

∣∣∣X=P

.

6.3.5 Line Designer

The Line Designer module aims to produce a set of N points GS = {G1, G2, ..., GN} ∈ DS, with all

the points in GS being along the line of increasing OF values. This line, henceforth referred to as

gradient line (GL), covers the direction defined by the masked gradient of the OF calculated at P, i.e.,∇Ym(X)

‖∇Ym(X)‖

∣∣∣X=P

= ∇G.

The GL can thus be parametrized by a single variable z ∈ R, independently of the DS dimensionality

n:

GL(z) = P + z · ∇G, (6.12)

where z ∈ [ζmin, ζmax], and ζmin, ζmax are a function of P,∇G and the DS boundaries lk, Uk (as defined

in Section 6.3.1).

Calculation of ζmin, ζmax yields the parametrization limits, or the minimum and maximum values ∇G

applied at P can be multiplied by until the DS limits are breached. To carry out this calculation, the

following algorithm was developped:

1. Set ζmax = 0;

2. Set zM = maxk(Uk−lk)‖∇G‖ , X = P + z · ∇G. This ensures X will be a point outside DS, provided

P ∈ DS;

3. Using notation from Section 6.3.4, calculate XRQ . If R and Q are not empty sets, make z = z

2 . Else,

make P = X and ζmax = ζmax + z;

4. If the distance from P to the closest DS border is smaller than the precision required, end ζmax

calculation. Else, return to 1;

5. For ζmin steps 1-3 are analogous, except for X = P +z ·∇ in step 1, which becomes X = P −z ·∇.

This algorithm converges exponentially fast.

Once the ζ are known, sampling of N points along GL is a trivial matter of selecting N values in the

range R = [ζmin, ζmax], as the corresponding points can be calculated via Eq. 6.12. The sampling

method implemented in the CoH is simple: Pick ζmin, ζmax, as well as N-2 randomly-selected points in

R, making the set gs = {ζmin, v1, ..., vN−2, ζmax}.

The output GS of this module is then GS = {G1, G2, ..., GN}, where Gk = GL(gsk) and gsk ∈ gs.

49

6.3.6 Line Builder

The Line Builder is a simple module whose purpose is to produce a SM of the OF (Objective Function)

with its domain restricted to the GL (Gradient Line). This makes the restricted OF (hereby denoted L(x))

a 1-dimensional manifold, with parameterizing variable z ∈ [ζmin, ζmax] as defined in Section 6.3.5.

For construction of the SM, L(x) : R 7−→ R, FT and OF value calculations are carried out at each

point from the Line Designer output set GS. A spline L(x) is built using the SPLT technique described in

Section 3.2.4, with the help of MACROS.

The output of the module is the SPLT SM L(x), represented internally by an array of tension coeffi-

cients.

6.3.7 Line Optimizer

Taking as input the L(x) SM built by the Line Builder, the Line Optimizer’s goal is to find the optimum

objective over L, i.e., X∗ = arg maxX L(X).

L being an analytical SM, optimization is performed near-instantaneously using a Quasi-Newton

descent optimization tool provided by MACROS.

Problems often arise near domain boundaries in optimization problems. For the CoH problem, this

issue arises often. As such, the semi-random sampling method implemented ensures that convergence

is not slowed down when optimizing different L over different iterations of the CoH on a boundary k of

the DS, if the flux Fk is outward-bound. If it is inward-bound, no issue arises since line design will be

done into the domain.

The output of this module is of course X∗.

50

Chapter 7

Results

7.1 Overview

The purpose of this Section is to present and validate the results obtained from the CoH algorithm.

Validation of the results is of paramount importance for an expensive real-time procedure such as the

one developed here. Validation comprises two distinct phases, each accompanied by a set of results:

1. Validation of the core performance calculations, that is, a simple MTOW calculation. Being the

underlying foundation of the CoH, any divergences found here would impact all the subsequent

results.

2. Validation of the process itself, that is, the CoH algorithm. For this, results from CoH executions

are compared to optimum values determined via an ’unlimited-budget’ surrogate modelling. This

consists in using the FT Simulator coupled with the Objective Calculator to exhaustively sample

the DS, a sampling from which a very accurate SM is built. It is over this ’unlimited-budget’ SM

model that optimization is then performed. Optimum obtained this way are considered the true

optimum, against which the CoH output can be meaningfully compared.

