estimation of precursors for extreme events using the

65
β—‹

Upload: others

Post on 13-Mar-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

Estimation of precursors for extreme events using

the adjoint based optimization approach

by

Rishabh Ishar

B.E., Punjab Engineering College (2018)

Submitted to the Department of Mechanical Engineeringin partial fulfillment of the requirements for the degree of

Master of Science in Mechanical Engineering

at the

MASSACHUSETTS INSTITUTE OF TECHNOLOGY

February 2020

c Massachusetts Institute of Technology 2020. All rights reserved.

Author . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Department of Mechanical Engineering

January 14, 2020

Certified by. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Themistoklis P. Sapsis

Doherty Associate Professor in Ocean UtilizationThesis Supervisor

Accepted by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Nicolas Hadjiconstantinou

Chairman, Committee on Graduate Students

2

Estimation of precursors for extreme events using the adjoint

based optimization approach

by

Rishabh Ishar

Submitted to the Department of Mechanical Engineeringon January 14, 2020, in partial fulfillment of the

requirements for the degree ofMaster of Science in Mechanical Engineering

Abstract

We formulate a generalized optimization problem for a non-linear dynamical systemgoverned by a set of differential equations. The plant under focus is the 2-D Kol-mogorov flow, as this flow has inherent turbulence which would give rise to chaosand intermittent bursts in a selected observable. As a first step, an observable withpotential extreme events in its time series is selected. In our case, we choose thekinetic energy of the flow field as the observable under study. The next step is toderive the adjoint equations for the kinetic energy that is the quantity of interestwith the velocity field as the optimizing variable. This obtained velocity field formsthe precursor for extreme events in the kinetic energy. The prediction capabilities forthis precursor are then explored in more detail. The goal is to select the precursorsuch that it predicts the extreme events in a given time horizon which can generatewarning signals effectively. We also present a coupled flow solver in Nek5000 andadjoint solver in MATLAB, the latter can be applied to any dynamical system tostudy the extreme events and obtain the relevant precursor. In a consecutive section,the results for extreme events in the kinetic energy and the lift coefficient for theflow over a 2-D airfoil are presented. As part of future work, the implementation andapplication of the solver for the flow past the airfoil and over a 3-D Ahmed body areproposed.

Thesis Supervisor: Themistoklis P. SapsisTitle: Doherty Associate Professor in Ocean Utilization

3

4

Acknowledgments

First and foremost, I would like to thank my supervisor, Professor Themis Sapsis,

for his constant guidance during the course of my research. This work wouldn’t have

been possible without his support, that he made sure I had enough, even during his

sabbatical.

I would like to express my gratitude to Dr. Antoine Blanchard, for his simplifying

explanations and great help during my research. I thank him for all his time. His

expertise in fluid mechanics and dynamical system analysis gave me a lot of insight

into these topics.

I want to thank Prof. Bernd R. Noack, who has been a great collaborator since

my third year of undergrad. He introduced me to topics like model reduction and

machine learning that were really important during my studies at MIT. I also want

to thank all my collaborators from USA, Germany, France, and Poland, including

Dr. Eurika Kaiser, Prof. Marek Morzynski, Prof. Steve Brunton, Prof. Wolfgang

Schroder, Prof. Richard Semaan, Mr. Daniel Fernex, Mr. Marian Albers, Dr. Pascal

Meysonnat, Ms. Camila Chovet, Mr. Guy Maceda, and Prof. Nathan Kutz.

I would like to thank my labmates Dr. Mohammad Farazmand, Mr. Alexis-

Tzianni Charalampopoulos, Mr. Zhong Yi Wan, Mr. Stephen Guth, and Dr. Hassan

Arbabi, for the stimulating discussions regarding coursework and research. I also

want to thank the Army Research Office for supporting me.

Finally, I want to thank my family and friends for their support during the course

of my degree.

5

6

Contents

1 Introduction 13

1.1 Prediction and control of extreme events . . . . . . . . . . . . . . . . 14

1.2 Motivation and contributions . . . . . . . . . . . . . . . . . . . . . . 15

2 Adjoint based optimization approach 17

2.1 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.1.1 The generalized optimization problem . . . . . . . . . . . . . . 18

2.1.2 Derivation of the adjoint equations . . . . . . . . . . . . . . . 18

2.1.3 Gradient descent . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.1.4 Layout for one complete loop/iteration of the adjoint solver . 25

2.2 Statistical analysis and prediction of extreme events . . . . . . . . . . 25

2.2.1 Joint and conditional statistics of the extreme events . . . . . 26

2.3 An overview of Nek5000 (flow solver) . . . . . . . . . . . . . . . . . . 28

3 Applications and results 29

3.1 Kolmogorov flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3.1.1 Flow solver, setup, and simulation . . . . . . . . . . . . . . . . 30

3.1.2 POD and extreme event analysis results . . . . . . . . . . . . 31

3.2 2-D Airfoil flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3.2.1 Flow solver, setup, and simulation . . . . . . . . . . . . . . . . 43

3.2.2 Lift coefficient over time analysis to determine the critical Reynolds

number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

7

3.2.3 Generating the stochastic inlet velocity and mode reduction

using Karhunen-Loeve expansion . . . . . . . . . . . . . . . . 45

3.2.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

4 Conclusions and future work 57

4.1 Summary of the results and discussion . . . . . . . . . . . . . . . . . 57

4.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

8

List of Figures

2-1 Layout of the loop to be completed in each iteration of the adjoint

problem till the solution converges . . . . . . . . . . . . . . . . . . . . 25

3-1 2-D mesh for the Kolmogorov flow. . . . . . . . . . . . . . . . . . . . 31

3-2 Time series plot of the kinetic energy. . . . . . . . . . . . . . . . . . . 32

3-3 Vorticity (the scale is -2 to 2) of the POD modes 𝑖 = βˆ‡ Γ— 𝑖(),

𝑖 = 1, . . . , 10 for the Kolmogorov flow. . . . . . . . . . . . . . . . . . . 33

3-4 The first 10 POD coefficients for the Kolmogorov flow. . . . . . . . . 34

3-5 Initial guess of the velocity field. . . . . . . . . . . . . . . . . . . . . . 35

3-6 Precursor obtained after 1500 iterations of the adjoint problem. . . . 35

3-7 πœ† and Kinetic energy (KE) vs. snapshot number (the time lag between

πœ† and KE is 1 time unit). . . . . . . . . . . . . . . . . . . . . . . . . 36

3-8 Zoomed peak near t = 2040 for the πœ† and Kinetic energy vs. snapshot

number plot. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

3-9 Plot showing correlation between πœ† and π·π‘š (the time lag between πœ†

and KE is 1 time unit). . . . . . . . . . . . . . . . . . . . . . . . . . . 37

3-10 Joint pdf of πœ† and π·π‘š (the time lag between πœ† and KE is 1 time unit). 37

3-11 Conditional pdf of πœ† and π·π‘š (the time lag between πœ† and KE is 1 time

unit). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

3-12 πœ† and Kinetic energy vs. snapshot number (10000 time units simula-

tion, the time lag between πœ† and KE is 1 time unit). . . . . . . . . . 39

3-13 Plot showing correlation between πœ† and π·π‘š (10000 time units simula-

tion, the time lag between πœ† and KE is 1 time unit). . . . . . . . . . 39

9

3-14 Joint pdf of πœ† and π·π‘š (10000 time units simulation, the time lag

between πœ† and KE is 1 time unit). . . . . . . . . . . . . . . . . . . . . 40

3-15 Conditional pdf of πœ† and π·π‘š (10000 time units simulation, the time

lag between πœ† and KE is 1 time unit). . . . . . . . . . . . . . . . . . . 40

3-16 πœ† and Dissipative energy (DE) vs. snapshot number (10000 time units

simulation, the time lag between πœ† and DE is 1 time unit). . . . . . . 41

3-17 Plot showing correlation between πœ† and π·π‘š (DE) (10000 time units

simulation, the time lag between πœ† and DE is 1 time unit). . . . . . . 41

3-18 Joint pdf of πœ† and π·π‘š (DE) (10000 time units simulation, the time lag

between πœ† and DE is 1 time unit). . . . . . . . . . . . . . . . . . . . . 42

3-19 Conditional pdf of πœ† and π·π‘š (DE) (10000 time units simulation, the

time lag between πœ† and DE is 1 time unit). . . . . . . . . . . . . . . . 42

3-20 2-D mesh of the flow around the airfoil. . . . . . . . . . . . . . . . . . 44

3-21 Vorticity snapshots at T=100 for Reynolds numbers near 𝑅𝑒𝑐: (a) 𝑅𝑒

= 90 (b) 𝑅𝑒 = 105 (c) 𝑅𝑒 = 110. . . . . . . . . . . . . . . . . . . . . 48

3-22 The kinetic energy plot for 3000 time units. . . . . . . . . . . . . . . 49

3-23 The inlet velocity input signal. . . . . . . . . . . . . . . . . . . . . . . 49

3-24 (a) Re = 90 for t = 15 to 200 (b) Re = 90, zoomed peaks (c) Re =

105, for t = 15 to 300 (d) Re = 105, zoomed peaks (e) Re = 110, for t

= 15 to 300 (f) Re = 110, zoomed peaks. . . . . . . . . . . . . . . . 50

3-25 Vorticity (the scale is -2 to 2) of the POD modes 𝑖 = βˆ‡ Γ— 𝑖(),

𝑖 = 1, . . . , 10 for the airfoil flow. . . . . . . . . . . . . . . . . . . . . . 51

3-26 The first 10 POD coefficients for the airfoil flow. . . . . . . . . . . . 52

3-27 (a) 𝐢𝐿 from t = 20 to 3000 (b) The averaged curve (c) fluctuations in

𝐢𝐿 calculated by removing the moving average (5 time units moving

window). π‘ˆ is taken as the inlet KLE velocity with zero mean. From

t = 20 to 3000. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

3-28 π‘₯ component of the velocities at the three probes (a) 𝑉π‘₯1 (b) 𝑉π‘₯2 (c)

