+ All Categories
Home > Documents > DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING...

DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING...

Date post: 03-Jul-2018
Category:
Upload: tranthien
View: 220 times
Download: 0 times
Share this document with a friend
74
DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING OF STOCHASTIC NONLINEAR WATER WAVES Saviz Mowlavi Supervisors Prof. Themistoklis Sapsis (MIT) & Prof. François Gallaire (EPFL) Submitted in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering at the Ecole Polytechnique Fédérale de Lausanne February 2015
Transcript
Page 1: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

DYNAMICALLY ORTHOGONAL REDUCED-ORDER

MODELING OF STOCHASTIC NONLINEAR WATER WAVES

Saviz Mowlavi

Supervisors

Prof. Themistoklis Sapsis (MIT) & Prof. François Gallaire (EPFL)

Submitted in partial fulfillment of the requirements for the degree ofMaster of Science in Mechanical Engineering

at the

Ecole Polytechnique Fédérale de Lausanne

February 2015

Page 2: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework
Page 3: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Acknowledgements

Working under the guidance of Prof. Themistoklis Sapsis during these past six months at MIT

has been an immense pleasure. His scientific advice, his open-mindedness, his availability and

his endless patience, all contributed to making my research here a very enjoyable experience.

I warmly thank him for all these reasons, and look forward to starting my Ph.D. under his

supervision.

It is through the passionate lectures of Prof. François Gallaire that I started developing a

profound interest for the fascinating field of fluid mechanics, and it is under his supervision

that I first entered the research world. He will long remain a source of inspiration for me. I

owe him a big debt of gratitude and thank him for always being so helpful and supportive in

my decisions.

I would also like to thank Will Cousins who has always been willing to spend all his time

answering my questions, my lab mates at SandLab for great company and my friends for

providing me with refreshing moments of escape from work.

Finally, words alone cannot express the gratitude that I feel towards my family for their

unconditional love and support in all circumstances. I feel very lucky to have them as a family

and they are the most precious thing I have in this world.

i

Page 4: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework
Page 5: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Abstract

In this thesis, we implement a reduced-order framework for the stochastic evolution of non-

linear water waves governed by the nonlinear Schrördinger (NLS) equation and subject to

random initial conditions. Our reduced-order model is based on the dynamically orthogonal

(DO) equations introduced by Sapsis & Lermusiaux (2009), and consists in the expansion of

the stochastic solution on a few time-dependent deterministic modes that capture the sub-

space where the dominant stochastic fluctuations reside, while an associated set of stochastic

coefficients describes the stochasticity within this subspace. Using a dynamical orthogonality

condition for the modes, a closed set of coupled evolution equations for the mean state, the

modes and the stochastic coefficients can be directly derived from the governing NLS equation.

This reduced-order set of DO equations enables the efficient computation of the stochastic

solution, and permits the visualization in phase space of its time-dependent structure. We

benchmark this reduced-order model against two well-known cases, that of a uniform wave-

train undergoing Benjamin-Feir instability and Fermi-Pasta-Ulam recurrence, and that of an

ensemble of waves with Gaussian spectrum and random phases. In both cases, we obtain

a very good agreement with results reported in the literature, validating our DO equations.

Finally, we exploit the benefits of the DO framework to study the nonlinear evolution of an

extreme wave subjected to small initial stochastic perturbations and we visualize its attractor

in phase space.