Regarding what follows, Section 7.2 addresses the aforementioned phase 1, whilst Sections 7.3-7.5

all tackle phase 2.

7.2 Single MTOW, Fixed Runway

The validation and results analysis process starts with simple MTOW calculations, amounting to testing

of a subpart of the Objective Calculator module. In this Section, 4000m is adopted as RL value, and

other sustained parameters are defined according to the FT data used to build the FT Simulator. As

explained earlier in the present work, no CZMAXFT data was available for the FT Simulator, hence the

4000m RL, assuring the segment under consideration is not limited by CZMAX.

This first MTOW calculation is done at a 1+F Configuration, defined in pre-flight as (Confidential)

Slat deflection, (Confidential) Flap deflection, using the pre-flight model. Next, still at (Confidential),

51

calculations are launched using LR and RSM interpolation techniques for the labels, with a varying

number of interpolation points and full overload of all label models. Results are displayed in Fig. 7.1.

(a) MTOW as function of interpolation technique.

(b) Interpolation error.

Figure 7.1: RSM and LR interpolation results.

See Confidential Appendix, Data 3.

Thus, the question that truly needs to be answered is: which interpolation method better represents

the FT data?

To attempt to answer this, the Lift and Drag polars used by OCTOSUP during the TO optimization

loop for all three cases (pre-flight, LR, RSM) were extracted and visualized (Fig. 7.2). Keep in mind this

was done using 4 different speed measures for each FP.

Bearing in mind that only the continuous, more dense parts of the curves are of importance, two

comments can be made regarding Fig. 7.2:

See Confidential Appendix, Data 4.

A look at the absolute modelling errors involved (Fig. 7.3) yields an important insight into MTOW

calculation sensibility versus label modelling errors.

See Confidential Appendix, Data 5.

This is an aspect of the utmost importance in the CoH algorithm: a proper sampling along the alphas

during FP measurement is vital for result accuracy. A RSM interpolation is adopted in the Sections that

follow.

52

Figure 7.2: Lift and drag (CZ and CX ) curve plots.

Figure 7.3: Lift and drag (CZ and CX ) error analysis.

53

7.3 Optimal Configuration, Fixed Runway

In the present Section, CoH algorithm results are compared against the so-called ’true’ results. These

true results are obtained from optimization over the DS using a SM, which in turn is built for a fixed RL.

This is known as Surrogate-Based Optimization (SBO). The SM takes as input a configuration, and it

outputs the corresponding MTOW. Keep in mind a sampling (training) set of the DS is needed to build

the SM.

At this stage two factors need now to be considered: the sample size S and DoE Technique. Ideally,

for purposes of comparison with CoH results, a SM where S → ∞ with a Full Factorial DoE would be

best, as it would be the closest to reality. However, it is of major interest to determine if a comparable

accuracy could be obtained with a significantly smaller S or not. A small enough S the generates an

accurate SM would qualify this Surrogate-based Optimization process as a worthy alternative to the

CoH.

The S∞ SM, henceforth considered as the reference, was built using a S = 225 FF DoE. From it,

the true OC and respective MTOW for a RL of 4000m were derived, and are both depicted in Figure 7.4.

More precisely, the OC found is indicated in Table 7.1.

Slats –Flaps –MTOW –

Table 7.1: Optimal Configuration for RL 4000m

See Confidential Appendix, Data 6.

A similar process was then carried out using smaller S values and different DoE techniques, taking

care to always sample the entire DS. The results obtained are shown in Figures 7.5 and 7.6.

See Confidential Appendix, Data 7.

54

(a) 3D view.

(b) Top view of the surrogate.

Figure 7.4: MTOW as function of configuration, using GP SM based on FF limit design. Note the OC,indicated by a yellow star.

55

Figure 7.5: MTOW value progression for the full DS, depending on DoE sample size S and technique.

Figure 7.6: Absolute MTOW error for the full DS, depending on DoE sample size S and technique.