𝑉π‘₯3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

3-29 𝑦 component of the velocities at the three probes (a) 𝑉𝑦1 (b) 𝑉𝑦2 (c) 𝑉𝑦3. 55

10

List of Tables

3.1 Location of the three probes. . . . . . . . . . . . . . . . . . . . . . . . 47

4.1 Summary of the extreme events prediction results for the Kolmogorov

flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

11

12

Chapter 1

Introduction

Extreme events [1] form an integral part of complex dynamical systems. Also called

rare events, they have a really low probability of occurrence but can have significant

outcomes. Their relevance stems from the fact that they occur in systems like the

ocean (rogue waves ([26]), freak waves ([63], [40], [23], [58], [84])), climate (hurricanes

([56]), stock market ([31], [30], [59]), etc. which can be highly influential. Though

rare, they can have catastrophic economic and social consequences which makes their

study even more important. A prominent feature of extreme events is that they

do not follow Gaussian statistics, and the number of their occurrences cannot be

accurately predicted by the Gaussian statistics. The variables involved follow non-

Gaussian statistics in the probability density function (pdf), an example can be a

heavy-tailed pdf where the values of the observed variables can be many standard

deviations away from the mean. Another example can be of a Poisson [13] or weibull

distribution [32]. Some examples of non-Gaussian statistics can be seen in [83], [48],

[57], [20], [47], [45]. This poses a key challenge for efficiently describing the behavior

in the pdf of a dynamical system, which could be non-linear, highly dimensional, or

have intermittent events in the response.

13

1.1 Prediction and control of extreme events

An informal definition of an extreme event is that they - (a) are rare (b) occur in ran-

dom intervals (c) there is a system variable that has intermittent bursts in its observed

values and takes on extreme values (d) they are the characteristic of a system, not

necessarily forced. In a mathematical perspective, an extreme event is said to occur

when the value of a variable crosses a threshold, within the upper and lower bounds

of the range of the data set. The properties of dynamical systems that cause extreme

events are stochasticity and non-linearities inhibited within the system. Undesirable

random processes that are inherently part of the system contribute to the stochastic

behavior. Hence, due to these random processes, extreme events occur spontaneously

with few warning signals which makes their quantification and prediction a difficult

task.

A lot of work being done on the prediction of extreme events is based on the

extreme value theory (EVT) and can be found in [41], [16],[19], [2], and more enclosed

references. The first EVT states that the maxima of the sample asymptotically follow

one of the three distributions, namely Gumbel (Type I), Frechet (Type II) or Weibull

(Type III). This is for the univariate case. For the multivariate case, [79], [21],

[22], [12], give a solid foundation on analyzing the extreme event statistics using

prescribed-size blocks or the observed values of the variable that exceed a certain

level of magnitude. J. H. J. Einmahl and others ([27]) proposed a nonparametric

model that generates estimators while satisfying the moment constraints and applied

it to fit an angular density to a bivariate distribution.

Time series analysis [35] is another approach towards the quantification of extreme

events. It has a strong connection with chaos theory ([33], [42], [54], [24], [25], [86]).

Degrees of freedom (DOF) of the system played an important role as it was discovered

that an infinite number of DOF (infinitely many dimensions) can be determined from

a finite number of DOF (low dimensional) in a non-linear system like the Lorenz

attractor [49]. Stochasticity in such systems was established using DOF. If the DOF

were finite then the system would be deterministic otherwise it would be stochastic

14

(infinitely many DOF). The underlying dynamics of the attractor were evaluated and

reconstructed by a ’method of delays’ which was further developed ([71] and [78]).

Due to noise in the practical/experimental data, only a base structure of the attractor

can be described, without all the intricacies.

A better physical understanding is provided by dynamical and stochastic modeling

of the system.

1.2 Motivation and contributions

The study of turbulent flows using invariant solutions of the Navier-Stokes equa-

tions has been an active area of research ([38]). There isn’t much information about

the analytical behavior of these invariant solutions due to their complex properties

in space and time. M. Farazmand [28] derived the adjoint equations to compute

the steady-state and traveling-wave solutions for a flow with incompressible Navier-

Stokes equations with periodic boundary conditions and a space-dependent forcing.

A steady-state solution is established as a precursor that gives rise to extreme events

in the dissipative energy of the flow.

O. Soto and others ([76]) used a scheme formulated for the continuous adjoint

problem and applied it to optimize the shape in fluid flows, examples include lift

control over a cylinder and the flow over a 3-D hypersonic wing (angle of attack =

5 degrees). The approach included obtaining the gradient of the target objective

function with respect to a selected number of design parameters, this was further im-

proved by a pseudo-shell approach. The purpose of shape optimization in an airfoil

is to improve its performance. M. Schramm and others ([73]) used the adjoint ap-

proach to maximize the lift to drag ratio in a flow over a 2-D airfoil with and without

constraints.

In this work, we develop a finite-time horizon approach to the adjoint problem

and build on the instantaneous problem as provided in [29] with a few modifications.

Also, P. J. Blonigan and others ([9]) studied the turbulent channel flow and identified

the precursors for extreme events from a high probability manifold of the state. A

15

very small time-horizon is used so that the results can be compared with the instan-

taneous problem. We use the kinetic energy of the flow as our observable for extreme

events, derive the associated adjoint equations for an arbitrary fluid flow with given

constraints, and solve the equations to obtain the precursor. This obtained precursor

is used with a time lag to estimate the prediction capabilities before the occurrence of

the extreme events. We also present a unified framework to solve the coupled adjoint

problem using Nek5000 and MATLAB. Hence, this can be applied to any fluid flow

system which has inherent extreme events and establish the precursors that lead to

intermittent bursts in the space of the targeted variable. Here, we present the results

as obtained for the Kolmogorov flow and the flow over a 2-D airfoil at an angle of

attack of 30 degrees.

16

Chapter 2

Adjoint based optimization approach

The goal of this chapter is to describe the generalized optimization problem for a

dynamical system that is governed by a set of differential equations which can be

potentially non-linear. Some examples can be seen in [53], [17], [68], [43], [36], [44],

[52], [65], [75], [15]. If there are constraints associated with the governing equations,

then it becomes a bounded and constrained optimization problem. An observable is

then chosen for which we optimize the adjoint equations. After this, we formulate the

Lagrangian (β„’). The gradient of this Lagrangian is set to zero for all the variables

involved in it, which is done to converge to the minima of the observable in the

problem. Standard gradient descent is then used to iterate and reach the updated

initial condition, that would be used as the new initial condition for the next iteration.

This process is continued until the norm difference between the new and old initial

condition falls below a certain threshold.

17

2.1 Methodology

2.1.1 The generalized optimization problem

In this section, we define the governing equations of a general dynamical system as

πœ•π‘‘π‘’ = 𝒩 (𝑒) (2.1a)

𝒦(𝑒) = 0 (2.1b)

𝑒(π‘₯, 𝑑0) = 𝑒0(π‘₯), (2.1c)

where 𝑒 determines the complete state of the system, 𝒩 and 𝒦 are differential oper-

ators that are based on the physical model of the system. They govern the evolution

of the system in space and time, and can have non-linear terms associated with them.

The focus of this section is to predict the occurrence of extreme events, via the inter-

mittent bursts in an observable. This observable is denoted by 𝐼 and it would have a

time series with rare extreme events after certain intervals. 𝑒(𝑑) is the driver for these

extreme occurrences and there are certain instances of 𝑒(𝑑) that would cause sudden

bursts in the observable 𝐼. We can imagine the attractor with regions of instability,

and as soon as the state 𝑒(𝑑) reaches this instability region, there would be a large

deviation from the general trajectory, which would cause an extreme event in the

time-series of the observable. We examine these instability regions and the associated

states. Also, these states should have a non-zero probability of occurrence, which can

be enforced by using a constraint. This prevents the adjoint solver from converging to

states that have negligible probability of occurrence and would not occur in practical

simulations/experiments.

2.1.2 Derivation of the adjoint equations

In this section, we derive the adjoint equations for a dynamical system that is gov-

erned by the Navier-Stokes equations. The flows considered here are assumed to be

18

incompressible. These set of equations can be applied to both 2-D and 3-D fluid flow

systems. The system can also be subjected to a number of constraints which enforce

the state of the system to stay close to the attractor. This means that during any

iteration of the solver, the state does not assume impractical states and the ones

which have really low probabilities of occurrence. The observable in which we are

interested in to study the extreme event statistics is the kinetic energy of the system.

This is given by

𝐼() =1

2

∫Ω

. 𝑑Ω

where is the velocity field and Ξ© is the flow domain. The velocity field is integrated

over each element in the whole domain, and then summed up to get the total kinetic

energy, 𝐼().

For a fluid flow, the constrained optimization problem becomes

max (𝐼((𝑇 )) βˆ’ 𝐼((0))) (2.2)

Through this, the deviation in the kinetic energy is being maximized, that is we aim

to solve for the initial condition (0) that gives rise to the state (𝑇 ) having the

maximum growth in the kinetic energy in the time horizon 𝑇 .

Here is subject to the following two governing equations and a feasibility constraint

given by

𝒩 () = πœ•π‘‘ = βˆ’.βˆ‡βˆ’βˆ‡π‘ + πœˆβˆ† + 𝑓 (2.3)

where 𝒩 () is the Navier-stokes incompressible momentum equation and contains

non-linear terms and

𝒦() = βˆ‡. = 0 (2.4)

where 𝒦() is the Navier-stokes incompressibility condition.