Key words: Water waves, nonlinear Schrödinger equation, reduced-order modeling, stochastic

dynamical systems.

iii

Page 6: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework
Page 7: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Contents

Acknowledgements i

Abstract iii

Introduction 1

1 Review of deep-water wave theory 3

1.1 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.1.1 Linear dispersion relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.1.2 Energy considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.2 Weakly nonlinear envelope equation . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.2.1 Weakly nonlinear narrow-band wavetrain . . . . . . . . . . . . . . . . . . 5

1.2.2 Nonlinear Schrödinger equation . . . . . . . . . . . . . . . . . . . . . . . . 6

1.3 Modulational instability and long-time evolution . . . . . . . . . . . . . . . . . . 7

1.3.1 Uniform Stokes wave solution . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.3.2 Benjamin-Feir instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.3.3 Fermi-Pasta-Ulam recurrence . . . . . . . . . . . . . . . . . . . . . . . . . 8

2 NLS equation under the DO framework 11

2.1 Review of probability theory and KL expansion . . . . . . . . . . . . . . . . . . . 11

2.1.1 Probability spaces and random variables . . . . . . . . . . . . . . . . . . . 11

2.1.2 Stochastic processes and random fields . . . . . . . . . . . . . . . . . . . . 12

2.1.3 Karhunen-Loève orthogonal expansion . . . . . . . . . . . . . . . . . . . . 13

2.2 Dynamically orthogonal NLS equation . . . . . . . . . . . . . . . . . . . . . . . . 15

2.2.1 Dynamically orthogonal expansion . . . . . . . . . . . . . . . . . . . . . . 15

2.2.2 On the choice of the inner product . . . . . . . . . . . . . . . . . . . . . . . 17

2.2.3 Dynamically orthogonal equations . . . . . . . . . . . . . . . . . . . . . . 18

2.3 Stochastic energy transfers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2.4 Numerical implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.4.1 Numerical schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.4.2 Initial condition formulation . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.4.3 Diagonalization of the covariance matrix . . . . . . . . . . . . . . . . . . . 23

2.4.4 Overview of the code structure . . . . . . . . . . . . . . . . . . . . . . . . . 24

v

Page 8: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Contents

3 Preliminary results and validation 27

3.1 Idealized Benjamin-Feir instability . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.2 Random Gaussian wavenumber spectrum . . . . . . . . . . . . . . . . . . . . . . 32

3.2.1 Statistical properties of the DO solution . . . . . . . . . . . . . . . . . . . . 34

3.2.2 Structure of the DO solution . . . . . . . . . . . . . . . . . . . . . . . . . . 37

4 Dynamics of an extreme wave 41

4.1 Nonlinear focusing of localized wave packets . . . . . . . . . . . . . . . . . . . . 41

4.2 Adaptivity of the DO modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

4.3 Attractor of an idealized extreme wave . . . . . . . . . . . . . . . . . . . . . . . . 43

5 Results with the MNLS equation 47

5.1 Modified nonlinear Schrödinger and DO equations . . . . . . . . . . . . . . . . . 47

5.2 Idealized Benjamin-Feir instability . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

5.3 Attractor of an idealized extreme wave . . . . . . . . . . . . . . . . . . . . . . . . 50

Conclusions and perspectives 55

A Complex coefficients in the DO framework 57

A.1 Dynamically orthogonal equations with complex coefficients . . . . . . . . . . . 57

A.2 Diagonalization of the complex covariance matrix . . . . . . . . . . . . . . . . . 58

B Dynamically orthogonal MNLS equation 61

Bibliography 63

vi

Page 9: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Introduction

Starting with the work of Stokes (1847), water waves have attracted the attention of scientists

and researchers alike. Because nonlinearity enters through the boundary conditions of their

governing equations, they are notoriously difficult to solve analytically and have continuously

posed great challenges. This probably explains why it was not until the work of Lighthill

(1965) followed by the experimental confirmation of Benjamin & Feir (1967) that weakly

nonlinear deep-water uniform wavetrains were found to be unstable to small modulational

perturbations, a surprising phenomenon that became known as the Benjamin-Feir instability.

Subsequent research focused on understanding the long-time evolution and dynamics of

these unstable wavetrains. A great progress in that direction was made by Zakharov (1968),

who derived a simple governing equation for the envelope of weakly nonlinear narrow-band

wavetrains to cubic order in nonlinearity, the nonlinear Schrödinger (NLS) equation. Through

experiments and numerical simulations of the NLS equation, Lake et al. (1977) showed that the

long-time behavior of the Benjamin-Feir instability included the generation of large coherent

structures through spatial focusing of the energy of the wave field. Later, Dysthe (1979) and

Trulsen & Dysthe (1996) introduced higher-order more accurate versions of the NLS equation.