56

Figure 7.7: MTOW value progression for a small sampling domain, depending on DoE sample size Sand technique.

Figure 7.8: Absolute MTOW error for a small sampling domain, depending on DoE sample size S andtechnique.

57

Even for a small (Confidential) sample, extremely precise results are obtained, regardless of the DoE

technique used. This fact, coupled with the Slat and Flap configuration convergence depicted in Figs.

7.9 and 7.10 respectively, position small-DS SBO as a valid tool for confirmation purposes. If a very

precise initial guess is already known (from a same-family aircraft, for example) confirmation of the OC

location could be quickly and cheaply carried out this way.

Figure 7.9: Slat deflection for a small sampling domain, depending on DoE sample size S and technique.

58

Figure 7.10: Flap deflection for a small sampling domain, depending on DoE sample size S and tech-nique.

59

At this stage, having obtained the ’true’ OC for reference (Table 7.1), all that is left to determine is the

correct parameterization for the CoH algorithm, namely:

• The minimum number of iterations needed until a satisfactory (small enough) error is obtained;

• Optimal number of Flight Points (FP) to use around each Design Point (DP) for gradient calculation;

• Optimal number of FP to use in the creation of the gradient line.

To accomplish this, a DoE OLHS set was first created using these 3 parameters (NumRuns, DoeSize,

LineSize). Additionally, a set of 10 random start-points (SP) was also created. Then, for each SP and

for each configuration in OLHS set, the CoH algorithm was launched, with the average over the SP set

of the results obtained being used to build a SM.

Figure 7.11 shows the MTOW (objective) evolution versus the number of iterations and DoE size.

Convergence to the true OC MTOW value with the increasing number of runs was expected, and is

clearly observable. Yet, the influence of the DoE size used in gradient calculation is very small, with

bigger gradient DoEs appearing to slightly slow convergence (mayhap due to gradient ’pollution’).

Figure 7.11: 3D GP SM cut at ISO-max of CoH output as function of local DoE pool sample size andnumber of iterations.

60

Similar 3D cuts for the number of points along the gradient line (LineSize) are depicted in Figs. 7.12a

and 7.12b. Besides the convergence with increasing number of iterations, the line size shows this time

a significant influence on convergence speed. The RSM model in Fig. 7.12b underlines this point, better

capturing the trend. Using only (Confidential) points to build the gradient line seems to be something to

avoid.

Concerning convergence speed, that is, how many runs are really needed to attain the true ’OC’, a

2D cut provides a good first answer (see Fig. 7.13). After just (Confidential) iterations, a very precise

estimate is already obtained, improving marginally from there onwards.

61

(a) GP 3D cut.

(b) RSM 3D cut.

Figure 7.12: 3D cuts at ISO-average of CoH output as function of gradient line sample size and numberof iterations.

62

Figure 7.13: 2D cuts showing MTOW convergence with increasing number of iterations.

63

Now, in order to quantify CoH performance, two curves are of paramount importance, and they are

described next:

Error Vs FP Budget

Consider the FP Budget of an entire CoH execution to be defined as

FP Budget = (DoE Size + Nb Grad Line Points)× Nb of Iterations + 1, (7.1)

where the last FP to be evaluated is the last optimum over the gradient line. To obtain a curve

relating the error obtained (compared to the ’true’ OC) with the total number of FP used, the

following optimization problem is posed:

minimized,g,n

FP Budget(d, g, n)

subject to |MTOW (d, g, n)−MTOW ∗| ≤ E, E = e1, . . . , em.

(7.2)

Where d, g are the number of FP used in each DoE and Gradient Line, respectively, and n is the

number of iterations. The problem is solved for different values of admissible error E, namely

e1, . . . , em. The parametric model constructed before is used to carry out the optimization. Recall

that the model was built using results averaged over 10 different random starting points.

Error Vs Number of Runs This second curve is a by-product of the optimization process previously

described, illustrating the effect of the error allowed on the number of CoH iterationS;

Note that curves describing the evolution of d and g also arise naturally from the optimization process.

Everything is depicted in Figures 7.14a and 7.14b.