Before we get to the feasibility constraint, we use proper orthogonal decomposition

19

(POD) to get the POD coefficients. Some studies on POD can be seen in [5], [18], [60],

[69], [85], [46]. Dynamic mode decomposition is another method that can be used

to extract coherent structures and dynamic information from flow fields as global

stability modes ([72], [61], [81]). Other areas of reduced order modeling have been

explored in [3], [7], [8], [6], [77]. When POD is used as an approximation method,

then any state on the attractor can be written as

(π‘₯, 𝑑) = (π‘₯) +π‘›βˆ‘

𝑖=1

πœ‰π‘–(𝑑) 𝑣𝑖(π‘₯) (2.5)

where (π‘₯) is the mean flow, 𝑣𝑖 are the POD mdoes, and πœ‰π‘– are the POD coefficients.

This is an approximation of a state (π‘₯, 𝑑) using the first 𝑛 POD modes. 𝑛 can be

chosen based on how much energy is desired in the state-space. More POD modes

provide better approximation of the state and capture more energy.

There is a feasibility constraint to ensure the initial states stay close to the attractor,

given byπ‘›βˆ‘

𝑖=1

πœ‰2π‘–πœ†π‘–

≀ π‘Ÿ0 (2.6)

where πœ‰π‘– is the 𝑖-th POD coefficient, πœ†π‘– is the 𝑖-th eigenvalue, π‘Ÿ0 is a constant that

can be obtained as an average of the values from all the snapshots, 𝑛 are the number

of modes to be used that contain more than 98% of the energy.

We define the Lagrangian, β„’(, 𝑝, *, 𝑝*, (0), (𝑇 )) as

β„’ = 𝐼((𝑇 )) βˆ’ 𝐼((0)) βˆ’βˆ« 𝑇

0

(*, πœ•π‘‘βˆ’π‘())π‘‘πœ

βˆ’βˆ« 𝑇

0

(𝑝*,βˆ‡.)π‘‘πœ βˆ’ πœ†(((0),π‘Š (0)) βˆ’ π‘Ÿ0) (2.7)

where 𝑝 is the pressure, * is the Lagrangian multiplier for the velocity field, 𝑝* is

the Lagrangian multiplier for the pressure field, π‘Š is the element-wise inverse of the

eigenvalue matrix, (0) is the initial velocity field, (𝑇 ) is the velocity field at time

𝑇 .

20

To evaluate the adjoint equations, we need to set the partial derivative of β„’ with

respect to each of the six parameters equal to zero, so that the minimization problem

can reach a minima in the state-space that corresponds to the maximum growth in

the kinetic energy of the system.

1.) First, we set the partial derivative of the Lagrangian with respect to the ve-

locity field at any point in space and time equal to zero. Hence, all the other 5

parameters can be considered as constants.

πœ•β„’πœ•

= βˆ’βˆ« 𝑇

0

(*, πœ•π‘‘βˆ’π‘())π‘‘πœ βˆ’βˆ« 𝑇

0

(𝑝*,βˆ‡.)π‘‘πœ = 0

∫ 𝑇

0

πœ•

πœ•

∫Ω

*(πœ•π‘‘βˆ’π‘()) 𝑑Ω π‘‘πœ +

∫ 𝑇

0

πœ•

πœ•

∫Ω

𝑝*βˆ‡. 𝑑Ω π‘‘πœ = 0

∫ 𝑇

0

πœ•

πœ•

∫Ω

*(πœ•π‘‘ + .βˆ‡ + βˆ‡π‘βˆ’ πœˆβˆ†βˆ’ 𝑓) 𝑑Ω π‘‘πœ +

∫ 𝑇

0

πœ•

πœ•

∫Ω

𝑝*βˆ‡. 𝑑Ω π‘‘πœ = 0

∫ 𝑇

0

∫Ω

(βˆ’πœ•π‘‘* βˆ’ .βˆ‡* + *.βˆ‡π‘‡ + πœˆβˆ†*) 𝑑Ω π‘‘πœ +

∫ 𝑇

0

πœ•

πœ•

∫Ω

(βˆ’βˆ‡π‘*) 𝑑Ω π‘‘πœ = 0

where we have used integration by parts and the divergence theorem

∫ 𝑇

0

∫Ω

(βˆ’πœ•π‘‘* βˆ’ .βˆ‡* + *.βˆ‡π‘‡ + πœˆβˆ†* βˆ’βˆ‡π‘*) 𝑑Ω π‘‘πœ = 0

which gives us the first adjoint Navier-stokes equation

πœ•π‘‘* + .βˆ‡* βˆ’ *.βˆ‡π‘‡ βˆ’ πœˆβˆ†* + βˆ‡π‘* = 0 (2.8)

This equation is not the same as the one we have for 𝒩 (𝑒), this is the adjoint version

of the incompressible Navier-stokes equation.

2.) Now we set the partial derivative of the Lagrangian with respect to the pres-

sure, 𝑝, equal to zero.

21

πœ•β„’πœ•π‘

=

∫ 𝑇

0

πœ•

πœ•π‘

∫Ω

*. βˆ‡π‘ 𝑑Ω π‘‘πœ = 0

∫ 𝑇

0

πœ•

πœ•π‘

∫Ω

(βˆ’βˆ‡.*)𝑝 𝑑Ω π‘‘πœ = 0

which gives us the second adjoint Navier-stokes equation

βˆ‡.* = 0 (2.9)

3.) Here, the partial derivative of the Lagrangian multiplier of is set to zero. That

is with respect to *

πœ•β„’πœ•* =

∫ 𝑇

0

πœ•

πœ•*

∫Ω

*(πœ•π‘‘ + .βˆ‡ + βˆ‡π‘βˆ’ πœˆβˆ†βˆ’ 𝑓) 𝑑Ω π‘‘πœ = 0

∫ 𝑇

0

∫Ω

(πœ•π‘‘ + .βˆ‡ + βˆ‡π‘βˆ’ πœˆβˆ†βˆ’ 𝑓) 𝑑Ω π‘‘πœ = 0

which gives us the first Navier-Stokes equation

πœ•π‘‘ = βˆ’.βˆ‡βˆ’βˆ‡π‘ + πœˆβˆ† + 𝑓 (2.10)

This equation is the same as the one we have for 𝒩 (𝑒), that is the incompressible

momentum Navier-stokes equation.

4.) Here, the partial derivative of the Lagrangian multiplier of 𝑝 is set to zero. That

is with respect to 𝑝*

πœ•β„’πœ•π‘*

=

∫ 𝑇

0

πœ•

πœ•π‘*

∫Ω

𝑝*βˆ‡. 𝑑Ω π‘‘πœ = 0

∫ 𝑇

0

∫Ω

βˆ‡. 𝑑Ω π‘‘πœ = 0

22

which gives us the second Navier-Stokes equation (the incompressibility condition)

βˆ‡. = 0 (2.11)

5.) With respect to the initial condition of the velocity field, (0), the partial deriva-

tive of the Lagrangian is set to zero.

πœ•β„’πœ•(0)

= βˆ’πœ•πΌ((0))

πœ•(0)+ *(0) βˆ’ πœ†

πœ•

πœ•(0)

∫Ω

(0)𝑇 π‘Š (0) 𝑑Ω = 0

which gives

2(𝐼 + πœ†π‘Š )(0) βˆ’ *(0) = 0 (2.12)

The equation above plays a major role in calculating the gradient descent for the

initial velocity field, the parameter with respect to which the kinetic energy growth

is being maximized. This is described in the next section.

6.) With respect to the initial condition of the velocity field, (𝑇 ), the partial deriva-

tive of the Lagrangian is set to zero.

πœ•β„’πœ•(𝑇 )

=πœ•πΌ((𝑇 ))

πœ•(𝑇 )βˆ’ *(𝑇 ) = 0

where *(𝑇 ) comes from the integration by parts done in part 1.)

πœ•

πœ•(𝑇 )

∫Ω

(𝑇 ) . (𝑇 ) 𝑑Ω βˆ’ *(𝑇 ) = 0

which gives

2(𝑇 ) βˆ’ *(𝑇 ) = 0 (2.13)

This equation provides the initial condition *(𝑇 ) for the adjoint problem, as it is

solved backwards in time to get to the final state *(0).

23

2.1.3 Gradient descent

This section aims to give a brief overview of the gradient descent optimization method.

Gradient descent ([14], [10], [51], [67], [70]) is used to find the minima of a function by

moving in the direction of the steepest descent. This is achieved through a number of

iterations, with the parameter in focus getting updated at each iteration. The direc-

tion of the steepest descent is given by the negative of the gradient of the functional

with respect to the parameter.

Gradient descent finds its use in various machine learning applications where a

model can have many parameters being updated at once. The size of the step which

dictates by how much we need to move in the direction of the steepest descent is

called the learning rate. This is denoted by 𝛼 in this section. When the learning rate

is high, convergence can be reached faster, as we take bigger steps. But there is a

risk of overshooting the minima in each iteration, which can make the solver to get

stuck in an infinite loop where it never reaches the minima. A safe option would be

choosing a low learning rate, but it would be slow and time consuming. The gradient

calculation is that of the function to be optimized with respect to a parameter.

In this thesis, the function would be given by equation 2.2 and the parameter is

the initial velocity field. The gradient of this function, πœ•π‘“πœ•(0)

, is the same as the gra-

dient of the Lagrangian with respect to the initial velocity. Hence, it can be replaced

by πœ•β„’πœ•(0)

Now, we use gradient descent to update the initial guess π‘œπ‘™π‘‘(0)

𝑛𝑒𝑀(0) = π‘œπ‘™π‘‘(0) βˆ’ π›Όπœ•β„’

πœ•(0)

which gives

𝑛𝑒𝑀(0) = π‘œπ‘™π‘‘(0) βˆ’ 𝛼(2(𝐼 + πœ†π‘Š )π‘œπ‘™π‘‘(0) βˆ’ *(0)) (2.14)

where 𝛼 is the learning rate.

This process is repeated until the initial condition solution reaches convergence, that

is the value of the gradient πœ•β„’πœ•(0)

tends to zero.