Nevertheless, in this thesis we will mainly use the NLS equation because of its reasonable

accuracy in the one-dimensional setting that we consider (Yuen & Lake, 1980).

Water waves are characterized by the interplay between two mechanisms. Dispersive effects

induce wave packet dispersion and mixing between different wavenumbers, while nonlinear

effects create energy transfers between modes and can result in the spatial focusing of wave

groups. Rogue waves are a dramatic example of the effects of nonlinearities. They are deep-

water gravity waves of much larger height than one would expect given the sea state (Dysthe

et al., 2008). Because of their potential catastrophic impact (Liu, 2007), they have recently

started to be an area of active research (Onorato et al., 2001). The Benjamin-Feir instability was

believed to be the basic physical mechanism at their origin, for it gives rise to large structures.

However, it applies to the idealized case of a uniform wavetrain while the ocean surface is

an irregular surface with energy spread over a range of wavenumbers. Nevertheless, Alber

(1978) showed that a similar type of instability persisted in random wavetrains characterized

by a narrow Gaussian spectrum with random phases. Numerical simulations of the NLS

equation and its higher-order derivatives confirmed that a sufficiently narrow spectrum leads

to an increased occurrence of these extreme waves (Janssen, 2003; Dysthe et al., 2003). More

1

Page 10: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Contents

recently, Ruban (2013) and Cousins & Sapsis (2015b,a) have shown that these extreme waves

are in fact induced by the nonlinear focusing of spatially localized wave groups that exceed a

certain critical length scale and amplitude.

In this thesis, we adopt a reduced-order stochastic approach towards the modeling of nonlin-

ear deep-water waves and we investigate the aforementioned phenomena in this reduced-

order stochastic setting. Specifically, we use the dynamically orthogonal (DO) framework

introduced by Sapsis & Lermusiaux (2009) to derive reduced-order equations for the stochastic

evolution of water waves governed by the NLS equation and subject to random initial con-

ditions. This DO framework is based on the expansion of the stochastic solution on a few

time-dependent deterministic modes with associated stochastic coefficients that efficiently

describe stochastic fluctuations around the mean state as the system evolves in time. From

the governing NLS equation and a dynamical orthogonality condition for the modes, can then

be derived a set of explicit equations that allow for the coupled simultaneous evolution of (i)

the mean state, (ii) the reduced-order subspace that contains the stochastic fluctuations (i.e.

the spatio-temporal form of the modes) and (iii) the stochasticity within this subspace (i.e. the

stochastic coefficients). Our choice of the DO framework follows from the time adaptivity of

its modes (critical in the highly transient context of water waves) and from the fact that no

prior knowledge on the modes is required.

This thesis is structured as follows. In Chapter 1 we provide a review of the theory related to

weakly nonlinear unidirectional waves on deep water. Then, in Chapter 2 we develop our DO

reduced-order framework for the stochastic modeling of water waves and we detail its numeri-

cal implementation. In Chapter 3, we illustrate the use of the obtained DO equations through

the computation of the stochastic solutions for two well-known situations. By comparing our

results with those from the literature, we validate the accuracy of our DO equations. Finally,

in Chapter 4 we study the nonlinear evolution of an extreme wave subjected to small initial

stochastic perturbations and we visualize its structure in phase space.

2

Page 11: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 1

Review of deep-water wave theory

In this first chapter, we provide a review of the theory related to weakly nonlinear unidirectional

waves on deep water. After presenting in Section 1.1 the governing equations for the surface

elevation, we show in Section 1.2 how perturbation methods can lead to a simplified nonlinear

equation for the envelope of a weakly nonlinear narrow-band wavetrain. We then introduce

in Section 1.3 the Benjamin-Feir modulational instability, which plays an important role in

the dynamics of nonlinear wavetrains and we discuss its long-time evolution.