Figure 7.14a shows results for CoH executions with a large error allowed. Possible conclusions are:

See Confidential Appendix, Data 8.

Analysis of Fig. 7.14b, traced for smaller error allowances, yields greater insights on the aforemen-

tioned results. A number of observations can be made:

See Confidential Appendix, Data 9.

This being said, from a business perspective the fundamental question is the grand total of FP used

to reach the OC. Figure 7.15 attempts to address this question showing the configuration and objective

convergence throughout CoH execution (as the number of FP increases). The green lines are the ’true’

optima.

Once again, after (Confidential) FP a satisfying result is already obtained (less than (Confidential) of

error). Flap behaviour, despite converging, behaves oddly. This can be explained both by the relative

’flatness’ of the objective function near the optimum as well as by the natural oscillations present in the

GP parametric SM used.

64

(a) Large error allowance parameterization results.

(b) Small error allowance parameterization results.

Figure 7.14: CoH full parameterization and run results.

65

(a) CoH Slap deflection convergence.

(b) CoH Flap deflection convergence.

(c) CoH MTOW objective convergence.

Figure 7.15: CoH Configuration and Objective convergence.

66

7.4 Different OC, Varying Runway

This subsection concerns itself with the presentation of TO network results. A TO network is simply a

2D graph of the MTOW Vs Runway Length (RL), for a given configuration, aircraft, conditions, etc. Ev-

erything so far has been done considering the RL fixed at 4000m, that is, a single point in a TO network.

Part of this was due to the absence of real-world CZMAXFT Results, which leads to artificial results for

2nd Segment limitations on a TO network. Nevertheless, the CoH algorithm can be legitimately validated

and used to build TO networks with the obtained configurations.

Presented in Figs. 7.17, 7.18 and 7.19 separately are the 3 traditional configurations and their

optimal counterparts. The optimal configurations were derived for (Confidential) RL for the 3,2 and 1+F

configurations respectively, by letting the CoH converge (see Table 7.2).

Configuration Slats (◦) Flaps (◦)

1+F – –1+F∗ – –

2 – –2∗ – –

3 – –3∗ – –

Table 7.2: 3 Configuration Sets, Traditional and Optimal (∗)

See Confidential Appendix, Data 10.

Of particular interest, and depicted in all TO networks, is the ’adaptive’ optimal configuration network.

This network was created from the MTOW values obtained using the OC for each RL. This provides an

upper theoretical bound on the TO performance that would be possible to obtain, a case where the

TO configuration would be adapted at each time for each RL. It also provides a means to assess the

’coverage’, or quality, of a given TO network, defined as:

Coverage(TO Network) =

∫ 4400m

1700mTO Network dRL∫ 4400m

1700mAdaptive TO Network dRL

× 100% (7.3)

The three configurations together do a remarkable job of covering the entire adaptive network (see

Fig. 7.16), with Conf 2 filling the gap. Qualitatively, Conf 2 (close-up Fig. 7.18) seems to provide the

best overall coverage, with Confs. 1+F and 3 (close-up in Figs. 7.17 and 7.19) excelling only near their

optima. In these two, substantial gains can be seen when comparing the OC with the traditional one.

Notice the overlap between each network with the adaptive network at the respective optimum RL value,

indicated by a circle (E.g., Conf 2 network overlaps with adaptive network at RL (Confidential)). Table

7.3 quantifies these conclusions.

For comparison purposes, note that a (Confidential) gain in coverage is roughly equivalent to a

(Confidential) gain in MTOW in all (Confidential).

See Confidential Appendix, Data 11.

The pre-flight Conf 2 is already optimal for the RL considered, whilst for Confs 1+F and 3 the gain

67

Figure 7.16: Traditional and CoH TO complete network coverage for the three different configurations.

Figure 7.17: Traditional and CoH configuration 1+F TO network detail.

68

Figure 7.18: Traditional and CoH configuration 2 TO network detail.

Figure 7.19: Traditional and CoH configuration 3 TO network detail.