24

2.1.4 Layout for one complete loop/iteration of the adjoint

solver

In this section, we summarize the steps involved in one iteration of the adjoint prob-

lem. First, a random velocity field is initialized in MATLAB while satisfying the

feasibility constraints. This initial condition is used to solve the direct problem for-

ward in time for a given time window using Nek5000. The final state of the direct

problem can used to solve for the initial state of the adjoint problem using equation

2.13. The adjoint problem can be solved backward in time to reach the final state.

This gives all the parameters required to calculate the gradient of the Lagrangian

with respect to the initial velocity field. We update the initial velocity using equation

2.14 and this process can be continued until the solution converges.

Figure 2-1: Layout of the loop to be completed in each iteration of the adjoint problemtill the solution converges

2.2 Statistical analysis and prediction of extreme events

In this section, we define the extreme event indicator πœ†, given as

πœ†(𝑑) =βŸ¨βˆ’ , 𝑒*

0 βˆ’ βŸ©βˆ’

𝑒*0 βˆ’

(2.15)

25

where is the velocity field snapshot at time 𝑑, is the mean of the velocity field

over the entire time domain, and 𝑒*0 is the velocity field of the precursor.

Here, we can see that the indicator value will be large when the velocity field snapshot

approaches the precursor velocity field. As the precursor is the cause of the extreme

events, the kinetic energy should also take a large value when the indicator does.

Hence, both the indicator and the kinetic energy would have extreme events almost

simultaneously. But given that we take a time window, the extreme event in the

indicator would precede the extreme event in the kinetic energy and that is how there

will be an early warning signal before the extreme event in the kinetic energy takes

place.

2.2.1 Joint and conditional statistics of the extreme events

The goal of this subsection is to quantify the prediction capabilities of the obtained

precursor from the adjoint-based optimization. For the quantitative analysis, we use

the joint and conditional statistics between the indicator πœ† and the extreme event

observable i.e. the kinetic energy.

Here, for a given random variable 𝑋𝑑, the goal is to find an indicator variable π‘Œπ‘‘

(random variable), which can predict or generate a signal for an extreme event about

to occur in 𝑋𝑑 in a given time frame. An extreme event in 𝑋𝑑 is defined as instant

where the value of 𝑋𝑑 crosses a given extreme event threshold π‘₯𝑒. For the random

variable 𝑋𝑑, the maximum value over a time interval [𝑑 + 𝑑0, 𝑑 + 𝑑0 + βˆ†π‘‘] for some 𝑑0,

and βˆ†π‘‘ β‰₯ 0, is

𝑑(𝑑0,βˆ†π‘‘) = max π‘‹π‘ π‘ βˆˆ[𝑑+𝑑0,𝑑+𝑑0+Δ𝑑]

(2.16)

Here, 𝑑(𝑑0,βˆ†π‘‘) is a new random variable and is a function of 𝑑0 and βˆ†π‘‘. This variable

is a measure of the maximum value observed in the variable 𝑋𝑑 over a future time

interval defined by 𝑠. The joint probability distribution of (𝑑, π‘Œπ‘‘) can be defined as

-

𝐹𝑋𝑑,π‘Œπ‘‘(π‘₯, 𝑦) = 𝑃 (𝑑 ≀ π‘₯, π‘Œπ‘‘ ≀ 𝑦) =

∫ π‘₯

βˆ’βˆž

∫ 𝑦

βˆ’βˆžπ‘π‘‹π‘‘,π‘Œπ‘‘

(π‘₯, 𝑦) 𝑑π‘₯ 𝑑𝑦 (2.17)

26

In the equation above, 𝑝𝑋𝑑,π‘Œπ‘‘is the probability density function (pdf) that is associated

with the probability distribution 𝐹𝑋𝑑,π‘Œπ‘‘. The instant where π‘Œπ‘‘ = 𝑦 and 𝑑 = π‘₯ at time

𝑑, is measured by the probability density as a function of π‘₯ and 𝑦. The conditional

probability of 𝑑 = π‘₯ given that π‘Œπ‘‘ = 𝑦 can be defined using the Bayes’ formula given

as

𝑝𝑋𝑑|π‘Œπ‘‘=

𝑝𝑋𝑑,π‘Œπ‘‘

π‘π‘Œπ‘‘

(2.18)

where π‘π‘Œπ‘‘ is the probability density of π‘Œπ‘‘. In the time interval 𝑠, the conditional

probability density function 𝑝𝑋𝑑|π‘Œπ‘‘indicates the predictive capability of the indicator

π‘Œπ‘‘ for extreme events in 𝑋𝑑. The probability of upcoming extreme events in 𝑋𝑑 for a

given threshold π‘₯𝑒 given the current value of π‘Œπ‘‘ at 𝑦

𝑃𝑒𝑒(𝑦) =

∫ ∞

π‘₯

𝑝𝑋𝑑,π‘Œπ‘‘(π‘₯, 𝑦) 𝑑π‘₯ (2.19)

Here, a large value of the indicator should indicate a higher probability of the oc-

curence of an extreme event in 𝑋𝑑. Hence, the function 𝑃𝑒𝑒 should have monotonic

characteristics. For a very low value of the indicator, the probability of an extreme

event should be very low as well (close to zero), and for very high values of the indica-

tor, close to one. For a threshold on the probability, we classify an event as extreme

only if 𝑃𝑒𝑒(𝑦𝑒) = 0.5. We can define various scenarios based on the thresholds π‘₯𝑒 and

𝑦𝑒 as

Correct Rejections (CR): 𝑑 < π‘₯𝑒 given π‘Œπ‘‘ < 𝑦𝑒

Correct Predictions (CP): 𝑑 > π‘₯𝑒 given π‘Œπ‘‘ > 𝑦𝑒

False Negatives (FN): 𝑑 > π‘₯𝑒 given π‘Œπ‘‘ < 𝑦𝑒

False Positives (FP): 𝑑 < π‘₯𝑒 given π‘Œπ‘‘ > 𝑦𝑒

From these prediction parameters, the indicator accuracy or prediction capabilities

can be defined as

27

Rate of successful predictions = 𝐢𝑃𝐢𝑃+𝐹𝑁

=

∫ ∞

π‘₯𝑒

∫ ∞

𝑦𝑒

𝑝𝑋𝑑,π‘Œπ‘‘(π‘₯, 𝑦) 𝑑π‘₯ π‘‘π‘¦βˆ« ∞

π‘₯𝑒

∫ ∞

βˆ’βˆžπ‘π‘‹π‘‘,π‘Œπ‘‘

(π‘₯, 𝑦) 𝑑π‘₯ 𝑑𝑦

Rate of successful rejections = 𝐢𝑅𝐢𝑅+𝐹𝑃

=

∫ π‘₯𝑒

βˆ’βˆž

∫ 𝑦𝑒

βˆ’βˆžπ‘π‘‹π‘‘,π‘Œπ‘‘

(π‘₯, 𝑦) 𝑑π‘₯ π‘‘π‘¦βˆ« π‘₯𝑒

βˆ’βˆž

∫ ∞

βˆ’βˆžπ‘π‘‹π‘‘,π‘Œπ‘‘

(π‘₯, 𝑦) 𝑑π‘₯ 𝑑𝑦

A good indicator would be one that has a higher number of correct predictions and

correct rejections, and a low amount of false positives and false negatives. Hence,

these are the desirable characteristics of a good indicator.

2.3 An overview of Nek5000 (flow solver)

As our fluid flow solver, we use Nek5000 which is a FORTRAN and C based solver

that can handle two- and three-dimensional domains that can be discretized into quad

or hex elements. It is an open source spectral element solver for computational fluid

dynamics simulations. It can be used to simulate unsteady incompressible fluid flows

with thermal and passive scalar transport. It is a code that is based on time-stepping

and supports steady Navier Stokes and heat conduction based problems.

It consists of three principal modules: the pre-processor π‘π‘Ÿπ‘’π‘₯, the solver π‘›π‘’π‘˜5000,

and the post-processor π‘π‘œπ‘ π‘‘π‘₯. Nek5000 uses MPI for message passing and some

LAPACK routines for the computation of eigenvalues. Some studies using Nek5000

can be seen in [74], [50], [62], [39], [34], [66].

28

Chapter 3

Applications and results

The goal of this chapter is to apply the adjoint framework as described in chapter

2 to various fluid flow systems. In general, the framework can be applied to any

dynamical system where a set of governing equations and constraints are known. An

observable is then chosen to begin the analysis for the extreme event prediction. Here,

we first consider a 2-D flow namely the Kolmogorov flow. The governing equations

are first described with the boundary conditions. A simulation is performed for 3000

time steps and the proper orthogonal decomposition is performed which provides us

with the POD modes. A specific number of modes are chosen which capture more

than 90% of the energy of the system. This helps us in reducing the dimensionality

of the problem as well as retain a major portion of the information. An analysis is

performed to determine the order of the constraint π‘Ÿ0 based on the snapshots that

were obtained from the simulation.

The adjoint setup with the governing and the constraints to be satisfied can then

be run. The solver iterates until it converges to a precursor and the order of the

magnitude of the Lagrangian tends to zero. The probability distributions for the cho-

sen parameters πœ† (the indicator) and π·π‘š (the extreme event in the observable in a

time horizon) are plotted with a time lag to see if the indicator jump above a certain

threshold precedes the extreme event in the observable. The rate of successful predic-

tions and successful rejections are presented which show the prediction capabilities of

the indicator πœ†.

29

A similar analysis is repeated for the 2-D airfoil at an angle of attack of 30 de-

grees case and results are presented. A slight difference in this case is that the inlet

velocity is used as a source of stochasticity. Choosing a critical Reynolds number

after which the flow becomes unstable is described. This is done by zooming in at

the lift coefficient time series plots. Once the peaks start growing, an unstable flow is

indicated and this would be the critical Reynolds number of the flow. To generate the

stochastic inlet velocity, the Karhunen-Loeve expansion on the signal is performed

for smoothing purposes and reducing the order of the signal. The mean is taken as

the critical Reynolds number and the standard deviation is based on the amplitude

of the variation around the mean Reynolds number.