1.1 Governing equations

Consider a two-dimensional system consisting of two layers of incompressible and inviscid

fluid, water at the bottom and air at the top. Since we are concerned with deep-water gravity

waves, we assume the water layer to be of infinite depth and we neglect surface tension

effects. Although water is not inviscid, here viscosity is neglected because it is effective only

for small-scale motion (Yuen & Lake, 1980). The air layer is assumed to remain at rest (which is

justified by the difference in densities), thus neglecting wind-wave interactions and effectively

restricting the problem to the water region. An (x, z) coordinate system is chosen in such a

way that the undisturbed water surface coincides with z = 0 and gravity points in the negative

z-direction. The fluid being irrotational, potential flow theory can be used and the system is

characterized by the surface elevation η(x, t ) of the water and its velocity potential φ(x, z, t ).

In the water domain, the velocity potential satisfies Laplace’s equation

∇2φ= 0 for −∞< z < η(x, t ), (1.1)

with the following boundary condition that enforces a vanishing velocity at the bed

∂φ

∂z= 0 when z →−∞. (1.2)

3

Page 12: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 1. Review of deep-water wave theory

There are two boundary conditions at the free surface. The kinematic boundary condition

ensures that fluid particles cannot traverse the surface, by equating the normal component of

the fluid velocity at the surface with that of the surface’s motion

∂η

∂t+ ∂φ

∂x

∂η

∂x− ∂φ

∂z= 0 at z = η(x, t ). (1.3)

Because we have neglected surface tension effects, the dynamic boundary condition states

that the fluid pressure at the surface equals the atmospheric pressure. The Bernoulli equation

applied at the surface therefore takes the following form

∂φ

∂t+ 1

2(∇φ)2 + g z = 0 at z = η(x, t ), (1.4)

where we have taken the atmospheric pressure to be zero without loss of generality, and g

denotes the acceleration of gravity. The set of equations (1.1) to (1.4) constitutes the govern-

ing equations of gravity waves on deep water. Although Laplace’s equation for the velocity

potential φ(x, z, t ) is linear, it applies to a domain for which one of the boundaries, the water

surface η(x, t ), is unknown a priori and itself part of the problem. Nonlinearity thus appears

implicitly through the boundary conditions at the unknown surface.

1.1.1 Linear dispersion relation

Restricting ourselves to small disturbances about the base state η = 0 and φ = 0, the free

surface boundary conditions can be linearized and applied at the undisturbed surface z = 0.

Assuming periodic conditions in the horizontal direction, the variables η and φ may then be

expanded in Fourier modes in x and t , leading to the linear solution

η(x, t ) = a cos(kx −ωt ), (1.5)

φ(x, z, t ) = ag

ωekz sin(kx −ωt ), (1.6)

where the wave frequency ω is related to the wavenumber k through the well-known linear

dispersion relation ω = √g k. The phase velocity of the waves is given by c = ω/k = √

g /k

while their group velocity is vg = ∂ω/∂k = c/2. The dependency of c on k is indicative of the

dispersive nature of the system, a property that will remain important for finite-amplitude

waves as linear dispersion will counteract nonlinearity.

1.1.2 Energy considerations

Finally, we note that because of the absence of dissipation and external forcing, the governing

equations (1.1) to (1.4) conserve the total energy of the fluid (see Janssen, 2004), expressed as

H = ρg∫ L

0

∫ η

−∞z dz dx + 1

∫ L

0

∫ η

−∞(∇φ)2dz dx, (1.7)

4

Page 13: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

1.2. Weakly nonlinear envelope equation

where the first term represents the potential energy of the fluid, the second term its kinetic

energy and we have considered a periodic domain x ∈ D = [0,L]. Similarly to an harmonic

oscillator, we can think of the wave motion as caused by the resonance between the kinetic

and potential energies. In fact, we can use the linear solution (1.5)–(1.6) to show that these

energies are equal to leading order, giving the following expression for the total energy