69

Configuration Coverage (%) Gain (%)

1+F 97.89881+F∗ 98.6999 +0.8011

2 99.47102∗ 99.4565 -0.01450

3 98.62393∗ 98.9310 +0.3071

1+F and 3 99.57991+F∗ and 3∗ 99.8036 +0.2237

All 99.8157All∗ 99.8971 +0.0814

Table 7.3: Different TO network coverage.

is considerable, as mentioned before. However, considering individual configuration gains may not be

the most appropriate metric to judge a configuration, since those configurations are only expected to

perform well in a vicinity of their respective optimal RLs. Thus, a coverage improvements for Confs

1+F/3 together, as well as for all Confs, are also calculated in Table 7.3. The final gain is small, which

was expected given the use of a pre-flight CZMAX.

Thanks to the adaptive TO network, it is now possible to trace the evolution of the OC deflections

with varying RL. This is shown in Figure 7.20.

Figure 7.20: Optimal Configuration evolution with Runway Length.

See Confidential Appendix, Data 12.

70

7.5 Statistically Optimal Configuration

To go a bit further in the definition of optimal configuration, and as thoroughly discussed in Section 6.3.3,

a redefinition of the objective function was needed. Results using a statistically optimal OF are here pre-

sented and discussed in Table 7.4.

RL Range (m) Slats (◦) Flaps (◦) E(MTOW ) (kg)

[1700, 3000] – – –[3000, 3500] – – –[3500, 4400] – – –[1700, 4400] – – –

Table 7.4: Statistical OC results and their standard counterparts.

See Confidential Appendix, Data 13.

The first column on the left indicates the range of airport RLs that was considered for uncertainty

propagation. The rationale behind the four chosen ranges was simple: picking three ranges that roughly

corresponds to each of the 3 traditional configuration domains and a range encompassing all the domain,

to see which would be the ideal TO configuration if only one configuration was allowed.

The last column on the right is the OF value. In this case, it answers the following question: ”Taking-

Off at random from airports with runway ∈ RL Range, and considering more important airports to have

more TakeOffs, what is the expected average MTOW the aircraft can offer?”

Interestingly enough, the configuration values found for each of the first three ranges are not far from

the corresponding OC found before for the 3 Confs in Table 7.2 in Section 7.4.

See Confidential Appendix, Data 14.

However, the 3 configurations determined before do NOT take into account the runway importance

distribution, making the results obtained even more relevant.

Additionaly, for the full domain range, something close to a Conf 2 avails itself as the optimal con-

figuration. This underscores the previously made remark that the Conf 2 is the configuration with the

highest coverage.

Finally, a word of caution is due. This statistically optimal procedure requires thrice the amount of

time an ordinary CoH execution would, meaning it may render execution in real-time for flight testing

impractical.

71

Chapter 8

Conclusion

The present work’s contribution to the state-of-the-art in the Aircraft Performance domain consisted in

the definition, implementation and validation of a novel gradient-based procedure for systematic deter-

mination of the optimal takeoff configurations of an aircraft. In addition, a second SBO procedure was

proposed, mainly for validation purposes, and its viability was also demonstrated. Both procedures were

shown to not only meet the requirements of finding an optimal solution, but also of making it possible to

do so in an easily extensible and economical way. An innovative statistical objective function was finally

proposed, its merits and feasibility discussed.

Firstly, the main idea behind TO optimization was presented, along with key elements and tech-

niques on surrogate modelling, design of experiment and uncertainty propagation. This comprised the

theoretical backbone of the algorithms implemented.

Next the gradient-based solution architecture, along with its implementation, was described in detail.

The algorithm developed, named CoH for ConfOpt Hunter, was then validated resorting to a second

SBO procedure.

Validation, presentation and discussion of the results and insights discovered was done gradually in

4 steps, synthesized next.

Stage 1 consisted on the validation of simple MTOW calculation capabilities, resorting to surrogate

models to model internal aerodynamic labels. This allowed for MTOW calculation at any given contin-

uous point in the configuration design space. It was found that RSM label interpolation provided better

results than a LR interpolation. The key insight to retain at this stage is the importance of an adequate

representation of the basic aerodynamic coefficients, meaning proper choice of alphas and interpolation

technique, in order to get sensible MTOW values for any general TO configuration. This is true since

relatively small interpolation errors on drag coefficients were found to have a heavy impact on the output.