3.1 Kolmogorov flow

The Kolmogorov flow is a two-dimensional turbulent flow and has been an important

plant to study the nature of turbulent flows in general. The flow state is governed by

the 2-D Navier-Stokes equations which have a monochromatic forcing. This forcing

is sinusoidal in nature and is assumed to be stationary. The source of stochasticity

in the Kolmogorov flow comes from the inherent turbulence. We analyze the results

for a Reynolds number such that the flow has turbulence. Stochasticity in the system

causes intermittent bursts in the kinetic energy that we will see later in the section.

3.1.1 Flow solver, setup, and simulation

The governing equations for the Kolmogorov flow are given below

πœ•π‘‘ = βˆ’.βˆ‡βˆ’βˆ‡π‘ + πœˆβˆ† + 𝑓 (3.1)

and

βˆ‡. = 0 (3.2)

There is a forcing term involved in the π‘₯ direction given by 𝑓 = 𝑠𝑖𝑛(π‘˜π‘“π‘¦)𝑒π‘₯ where

π‘˜π‘“ is the wave forcing number. π‘˜π‘“ is chosen to be 4, as with this and 𝑅𝑒 = 40, the

30

flow will have inherent chaos and intermittent bursts in the kinetic energy. This is

shown in figure 3-2. The boundary conditions for the flow are assumed to be periodic

both in the π‘₯ and 𝑦 directions. The mesh (figure 3-1) for the Kolmogorov flow is

shown below. It contains 256 elements and order of the polynomial for the flow solver

is 7. The Reynolds number - 𝑅𝑒 is taken to be 40 which ensures turbulence. The

direct numerical simulation is performed for a total of 3000 convective time units

using Nek5000. Figure 3-2 shows the kinetic energy of the system plot vs. time.

The intermittent bursts can be seen which would be considered as the extreme events

in the observable 𝐼. The goal of the precursor and the extreme event indicator πœ†

would be to predict these bursts before a certain time frame so that we can have

some warning signals.

Figure 3-1: 2-D mesh for the Kolmogorov flow.

3.1.2 POD and extreme event analysis results

The goal of this subsection is to show the proper orthogonal decompositions results,

particularly the POD modes and the associated coefficients, the obtained precursor

and the prediction capabilities of the extreme event indicator πœ†. For the analysis, the

first 500 snapshots of the simulation are neglected so that the transient states can be

removed as they contain random spikes that need to die down. The initial velocity

31

0 500 1000 1500 2000 2500 3000t

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9K

E

Figure 3-2: Time series plot of the kinetic energy.

field is initialized randomly based on the number 𝑛 of the POD modes subspace. This

initial condition should lie on the attractor and that is ensured by making it satisfy the

feasibility constraint that makes the guess lie on the attractor. 10 degrees of freedom

are used for the POD coefficients as they capture more than 90% of the energy. The

vorticity snapshots of the POD modes are shown in figure 3-3 and the associated POD

coefficients are shown in figure 3-4. The first mode shows the predominant feature

of the flow and captures the most energy. The modes are arranged in an order of

decreasing energy, and the first 10 modes cumulatively sum up to 90% energy.

The randomly initialized velocity field is shown in figure 3-5, which is used as

the input for the solver for the first iteration. As the solver iterates, this field gets

updated based on the lagrangian until it converges to the precursor. After around

1500 iterations, the field converges to the required local minima velocity field which

would be the precursor as shown in figure 3-6. The time horizon used to evolve the

adjoint problem per iteration is 2e-5 time units (time step size is 2e-6) as this is the

instantaneous time horizon to make 𝑑 tend to 0. The learning rate is chosen to be

𝛼 = 0.05.

32

πœ”1 πœ”6

πœ”2 πœ”7

πœ”3 πœ”8

πœ”4 πœ”9

πœ”5 πœ”10

Figure 3-3: Vorticity (the scale is -2 to 2) of the POD modes 𝑖 = βˆ‡ Γ— 𝑖(),𝑖 = 1, . . . , 10 for the Kolmogorov flow.

Figure 3-7 shows the indicator πœ† and π·π‘š (the max of the KE in each 1 time unit

window) on the same figure. It can be seen that the bursts in both the variables are

concurrent. Also, due to the time lag between πœ† and π·π‘š, the extreme events in π·π‘š

can be predicted beforehand, by seeking states when πœ† crosses a certain threshold.

It can be observed that the extreme event in πœ† occurs before the extreme event in

KE, hence giving an early warning signal. See figure 3-8. A required property of

the extreme event indicator πœ† is to maximize the number of correct predictions and

correct rejections. The results for these based on the formulas as derived earlier are

33

500 1000 1500 2000 2500t

-5

0

5a 1

500 1000 1500 2000 2500t

-3

-2

-1

0

1

2

a 6

500 1000 1500 2000 2500t

-5

0

5

a 2

500 1000 1500 2000 2500t

-2

-1

0

1

a 7

500 1000 1500 2000 2500t

-6

-4

-2

0

2

4

a 3

500 1000 1500 2000 2500t

-2

-1

0

1

2

3

a 8

500 1000 1500 2000 2500t

-4

-2

0

2

4

6

a 4

500 1000 1500 2000 2500t

-2

-1

0

1

2

a 9

500 1000 1500 2000 2500t

-1

-0.5

0

0.5

1

1.5

a 5

500 1000 1500 2000 2500t

-3

-2

-1

0

1

2

a 10

Figure 3-4: The first 10 POD coefficients for the Kolmogorov flow.

34

Figure 3-5: Initial guess of the velocity field.

Figure 3-6: Precursor obtained after 1500 iterations of the adjoint problem.

as follows -

Rate of successful predictions = 90.59%

Rate of successful rejections = 96.93%

We take the extreme event threshold π‘₯𝑒 to be = mean(KE) + 2*std(KE)

πœ†π‘’ is computed by setting 𝑃𝑒𝑒(πœ†π‘’) = 0.5

The probability distribution function plots as shown in Figures 3-9 and 3-10, describe

the appromixate linearity between the values of πœ† and π·π‘š. Small values of π·π‘š and

πœ† are highly populated, which show the local minimums being predicted successfully.

35

500 1000 1500 2000 2500Snapshot #

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

KE

Figure 3-7: πœ† and Kinetic energy (KE) vs. snapshot number (the time lag between πœ†and KE is 1 time unit).

2025 2030 2035 2040 2045 2050 2055 2060Snapshot #

0.21

0.22

0.23

0.24

0.25

0.26

0.27

0.28

0.29

0.3

0.31

Figure 3-8: Zoomed peak near t = 2040 for the πœ† and Kinetic energy vs. snapshotnumber plot.

36

0 0.1 0.2 0.3 0.4D

m

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

Figure 3-9: Plot showing correlation between πœ† and π·π‘š (the time lag between πœ† andKE is 1 time unit).

Figure 3-10: Joint pdf of πœ† and π·π‘š (the time lag between πœ† and KE is 1 time unit).

High values of π·π‘š and πœ† are low in population, which show the global minimums or

the extreme events being predicted successfully. A bad indicator would be one which

37

Figure 3-11: Conditional pdf of πœ† and π·π‘š (the time lag between πœ† and KE is 1 timeunit).

has data points all over the joint or conditional probability distribution function plots.

A simulation for 10,000 time units is also performed and the calculations are made

using the precursor obtained previously. From calculations, the extreme events re-

sults are as follows -

Rate of successful predictions = 96.14%

Rate of successful rejections = 94.86%

The extreme events in the dissipative energy (DE) can also be predicted with a

good accuracy by using the same precursor. The results are as follows -

Rate of successful predictions = 80.71%

Rate of successful rejections = 83.93%

38

1000 2000 3000 4000 5000 6000 7000 8000 9000Snapshot #

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

KE

Figure 3-12: πœ† and Kinetic energy vs. snapshot number (10000 time units simulation,the time lag between πœ† and KE is 1 time unit).

0 0.1 0.2 0.3 0.4 0.5D

m

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

Figure 3-13: Plot showing correlation between πœ† andπ·π‘š (10000 time units simulation,the time lag between πœ† and KE is 1 time unit).

39

Figure 3-14: Joint pdf of πœ† andπ·π‘š (10000 time units simulation, the time lag betweenπœ† and KE is 1 time unit).

Figure 3-15: Conditional pdf of πœ† and π·π‘š (10000 time units simulation, the time lagbetween πœ† and KE is 1 time unit).

40

1000 2000 3000 4000 5000 6000 7000 8000 9000Snapshot #

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

KE

Figure 3-16: πœ† and Dissipative energy (DE) vs. snapshot number (10000 time unitssimulation, the time lag between πœ† and DE is 1 time unit).

0.4 0.5 0.6 0.7 0.8 0.9 1D

m

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

Figure 3-17: Plot showing correlation between πœ† and π·π‘š (DE) (10000 time unitssimulation, the time lag between πœ† and DE is 1 time unit).

41

Figure 3-18: Joint pdf of πœ† and π·π‘š (DE) (10000 time units simulation, the time lagbetween πœ† and DE is 1 time unit).

Figure 3-19: Conditional pdf of πœ† and π·π‘š (DE) (10000 time units simulation, thetime lag between πœ† and DE is 1 time unit).

42

3.2 2-D Airfoil flow

In this section, we take the fluid flow over a 2-D airfoil at a certain angle of attack.

The source of stochasticity in this case is the stochastic inlet velocity that is generated

using a stochastic differential equation. The inlet velocity signal is further smoothed

out using the Karhuhen-Loeve Expansion (KLE). This is done to preserve the dynam-

ics and also capture the most important features in the signal. A critical Reynolds

number is established on which the inlet velocity is based using a certain amplitude.