H = 1

2ρg

∫ L

0η2dx + 1

4ρg a2

∫ L

0(1+2kη+O (η2))dx = 1

2ρg a2L+O (a4). (1.8)

1.2 Weakly nonlinear envelope equation

The linear solution (1.5)–(1.6) describes a uniform wavetrain of perfect periodicity and con-

stant infinitesimal amplitude. It provides a sufficient description for linear systems, as the

properties of an entire wave system can then be determined by superposition. However,

this no longer holds true for nonlinear systems like the one given by governing equations

(1.1)–(1.4), hence we rather seek a generalization of the idealized uniform wavetrain where we

would allow for slow (with respect to the wavelength and wave period) variations in space and

time of the amplitude, wavenumber and frequency.

1.2.1 Weakly nonlinear narrow-band wavetrain

Specifically, we introduce the weakly nonlinear narrow-band wavetrain through the concept

of a carrier wave with carrier wavenumber k0 and frequency ω0, that is modulated by a small

but finite and slowly varying complex envelope function u(x, t ) = a(x, t )e iθ(x,t ). The respective

roles of the modulus a(x, t ) and the phase θ(x, t ) can be made explicit through the following

(single-harmonic) expression for the surface elevation of the resulting wavetrain

η(x, t ) = Reu(x, t )e i (k0x−ω0t ) = Rea(x, t )e iθ(x,t )e i (k0x−ω0t ), (1.9)

where it is seen that a(x, t ) represents a slowly varying modulation amplitude, and θ(x, t ) is a

slowly varying phase modulation function that describes small variations in wavenumber and

frequency of the wavetrain about the carrier wavenumber and frequency. Specifically, we have

the modulation wavenumber ∆k and modulation frequency ∆ω

∆k = ∂θ

∂x¿ k0, ∆ω=−∂θ

∂t¿ω0, (1.10)

both of which are small with respect to k0 and ω0, hence the ‘narrow-band’ designation

as departures in wavenumber k = k0 +∆k and frequency ω = ω0 +∆ω of the wavetrain are

restricted to a narrow-band window around the corresponding carrier properties.

5

Page 14: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 1. Review of deep-water wave theory

1.2.2 Nonlinear Schrödinger equation

We now seek to obtain the equation governing the evolution of the slowly varying complex

envelope u(x, t). This can be done by taking advantage of both the small amplitude a and

the slow time and space variation of u, through a perturbation expansion combined with a

multiple-scales method. As a small parameter, we define the wave steepness ε= k0a ¿ 1 and

we require the modulation bandwidth to be of the same order of magnitude i.e. ∆k/k0 =O (ε),

meaning that u varies on the slow space and time scales εx and εt . We then employ the

following harmonic expansions for η(x, t ) and φ(x, z, t )

η(x, t ) = Reu(εx,εt )e i (k0x−ω0t ) +u2(εx,εt )e2i (k0x−ω0t ) + . . . , (1.11)

φ(x, z, t ) = Rev(εx, z,εt )ek0z e i (k0x−ω0t ) + v2(εx, z,εt )e2k0z e2i (k0x−ω0t ) + . . . , (1.12)

where the slowly varying complex functions u, u2, ..., v , v2, ... are O (ε/k0). Inserting these

expansions into the governing equations (1.1)–(1.4) and expanding the free-surface boundary

conditions (1.3)–(1.4) in Taylor series about the undisturbed surface z = 0, the equations can

be solved in orders of ε. At the first order, the usual linear dispersion relation ω0 =√

g k0 for

the carrier wave is recovered. At the second order, it is found that the envelope is advected at

the group velocity of the carrier wave. Taking the perturbation expansion for the first harmonic

to the third order gives the following evolution equation for the complex envelope

i

(∂u

∂t+ ω0

2k0

∂u

∂x

)− ω0

8k20

∂2u

∂x2 − 1

2ω0k2

0 |u|2u = 0, (1.13)