Stage 2 handled the optimal configuration determination itself, albeit for a fixed runway. At this point,

the viability of a SBO procedure was assessed. This SBO procedure, consisting in sampling the entire

configuration domain and optimizing over a model built from those samples, was also used to evaluate

its gradient-based counterpart. To do this, an extensive sampling pool was used to build a very high-

fidelity model, from which the ’true’ optimum was derived. Using this ’true’ optimum, the gradient method

72

was demonstrated to be a viable and flexible choice to find the OC. However, and this is the key idea to

take from this step, a SBO done over a restricted domain of the design space is also an equally good

alternative for this purpose.

Stage 3 introduced varying runway values, making it possible to validate the CoH algorithm for the

general case. CoH results were used to build different TO networks, and significant gains were shown

to be possible to obtain. The most valuable contribution this step provided was the possibility to trace an

adaptive TO network, that is, a network showing the maximum attainable MTOW at each runway length.

The importance of this adaptive network is significant, as it provides an upper bound against which to

compare all other networks and assess the quality of a particular set of configurations.

The fourth and final step was the statistical objective function definition. Although powerful, due to

the need for more calculations, this statistical OF might prove to be too slow for real-time applications.

Nevertheless, the most important information this step provided was one of confirmation. In other words,

the optima found using this OF are close to those found before, despite the fact that now information

on the relative importance of each runway length is incorporated in the calculation. This valuable infor-

mation was previously ignored. Eventually, this method will allow for slight non-intuitive corrections on

future OC settings.

Despite this being essentially an optimization work, the absolute most valuable insight the project

offered was not one of how much ’better’ the optimal configurations found are. Instead, the current work

proved those optima can be reached faster and cheaper than in the past, using a method that is easily

expandable to any number of problem dimensions. This means that if in the future AIRBUS envisions

optimizing more than 2 parameters for the OC, that can be easily done. Not only that, the methods

implemented, using state-of-art tools, are only the tip of the iceberg regarding the possibilities in this

domain: for example, cruise flight configurations, which resort to Variable Camber (DFS flap settings),

could be optimized using this method.

Future work directions on this topic are thus numerous, with a non-exhaustive list of three examples

reading as follows:

• Determination of an effective real-time alpha/speed selection method for label interpolation, or

interpolation solutions in general. It was found that even minor label interpolation errors could lead

to grave errors in the objective functions. A full-proof method of interpolation would be an essential

first step in rendering the CoH algorithm 100% robust.

• Study of the optimal number of configurations to include in the aircraft. Although the more configu-

rations are included the better the overall coverage of the TO network, these incur extra costs that

are not limited to the Performance domain. The possibility of using only one or two configurations

remains open.

• Algorithm innovations at a real-time level. Due to the real-time nature of flight-testing, different

procedures and/or order of evaluation of FP can be put into place to minimize aircraft idle flight

time.

73

74

Bibliography

[1] S. Koziel, D. E. Ciaurri, and L. Leifsson, “Surrogate-based methods,” in Computational Optimization,

Methods and Algorithms, pp. 33–59, Springer, 2011.

[2] J. Gray, K. T. Moore, T. A. Hearn, and B. A. Naylor, “A standard platform for testing and compari-

son of mdao architectures,” in 8th AIAA multidisciplinary design optimization specialist conference

(MDO), Honolulu, pp. 1–26, 2012.

[3] J. Sacks, W. J. Welch, T. J. Mitchell, and H. P. Wynn, “Design and analysis of computer experi-

ments,” Statistical science, pp. 409–423, 1989.

[4] N. V. Queipo, R. T. Haftka, W. Shyy, T. Goel, R. Vaidyanathan, and P. Kevin Tucker, “Surrogate-

based analysis and optimization,” Progress in aerospace sciences, vol. 41, no. 1, pp. 1–28, 2005.

[5] A. I. Forrester and A. J. Keane, “Recent advances in surrogate-based optimization,” Progress in

Aerospace Sciences, vol. 45, no. 1, pp. 50–79, 2009.