For the original signal and the one generated by KLE to have similar statistics, we

keep the mean and standard deviation of the KLE signal close enough to the original

signal through running multiple iterations of random signals. The random signal that

has similar statistics under a small margin of error is chosen as the inlet velocity.

3.2.1 Flow solver, setup, and simulation

For the setup, we consider the NACA-0012 airfoil with an angle of attack of 30∘.

The mesh (figure 3-20) for the given geometry is shown below (618 elements and

polynomial order of 9). A constant velocity profile (π‘ˆπ‘₯ = 1) characterized by the

Reynolds number is given at the inlet. We initialize the flow field with a low amplitude

disturbance (π‘ˆπ‘₯ = 1 + 10βˆ’5𝑦). The Reynolds number in terms of the dimensional

parameters is 𝑅𝑒 = 𝜌 π‘ˆπΏπœ‡

.

We perform a DNS simulation for the 2D airfoil flow using Nek5000. The simu-

lation is performed for a total of 3000 convective time units. The critical Reynolds

number (𝑅𝑒𝑐) is defined to be the one from which the flow starts becoming un-

steady i.e. doesn’t reach a fixed steady state solution. Vorticity snapshots for various

Reynolds numbers near the 𝑅𝑒𝑐 at 100 time units instance are given in figure 3-21

below, part (a) shows the converged steady state solution at a Reynolds number be-

low the critical Reynolds number, part (b) shows the beginning of the unsteady state

of the flow, above this Reynolds number, all solutions would be unsteady (see part(c)

as an example).

43

Figure 3-20: 2-D mesh of the flow around the airfoil.

3.2.2 Lift coefficient over time analysis to determine the crit-

ical Reynolds number

To study the unsteadiness in the flow and establish a critical Reynolds number, we

consider the lift coefficient over time. Figure 3-24 displays the lift coefficient over

time plots for the three Reynolds numbers as in figure 3-21. 𝑅𝑒 = 90, decays at a very

rapid rate and the oscillations become negligible in magnitude after 100 convective

time units. On zooming the regions, it can be observed that the peaks are decreasing

with time which means there will be a point where steady state would be established.

For 𝑅𝑒 = 105, the decay is at a very slow rate over time (as seen when zooming the

peaks in the plots). From T=280, it can be seen that the peaks start to stay where

they are in magnitude. This shows there will be oscillations with some magnitude

44

that won’t die out over time. Whereas, for 𝑅𝑒 = 110, a chaotic behavior can be

observed which marks the onset of a transition. The peaks start to show oscillations

in themselves, meaning that there will be continued transience in the flow state for

all time. The critical Reynolds number (𝑅𝑒𝑐) can hence be said to be between 105

and 110 (𝑅𝑒𝑐 = 107.5 Β± 2.5).

3.2.3 Generating the stochastic inlet velocity and mode re-

duction using Karhunen-Loeve expansion

The inlet velocity signal (π‘ˆ) that will be the source of stochasticity in the fluid flow

dynamics is generated by simulating the stochastic differential equation as given below

-

π‘‘π‘ˆ = βˆ’πœ† π‘ˆ 𝑑𝑑 + πœ‡ π‘‘π‘Š (3.3)

with the parameters given as

π‘ˆ0 = π‘ˆ(1) = 0

𝑁 = 2000000 : Number of time steps in our fluid simulation

𝑑𝑑 = 1/𝑁

π‘‘π‘Šπ‘– = π‘§π‘–βˆšπ‘‘π‘‘ :

where 𝑧𝑖 is a random variable chosen from a normal distribution with mean = 0 and

std = 1

Hence, to simulate it numerically we write it as simply a finite time difference and

propagating forward in time

π‘ˆπ‘–+1 = π‘ˆπ‘– βˆ’ πœ† π‘ˆπ‘– 𝑑𝑑 + πœ‡ π‘‘π‘Šπ‘–

If this signal is used by itself, then it gives rise to a noisy lift coefficient time se-

ries. This is due to the fact that the flow is non-symmetric and changing the inlet

45

velocity rapidly over time would make the flow field chaotic. Hence, there is a need

to reduce the order of the stochastic inlet velocity signal.

We generate the long time scale fluctuations and reduce the number of modes to

smooth out the signal using KLE. In essence, the Karhunen-Loeve expansion theorm

(KLE) represents a stochastic process as a linear combination of an infinite number

of orthogonal functions. This is similar to the principal component analysis (PCA)

and the fourier series expansion of a given function. A more detailed explanation can

be seen in [82]. The fluctuating inlet velocity using KLE is given by -

π‘ˆ(𝜏) =βˆ‘π‘–β‰₯1

βˆšπœ†π‘– 𝑒𝑖(𝜏) 𝜈

where 𝜈 is a normally distributed random variable. To obtain the eigenvalues (πœ†π‘–)

and eigenvectors (𝑒𝑖), we solve the eigenvalue problem given by

∫Ω

𝐢(𝜏, 𝜏 β€²) 𝑒𝑖(πœβ€²) π‘‘πœ = πœ†π‘– 𝑒𝑖(𝜏)

where 𝐢(𝜏, 𝜏 β€²) is the correlation function of the process π‘ˆ(𝜏)

We choose the number of modes by satisfying the minimum k condition -

π‘˜βˆ‘π‘–=1

πœ†π‘– β‰₯ 0.99π‘›βˆ‘

𝑖=1

πœ†π‘–

To formulate the correlation matrix, we use the MATLAB snippet -

for i=1:length(π‘ˆπ‘)

for j=1:length(π‘ˆπ‘)

diff = abs((i-j)*0.5);

C(i,j) = variance * exp(-diff/b);

end

end

46

Probe # x y zProbe 1 1.5 0.5 0.0Probe 2 3.0 0.2 0.0Probe 3 5.0 0.0 0.0

Table 3.1: Location of the three probes.

where 𝑏 is the correlation length taken to be 200 (large enough than the strouhal

period of 5). This has been chosen out of b=20, 50, 100 and 200. The inlet velocity

signal is generated with a mean of Reynolds number = 100 and amplitude = 20. The

signal can be seen in figure 3-23.

3.2.4 Results

The goal of this section is to show the proper orthogonal decomposition results and

the extreme events happening in the kinetic energy and the lift coefficient. The first

ten POD modes and the corresponding POD coefficients are displayed in the figures

3-25 and 3-26 respectively. They show the modes containing the most dominant

features in the flow simulation in decreasing order of energy.

The extreme events in the lift coefficient (𝐢𝐿) can be seen in figure 3-27. The

fluctuations in the lift coefficient has a strong correlation with the inlet velocity as

can be seen in part(c), extreme events in the inlet velocity lead to extreme events

in the lift coefficient most of the times. Here, we could refer to an extreme event

as a significant jump or drop from the mean of the lift coefficient time series. We

also consider the x (figure 3-28) and y (figure 3-29) components of the velocity at 3

probes with their locations given as (see table 3.1)

47

(a)

(b)

(c)

Figure 3-21: Vorticity snapshots at T=100 for Reynolds numbers near 𝑅𝑒𝑐: (a) 𝑅𝑒= 90 (b) 𝑅𝑒 = 105 (c) 𝑅𝑒 = 110.

48

0 500 1000 1500 2000 2500 3000t

30

35

40

45

50

55

60K

E

Figure 3-22: The kinetic energy plot for 3000 time units.

0 500 1000 1500 2000 2500 3000t

80

85

90

95

100

105

110

115

120

Re

KLE, b = 200, std = 6.7261, std-formula=6.67, mean=101.7132, mean-base=100

Figure 3-23: The inlet velocity input signal.

49

(a) (b)

0 50 100 150 200t

0.9

0.95

1

1.05

1.1

1.15

1.2C

L

140 150 160 170 180 190 200t

0.91955

0.9196

0.91965

0.9197

0.91975

CL

(c) (d)

0 50 100 150 200 250 300t

0.85

0.9

0.95

1

1.05

1.1

1.15

CL

220 240 260 280 300t

0.935686302

0.935686304

0.935686306

0.935686308

0.93568631

0.935686312

0.935686314

0.935686316

0.935686318

0.93568632

(e) (f)

0 50 100 150 200 250 300t

0.85

0.9

0.95

1

1.05

1.1

1.15

CL

200 220 240 260 280t

0.948027234

0.948027236

0.948027238

0.94802724

0.948027242

0.948027244

Figure 3-24: (a) Re = 90 for t = 15 to 200 (b) Re = 90, zoomed peaks (c) Re =105, for t = 15 to 300 (d) Re = 105, zoomed peaks (e) Re = 110, for t = 15 to 300(f) Re = 110, zoomed peaks.

50

πœ”1 πœ”6

πœ”2 πœ”7

πœ”3 πœ”8

πœ”4 πœ”9

πœ”5 πœ”10

Figure 3-25: Vorticity (the scale is -2 to 2) of the POD modes 𝑖 = βˆ‡ Γ— 𝑖(),𝑖 = 1, . . . , 10 for the airfoil flow.

51

500 1000 1500 2000 2500t

-1

0

1

a 1

500 1000 1500 2000 2500t

-0.1

-0.05

0

0.05

0.1

a 6

500 1000 1500 2000 2500t

-0.5

0

0.5

a 2

500 1000 1500 2000 2500t

-0.1

-0.05

0

0.05

0.1

a 7

500 1000 1500 2000 2500t

-0.5

0

0.5

a 3

500 1000 1500 2000 2500t

-0.05

0

0.05

a 8

500 1000 1500 2000 2500t

-0.1

0

0.1

0.2

a 4

500 1000 1500 2000 2500t

-0.02

0

0.02

0.04

a 9

500 1000 1500 2000 2500t

-0.1

-0.05

0

0.05

0.1

a 5

500 1000 1500 2000 2500t

-0.04

-0.02

0

0.02

0.04

a 10

Figure 3-26: The first 10 POD coefficients for the airfoil flow.