where the first term indicates advection of the envelope at the group velocity of the carrier

wave, the second term represents effects of dispersion, and the third term describes nonlinear

effects to the lowest-order. This equation is the nonlinear Schrödinger (NLS) equation. It has

widespread use in physics across a number of fields that involve nonlinear dispersive waves

and was first obtained by Benney & Newell (1967) in a general context. For deep-water waves,

it has been derived using a variety of methods, first by Zakharov (1968) using a spectral method,

then by Hasimoto & Ono (1972) and Davey (1972) using multiple-scale methods and finally by

Yuen & Lake (1975) through a Lagrangian variational approach from Whitham (1965). Finally,

from the perturbation expansion for the second harmonic we have u2 = k0u2/2, implying that

the dynamics of the higher harmonics are tied to the first one. The waves pertaining to the

higher harmonics are therefore referred to as the bound waves, while those lying within the

bandwidth of the first harmonic are called the free waves. Note also that u2 =O (ε2/k20), thus

the first harmonic provides a sufficient description of the surface elevation

η(x, t ) = Reu(εx,εt )e i (k0x−ω0t )+O (ε2/k20). (1.14)

In studying solutions to the NLS equation, it is helpful to use a reference frame moving at the

6

Page 15: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

1.3. Modulational instability and long-time evolution

group velocity of the carrier wave, which is achieved through the change of coordinates

x∗ = x − ω0

2k0t . (1.15)

Additionally, we use the length and time scales imposed by the wavelength and period of the

carrier wave to introduce the following nondimensional variables

t =ω0t , x = k0x∗, u = k0u, ∆k = ∆k

k0, ∆ω= ∆ω

ω0, (1.16)

leading to the following nondimensionalized form of the NLS equation (1.13)

∂u

∂t=− i

8

∂2u

∂x2 − i

2|u|2u. (1.17)

and the surface elevation to lowest order is expressed as η(x, t ;ω) = Reu(x, t ;ω)e i (x−t/2). Note

that under this choice of length scale, the wave steepness ε= k0a becomes equivalent to the

nondimensional wave amplitude |u|. Unless specified otherwise, nondimensional variables

will be used in the remainder of this thesis so we hereafter drop the prime symbols.

1.3 Modulational instability and long-time evolution

1.3.1 Uniform Stokes wave solution

As a first step in investigating the properties of weakly nonlinear wavetrains, we look at the

simplest solution of the NLS equation (1.17), given by the following spatially constant envelope

u(x, t ) = a0 e−(i /2)a20 t , a0 ∈R, (1.18)

which describes a uniform wavetrain similar to the linear solution (1.5), but with a nonlinear

correction to the linear dispersion relation ω0 = √g k0. This can be shown by writing the

corresponding expression for the surface elevation in dimensional form

η(x, t ) = Rea0 e−(i /2)k20 a2

0ω0t e i (k0x−ω0t )+O (a20)

= a0 cos[k0x −ω0(1+k20 a2

0/2)t ]+O (a20). (1.19)

We obtain a uniform wave with dispersion relation ω=ω0(1+k20 a2

0/2) that depends not only

on the wavenumber k0, but also on the finite amplitude a0. This nonlinear correction was

already found by Stokes (1847) through a weakly nonlinear harmonic expansion of the uniform

wavetrain. He also found that the higher harmonics resulted in an altered profile with sharp

crests and flat troughs, resulting in the so-called Stokes wave. Note that the dispersion relation

for the finite-amplitude uniform wave gives

1

2

∂2ω

∂k20

=− ω0

8k20

and − ∂ω

∂a20

=−1

2ω0k2

0 , (1.20)

7

Page 16: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 1. Review of deep-water wave theory

which shows that the second and third terms of the NLS equation (1.13) indeed relate to

dispersion and nonlinear effects, respectively. This can be shown in a more rigorous way

from a heuristic derivation of the NLS equation based on a Taylor expansion of the dispersion

relation around wavenumber k0 and zero amplitude (Yuen & Lake, 1982; Janssen, 2004).