[6] S. R. Schmidt and R. G. Launsby, Understanding industrial designed experiments. Air Academy

Press, 1989.

[7] Z. Qian, C. C. Seepersad, V. R. Joseph, J. K. Allen, and C. J. Wu, “Building surrogate models

based on detailed and approximate simulations,” Journal of Mechanical Design, vol. 128, no. 4,

pp. 668–677, 2006.

[8] A. G. Airbus, Getting to Grips with Aircraft Performance. Customer Services, Blagnac, 2002.

[9] J. A. Authorities, “Joint aviation requirements. jar-25. large aeroplanes,” Civil Aviation Authority

Printing & Publication Services, Greville House, vol. 37, 1994.

[10] A. Forrester, A. Sobester, and A. Keane, Engineering design via surrogate modelling: a practical

guide. John Wiley & Sons, 2008.

[11] T. Hastie, R. Tibshirani, J. Friedman, and J. Franklin, “The elements of statistical learning: data

mining, inference and prediction,” The Mathematical Intelligencer, vol. 27, no. 2, pp. 83–85, 2005.

[12] G. E. Box and N. R. Draper, Empirical model-building and response surfaces. John Wiley & Sons,

1987.

[13] C. L. Lawson and R. J. Hanson, Solving least squares problems, vol. 161. SIAM, 1974.

75

[14] E. Burnaev, M. Belyaev, and A. Lyubin, “Construction of approximation based on linear expansion

in heterogeneous nonlinear functions,”

[15] T. Zou, M. Cai, and X. Shu, “Response surface methodology and improved interval analysis

method—for analyzing uncertainty in accident reconstruction,” Forensic science international,

vol. 222, no. 1, pp. 306–312, 2012.

[16] R. H. Myers, D. C. Montgomery, and C. M. Anderson-Cook, Response surface methodology: pro-

cess and product optimization using designed experiments, vol. 705. John Wiley & Sons, 2009.

[17] M. L. Stein, Interpolation of spatial data: some theory for kriging. Springer, 1999.

[18] N. Cressie, “The origins of kriging,” Mathematical geology, vol. 22, no. 3, pp. 239–252, 1990.

[19] C. Paciorek and M. Schervish, “Nonstationary covariance functions for gaussian process regres-

sion,” Advances in neural information processing systems, vol. 16, pp. 273–280, 2004.

[20] P. C. Mahalanobis, “On the generalized distance in statistics,” Proceedings of the National Institute

of Sciences (Calcutta), vol. 2, pp. 49–55, 1936.

[21] C. K. Williams and C. E. Rasmussen, “Gaussian processes for machine learning,” the MIT Press,

vol. 2, no. 3, p. 4, 2006.

[22] R. Renka, “Interpolatory tension splines with automatic selection of tension factors,” SIAM Journal

on Scientific and Statistical Computing, vol. 8, no. 3, pp. 393–415, 1987.

[23] S. Pruess, “An algorithm for computing smoothing splines in tension,” Computing, vol. 19, no. 4,

pp. 365–373, 1978.

[24] L. Finschi, “Quasi-monte carlo: An empirical study on low-discrepancy sequences,” tech. rep.,

Working paper, Swiss Federal Institute of Technology, 1996.

[25] F. A. C. Viana, V. Steffen Jr, G. Venter, and V. Balabanov, “On how to implement an affordable

optimal latin hypercube,” 2007.

[26] H. Faure, “Discrepance de suites associees a un systeme de numeration (en dimension s),” Acta

Arithmetica, vol. 41, no. 4, pp. 337–351, 1982.

[27] J. H. Halton, “Algorithm 247: Radical-inverse quasi-random point sequence,” Communications of

the ACM, vol. 7, no. 12, pp. 701–702, 1964.

[28] I. M. Sobol, “Uniformly distributed sequences with an additional uniform property,” USSR Compu-

tational Mathematics and Mathematical Physics, vol. 16, no. 5, pp. 236–242, 1976.

[29] M. Stein, “Large sample properties of simulations using latin hypercube sampling,” Technometrics,

vol. 29, no. 2, pp. 143–151, 1987.