52

(a)

500 1000 1500 2000 2500t

0.6

0.7

0.8

0.9

1

1.1

1.2

1.3

1.4

CL

CL

mean(CL

)

(b)

0 500 1000 1500 2000 2500 3000t

0.6

0.7

0.8

0.9

1

1.1

1.2

1.3

mea

n(C

L)

(c)

500 1000 1500 2000 2500t

-0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

0.2

CL

fluc(CL

)

U

Figure 3-27: (a) 𝐢𝐿 from t = 20 to 3000 (b) The averaged curve (c) fluctuations in𝐢𝐿 calculated by removing the moving average (5 time units moving window). π‘ˆ istaken as the inlet KLE velocity with zero mean. From t = 20 to 3000.

53

(a)

500 1000 1500 2000 2500t

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

Vx1

(b)

500 1000 1500 2000 2500t

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

0.55

Vx2

(c)

500 1000 1500 2000 2500t

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Vx3

Figure 3-28: π‘₯ component of the velocities at the three probes (a) 𝑉π‘₯1 (b) 𝑉π‘₯2 (c) 𝑉π‘₯3.

54

(a)

500 1000 1500 2000 2500t

-0.2

-0.15

-0.1

-0.05

0

0.05

0.1

Vy1

(b)

500 1000 1500 2000 2500t

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Vy2

(c)

500 1000 1500 2000 2500t

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

Vy3

Figure 3-29: 𝑦 component of the velocities at the three probes (a) 𝑉𝑦1 (b) 𝑉𝑦2 (c) 𝑉𝑦3.

55

56

Chapter 4

Conclusions and future work

4.1 Summary of the results and discussion

In this thesis, we have derived the adjoint equations for the Navier-stokes equations of

a 2-D fluid flow. The extreme events are considered in the observable taken as kinetic

energy of the flow field. We have presented a coupled fluid (Nek5000) and adjoint

solver (MATLAB) that can be applied to any dynamical system with non-linear terms

in the flow map and subjected to constraints. The prediction capabilities of the solver

are demonstrated for the 2-D Kolmogorov flow, with the results summarized in the

table below 4.1.

The extreme events for the kinetic energy and the lift coefficient for a 2-D airfoil

flow at an angle of attack of 30 degrees are also presented. The extreme events bear

a strong correlation with the extreme events in the stochastic inlet velocity, which is

considered to be the source of the extreme events in this case.

Simulation Rate of successful predictions Rate of successful rejections3000 time units 90.59% 96.93%10000 time units 96.14% 94.86%

Table 4.1: Summary of the extreme events prediction results for the Kolmogorov flow

57

4.2 Future work

Future work would include the application of this coupled solver on the 2-D airfoil

flow presented in chapter 3. The goal would be to evaluate the prediction capabilities

of the obtained precursor and the time window in which it can predict extreme events

in the kinetic energy effectively. Another interesting case would be the 3-D Ahmed

body flow ([64], [11], [80], [37], [55], [4]), where the source of stochasticity would

be turbulence. The prediction capabilities can also be challenged by increasing the

time window in which the extreme event has to be predicted. This would expose the

maximum time before which the extreme events could be predicted without a large

decline in the prediction accuracy of the precursor.

58

Bibliography

[1] S. Albeverio, V. Jentsch, and H. Kantz. Extreme events in nature and society.Springer Science & Business Media, 2006.

[2] S. Albeverio and V. Piterbarg. Mathematical methods and concepts for theanalysis of extreme events. In Extreme events in nature and society, pages 47–68. Springer, 2006.

[3] H Babaee. An observation-driven time-dependent basis for a reduced de-scription of transient stochastic systems. Proceedings of the Royal Society A,475(2231):20190506, 2019.

[4] I. Bayraktar, D. Landman, and O. Baysal. Experimental and computationalinvestigation of ahmed body for ground vehicle aerodynamics. Technical report,SAE Technical Paper, 2001.

[5] G. Berkooz, P. Holmes, and J. L. Lumley. The proper orthogonal decompositionin the analysis of turbulent flows. Annual review of fluid mechanics, 25(1):539–575, 1993.

[6] A. Blanchard, S. Mowlavi, and T. P. Sapsis. Control of linear instabilities by dy-namically consistent order reduction on optimally time-dependent modes. Non-linear Dynamics, 95(4):2745–2764, 2019.

[7] A. Blanchard and T. P. Sapsis. Analytical description of optimally time-dependent modes for reduced-order modeling of transient instabilities. SIAMJournal on Applied Dynamical Systems, 18(2):1143–1162, 2019.

[8] A. Blanchard and T. P. Sapsis. Stabilization of unsteady flows by reduced-order control with optimally time-dependent modes. Physical Review Fluids,4(5):053902, 2019.

[9] P. J. Blonigan, M. Farazmand, and T. P. Sapsis. Are extreme dissipation eventspredictable in turbulent fluid flows? Physical Review Fluids, 4(4):044606, 2019.

[10] L. Bottou. Large-scale machine learning with stochastic gradient descent. InProceedings of COMPSTAT’2010, pages 177–186. Springer, 2010.

59

[11] C. Bruneau, E. CreusΓ©, D. Depeyras, P. GilliΓ©ron, and I. Mortazavi. Couplingactive and passive techniques to control the flow past the square back ahmedbody. Computers & Fluids, 39(10):1875–1892, 2010.

[12] A. BΓΌcher and J. Segers. Extreme value copula estimation based on block maximaof a multivariate stationary time series. Extremes, 17(3):495–528, 2014.

[13] A. Bunde, J. F. Eichner, S. Havlin, and J. W. Kantelhardt. The effect of long-term correlations on the return periods of rare events. Physica A: StatisticalMechanics and its Applications, 330(1-2):1–7, 2003.

[14] C. Burges, T. Shaked, E. Renshaw, A. Lazier, M. Deeds, N. Hamilton, and G. N.Hullender. Learning to rank using gradient descent. In Proceedings of the 22ndInternational Conference on Machine learning (ICML-05), pages 89–96, 2005.

[15] Y. Cao, S. Li, L. Petzold, and R. Serban. Adjoint sensitivity analysis fordifferential-algebraic equations: The adjoint dae system and its numerical so-lution. SIAM journal on scientific computing, 24(3):1076–1089, 2003.

[16] E. Castillo. Extreme value theory in engineering. Elsevier, 2012.

[17] M. A. Celia, T. F. Russell, I. Herrera, and R. E. Ewing. An eulerian-lagrangianlocalized adjoint method for the advection-diffusion equation. Advances in WaterResources, 13(4):187–206, 1990.

[18] A. Chatterjee. An introduction to the proper orthogonal decomposition. Currentscience, pages 808–817, 2000.

[19] V. A. Cicirello and S. F. Smith. Heuristic selection for stochastic search opti-mization: Modeling solution quality by extreme value theory. In InternationalConference on Principles and Practice of Constraint Programming, pages 197–211. Springer, 2004.

[20] P. Coles and J. D. Barrow. Non-gaussian statistics and the microwave backgroundradiation. Monthly Notices of the Royal Astronomical Society, 228(2):407–426,1987.

[21] L. De Haan and A. Ferreira. Extreme value theory: an introduction. SpringerScience & Business Media, 2007.

[22] L. De Haan and S. Resnick. On regular variation of probability densities. Stochas-tic processes and their applications, 25:83–93, 1987.

[23] R. G. Dean. Freak waves: a possible explanation. In Water wave kinematics,pages 609–612. Springer, 1990.

[24] R. Devaney. An introduction to chaotic dynamical systems. CRC Press, 2018.

[25] R. L. Devaney. A first course in chaotic dynamical systems: theory and experi-ment. CRC Press, 2018.

60

[26] K. Dysthe, H. E. Krogstad, and P. MΓΌller. Oceanic rogue waves. Annu. Rev.Fluid Mech., 40:287–310, 2008.

[27] J. H. J. Einmahl, J. Segers, et al. Maximum empirical likelihood estimation ofthe spectral measure of an extreme-value distribution. The Annals of Statistics,37(5B):2953–2989, 2009.

[28] M. Farazmand. An adjoint-based approach for finding invariant solutions ofnavier–stokes equations. Journal of Fluid Mechanics, 795:278–312, 2016.

[29] M. Farazmand and T. P. Sapsis. A variational approach to probing extremeevents in turbulent dynamical systems. Science advances, 3(9):e1701533, 2017.

[30] V. FernΓ‘ndez. Risk management under extreme events. International Review ofFinancial Analysis, 14(2):113–148, 2005.

[31] P. H. Franses and D. Van Dijk. Forecasting stock market volatility using (non-linear) garch models. Journal of Forecasting, 15(3):229–235, 1996.

[32] E. M. Furrer and R. W. Katz. Improving the simulation of extreme precipitationevents by stochastic weather generators. Water Resources Research, 44(12), 2008.

[33] J. Gleick. Chaos: Making a new science. Open Road Media, 2011.

[34] J. Gong, S. Markidis, M. Schliephake, E. Laure, D. Henningson, P. Schlatter,A. Peplinski, A. Hart, J. Doleschal, D. Henty, et al. Nek5000 with openacc. InInternational Conference on Exascale Applications and Software, pages 57–68.Springer, 2014.

[35] J. D. Hamilton. Time series analysis, volume 2. Princeton university pressPrinceton, NJ, 1994.

[36] R. W. Healy and T. F. Russell. A finite-volume eulerian-lagrangian localized ad-joint method for solution of the advection-dispersion equation. Water ResourcesResearch, 29(7):2399–2413, 1993.

[37] C. Hinterberger, M. Garcia-Villalba, and W. Rodi. Large eddy simulation of flowaround the ahmed body. In The aerodynamics of heavy vehicles: trucks, buses,and trains, pages 77–87. Springer, 2004.

[38] E. Hopf. A mathematical example displaying features of turbulence. Communi-cations on Pure and Applied Mathematics, 1(4):303–322, 1948.

[39] I. Ivanov, J. Gong, D. Akhmetova, I. B. Peng, S. Markidis, E. Laure, R. Machado,M. Rahn, V. Bartsch, A. Hart, et al. Evaluation of parallel communicationmodels in nekbone, a nek5000 mini-application. In 2015 IEEE InternationalConference on Cluster Computing, pages 760–767. IEEE, 2015.

[40] P. A. E. M. Janssen. Nonlinear four-wave interactions and freak waves. Journalof Physical Oceanography, 33(4):863–884, 2003.

61

[41] R. W. Katz, M. B. Parlange, and P. Naveau. Statistics of extremes in hydrology.Advances in water resources, 25(8-12):1287–1304, 2002.

[42] S. H. Kellert. In the wake of chaos: Unpredictable order in dynamical systems.University of Chicago press, 1993.

[43] H. Kim, D. Sasaki, S. Obayashi, and K. Nakahashi. Aerodynamic optimizationof supersonic transport wing using unstructured adjoint method. AIAA journal,39(6):1011–1020, 2001.

[44] S. Kim, J. Alonso, and A. Jameson. Design optimization of high-lift configu-rations using a viscous continuous adjoint method. In 40th AIAA AerospaceSciences Meeting & Exhibit, page 844, 2002.

[45] A. Kogut, A. J. Banday, C. L. Bennett, K. M. Gorski, G. Hinshaw, G. F. Smoot,and E. L. Wright. Tests for non-gaussian statistics in the dmr four-year skymaps. The Astrophysical Journal Letters, 464(1):L29, 1996.

[46] O. Lehmann, M. Luchtenburg, B. R. Noack, R. King, M. Morzynski, and G. Tad-mor. Wake stabilization using pod galerkin models with interpolated modes. InProceedings of the 44th IEEE Conference on Decision and Control, pages 500–505. IEEE, 2005.

[47] Y. Li and C. Meneveau. Origin of non-gaussian statistics in hydrodynamic tur-bulence. Physical review letters, 95(16):164502, 2005.

[48] B. Liu and J. Goree. Superdiffusion and non-gaussian statistics in a driven-dissipative 2d dusty plasma. Physical review letters, 100(5):055003, 2008.

[49] E. N. Lorenz. Deterministic nonperiodic flow. Journal of the atmospheric sci-ences, 20(2):130–141, 1963.

[50] S. Markidis, J. Gong, M. Schliephake, E. Laure, A. Hart, D. Henty, K. Heisey,and P. Fischer. Openacc acceleration of the nek5000 spectral element code. TheInternational Journal of High Performance Computing Applications, 29(3):311–319, 2015.

[51] L. Mason, J. Baxter, P. L. Bartlett, and M. R. Frean. Boosting algorithms asgradient descent. In Advances in neural information processing systems, pages512–518, 2000.

[52] K. Maute, M. Nikbay, and C. Farhat. Sensitivity analysis and design optimizationof three-dimensional non-linear aeroelastic systems by the adjoint method. In-ternational Journal for Numerical Methods in Engineering, 56(6):911–933, 2003.

[53] A. McNamara, A. Treuille, Z. PopoviΔ‡, and J. Stam. Fluid control using theadjoint method. In ACM Transactions On Graphics (TOG), volume 23, pages449–456. ACM, 2004.

62

[54] A. Medio and G. Gallo. Chaotic dynamics. Technical report, Cambridge Uni-versity Press, 1995.

[55] M. Minguez, R. Pasquetti, and E. Serre. High-order large-eddy simulation offlow over the β€œahmed body” car model. Physics of fluids, 20(9):095101, 2008.

[56] J. F. B. Mitchell, J. Lowe, R. A. Wood, and M. Vellinga. Extreme events due tohuman-induced climate change. Philosophical Transactions of the Royal SocietyA: Mathematical, Physical and Engineering Sciences, 364(1845):2117–2133, 2006.

[57] A. Montina, U. Bortolozzo, S. Residori, and F. T. Arecchi. Non-gaussian statis-tics and extreme waves in a nonlinear optical cavity. Physical review letters,103(17):173901, 2009.

[58] N. Mori and P. A. E. M. Janssen. On kurtosis and occurrence probability offreak waves. Journal of Physical Oceanography, 36(7):1471–1483, 2006.

[59] G. Mu, W. Zhou, W. Chen, and J. KertΓ©sz. Order flow dynamics around ex-treme price changes on an emerging stock market. New Journal of Physics,12(7):075037, 2010.

[60] B. R. Noack, K. Afanasiev, M. MorzyΕ„ski, G. Tadmor, and F. Thiele. A hierarchyof low-dimensional models for the transient and post-transient cylinder wake.Journal of Fluid Mechanics, 497:335–363, 2003.

[61] B. R. Noack, W. Stankiewicz, M. MorzyΕ„ski, and P. J. Schmid. Recursive dy-namic mode decomposition of transient and post-transient wake flows. Journalof Fluid Mechanics, 809:843–872, 2016.

[62] N. Offermans, O. Marin, M. Schanen, J. Gong, P. Fischer, P. Schlatter, A. Ob-abko, A. Peplinski, M. Hutchinson, and E. Merzari. On the strong scaling ofthe spectral element solver nek5000 on petascale systems. In Proceedings of theExascale Applications and Software Conference 2016, page 5. ACM, 2016.

[63] M. Onorato, A. R. Osborne, M. Serio, and S. Bertone. Freak waves in randomoceanic sea states. Physical review letters, 86(25):5831, 2001.

[64] J. Γ–sth, B. R. Noack, S. KrajnoviΔ‡, D. Barros, and J. BorΓ©e. On the needfor a nonlinear subscale turbulence term in pod models as exemplified for ahigh-reynolds-number flow over an ahmed body. Journal of Fluid Mechanics,747:518–544, 2014.

[65] D.I. Papadimitriou and K.C. Giannakoglou. A continuous adjoint method withobjective function derivatives based on boundary integrals, for inviscid and vis-cous flows. Computers & Fluids, 36(2):325–341, 2007.

[66] A. Peplinski, P. Schlatter, P. F. Fischer, and D. S. Henningson. Stability tools forthe spectral-element code nek5000: Application to jet-in-crossflow. In Spectraland High Order Methods for Partial Differential Equations-ICOSAHOM 2012,pages 349–359. Springer, 2014.

63

[67] N. Qian. On the momentum term in gradient descent learning algorithms. Neuralnetworks, 12(1):145–151, 1999.

[68] C. Qiu and Q. Xu. A simple adjoint method of wind analysis for single-dopplerdata. Journal of Atmospheric and Oceanic Technology, 9(5):588–598, 1992.

[69] C. W. Rowley. Model reduction for fluids, using balanced proper orthogonal de-composition. International Journal of Bifurcation and Chaos, 15(03):997–1013,2005.

[70] S. Ruder. An overview of gradient descent optimization algorithms. arXivpreprint arXiv:1609.04747, 2016.

[71] D. Ruelle. Small random perturbations of dynamical systems and the definitionof attractors. Communications in Mathematical Physics, 82(1):137–151, 1981.

[72] P. J. Schmid. Dynamic mode decomposition of numerical and experimental data.Journal of fluid mechanics, 656:5–28, 2010.

[73] M. Schramm, B. Stoevesandt, and J. Peinke. Adjoint optimization of 2d-airfoilsin incompressible flows. In 11th World Congress on Computational Mechanics,Wccm, pages 6200–6211, 2014.

[74] J. Shin, M. W. Hall, J. Chame, C. Chen, P. F. Fischer, and P. D. Hovland.Speeding up nek5000 with autotuning and specialization. In Proceedings of the24th ACM International Conference on Supercomputing, pages 253–262. ACM,2010.

[75] O. Soto and R. LΓΆhner. On the computation of flow sensitivities from boundaryintegrals. AIAA paper, 112:2004, 2004.

[76] O. Soto, R. LΓΆhner, and C. Yang. An adjoint-based design methodology for cfdproblems. International Journal of Numerical Methods for Heat & Fluid Flow,14(6):734–759, 2004.

[77] G. Tadmor, O. Lehmann, B. R. Noack, L. Cordier, J. Delville, J. Bonnet, andM. MorzyΕ„ski. Reduced-order models for closed-loop wake control. PhilosophicalTransactions of the Royal Society A: Mathematical, Physical and EngineeringSciences, 369(1940):1513–1524, 2011.

[78] F. Takens. Detecting strange attractors in turbulence. In Dynamical systemsand turbulence, Warwick 1980, pages 366–381. Springer, 1981.

[79] J. A. Tawn. Modelling multivariate extreme value distributions. Biometrika,77(2):245–253, 1990.

[80] A. Thacker, S. Aubrun, A. Leroy, and P. Devinant. Effects of suppressing the3d separation on the rear slant on the flow structures around an ahmed body.Journal of Wind Engineering and Industrial Aerodynamics, 107:237–243, 2012.

64

[81] G. Tissot, L. Cordier, N. Benard, and B. R. Noack. Model reduction usingdynamic mode decomposition. Comptes Rendus MΓ©canique, 342(6-7):410–416,2014.

[82] J. Venkatramani, S. K. Kumar, S. Sarkar, and S. Gupta. Physical mechanismof intermittency route to aeroelastic flutter. Journal of Fluids and Structures,75:9–26, 2017.

[83] J. Wenger, R. Tualle-Brouri, and P. Grangier. Non-gaussian statistics from in-dividual pulses of squeezed light. Physical review letters, 92(15):153601, 2004.

[84] B. S. White and B. Fornberg. On the chance of freak waves at sea. Journal offluid mechanics, 355:113–138, 1998.

[85] K. Willcox and J. Peraire. Balanced model reduction via the proper orthogonaldecomposition. AIAA journal, 40(11):2323–2330, 2002.

[86] G. Williams. Chaos theory tamed. CRC Press, 1997.

65