[30] J. C. Helton and F. J. Davis, “Latin hypercube sampling and the propagation of uncertainty in analy-

ses of complex systems,” Reliability Engineering & System Safety, vol. 81, no. 1, pp. 23–69, 2003.

76

[31] R. L. Iman, Latin hypercube sampling. Wiley Online Library, 2008.

[32] J.-S. Park, “Optimal latin-hypercube designs for computer experiments,” Journal of statistical plan-

ning and inference, vol. 39, no. 1, pp. 95–111, 1994.

[33] F. A. Viana, “Things you wanted to know about the latin hypercube design and were afraid to ask,”

in 10th World Congress on Structural and Multidisciplinary Optimization, Orlando, Florida, USA (cf.

p. 69), 2013.

[34] Proceedings of the 58th ISI world statistics congress, (Dublin, Ireland), 2011.

[35] R. S. John and N. R. Draper, “D-optimality for regression designs: a review,” Technometrics, vol. 17,

no. 1, pp. 15–23, 1975.

[36] H. Dette and T. E. O’Brien, “Optimality criteria for regression models based on predicted variance,”

Biometrika, vol. 86, no. 1, pp. 93–106, 1999.

[37] N.-K. Nguyen and A. J. Miller, “A review of some exchange algorithms for constructing discrete

d-optimal designs,” Computational Statistics & Data Analysis, vol. 14, no. 4, pp. 489–498, 1992.

[38] W. Chen, J. K. Allen, K.-L. Tsui, and F. Mistree, “A procedure for robust design: minimizing varia-

tions caused by noise factors and control factors,” Journal of Mechanical Design, vol. 118, no. 4,

pp. 478–485, 1996.

[39] G.-J. Park, T.-H. Lee, K. H. Lee, and K.-H. Hwang, “Robust design: an overview,” AIAA journal,

vol. 44, no. 1, pp. 181–191, 2006.

[40] M. Padulo, M. S. Campobasso, and M. D. Guenov, “Novel uncertainty propagation method for robust

aerodynamic design,” AIAA journal, vol. 49, no. 3, pp. 530–543, 2011.

[41] P. Mattia, C. M. Sergio, G. M. Dimitrov, et al., “Comparative analysis of uncertainty propagation

methods for robust engineering design,” Guidelines for a Decision Support Method Adapted to

NPD Processes, 2007.

[42] L. Lewis and A. Parkinson, “Robust optimal design using a second-order tolerance model,” Re-

search in Engineering Design, vol. 6, no. 1, pp. 25–37, 1994.

[43] L. L. Green, H.-Z. Lin, and M. R. Khalessi, “Probabilistic methods for uncertainty propagation ap-

plied to aircraft design,” AIAA Paper, vol. 3140, p. 2002, 2002.

[44] A. Parkinson, C. Sorensen, and N. Pourhassan, “A general approach for robust optimal design,”

Journal of Mechanical Design, vol. 115, no. 1, pp. 74–80, 1993.

[45] S. J. Julier and J. K. Uhlmann, “A general method for approximating nonlinear transformations

of probability distributions,” tech. rep., Technical report, Robotics Research Group, Department of

Engineering Science, University of Oxford, 1996.

77

[46] H. Xu and S. Rahman, “A generalized dimension-reduction method for multidimensional integra-

tion in stochastic mechanics,” International Journal for Numerical Methods in Engineering, vol. 61,

no. 12, pp. 1992–2019, 2004.

[47] M. Cioffi, A. Formisano, R. Martone, G. Steiner, and D. Watzenig, “A fast method for statistical

robust optimization,” Magnetics, IEEE Transactions on, vol. 42, no. 4, pp. 1099–1102, 2006.

[48] D. G. Luenberger, Introduction to linear and nonlinear programming, vol. 28. Addison-Wesley Read-

ing, MA, 1973.

[49] J. Snyman, Practical mathematical optimization: an introduction to basic optimization theory and

classical and new gradient-based algorithms, vol. 97. Springer, 2005.

[50] P. Wolfe, “Convergence conditions for ascent methods,” SIAM review, vol. 11, no. 2, pp. 226–235,

1969.

78