1.3.2 Benjamin-Feir instability

The linear stability of this uniform wavetrain can be investigated by imposing small Fourier

mode perturbations representing infinitesimal modulations in amplitude and phase

u(x, t ) = a0 (1+b1e i (∆kx−∆ωt ) + i b2e i (∆kx−∆ωt ))e−(i /2)a20 t , (1.21)

where |b1| and |b2| are infinitesimal real numbers and ∆k, ∆ω represent respectively the

nondimensional modulation wavenumber and frequency around the carrier wavenumber

and frequency. Linearizing the NLS equation about the uniform solution and solving the

resulting eigenvalue problem results in the following dispersion relation

∆ω2 = ∆k2

8

(∆k2

8−a2

0

). (1.22)

This indicates that perturbations with nondimensional wavenumber in the range

0 <∆k <∆kc = 2p

2a0 (1.23)

have positive growth rate σ= Im∆ω, hence are linearly unstable, while ∆kc = 2p

2a0 repre-

sents a cut-off wavenumber above which there is restabilization. The instability is maximum at

∆km = 2a0 with an associated nondimensional maximum growth rate of σm = a20/2, implying

that it occurs on a nondimensional timescale of O (a−20 ). This instability is called the Benjamin-

Feir (BF) or modulational instability and was initially discovered for very long perturbation

wavelengths by Lighthill (1965), but it was Benjamin & Feir (1967) who first obtained equation

(1.22) and found the restabilization at higher perturbation wavenumbers. A plot of the growth

rate versus perturbation wavenumber is shown in Figure 1.1. Note that the instability region

depends on the amplitude and disappears as a0 tends to 0, thus recovering the linear result of

a marginally stable water surface around z = 0.

1.3.3 Fermi-Pasta-Ulam recurrence

Through experiments in a wave tank and simulations of the NLS equation, Lake et al. (1977)

studied the long-time behavior of the Benjamin-Feir instability. They showed that after an

initial period of exponential growth, unstable modulations grow to a maximum and saturate

before decaying, and the wavetrain returns to a nearly-uniform state, after which a new cycle

starts. This surprising behavior had previously been discovered in another context by Fermi

et al. (1955) and thus became known as the Fermi-Pasta-Ulam (FPU) recurrence.

8

Page 17: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

1.3. Modulational instability and long-time evolution

"k/a0

0 0.5 1 1.5 2 2.5 3<

/<m

0

0.2

0.4

0.6

0.8

1

Figure 1.1 – Normalized growth rate σ/σm of the Benjamin-Feir instability versus normalizedperturbation wavenumber ∆k/a0. Note that the instability is maximum at ∆km = 2a0 and thecut-off occurs at ∆kc = 2

p2a0.

Figure 1.2 – Benjamin-Feir instability and Fermi-Pasta-Ulam recurrence in a slightly modu-lated uniform wavetrain. Left, spatio-temporal evolution of the complex envelope modulus|u(x, t )|. Right, surface elevation η(x, t ) (blue) and envelope modulus |u(x, t )| (orange) at threedifferent times.

We illustrate the Benjamin-Feir instability and subsequent Fermi-Pasta-Ulam recurrence in

Figures 1.2 and 1.3, in which we have computed the numerical solution to a periodic initial

condition of the form u(x,0) = a0(1+0.1cos∆km x), with a0 = 0.1, representing a uniform

wavetrain that is slightly modulated by the linearly most unstable modulation wavenumber

∆km = 2a0 = 0.2. In the left-hand side plot of Figure 1.2, we plot the spatio-temporal evolution

of the complex envelope modulus |u(x, t )|. After an initial growth, the wavetrain is observed

to reach a strongly modulated state before returning to its initial state and undergoing a

new cycle. The right-hand side plot displays the surface elevation η(x, t) (blue) together

with the envelope modulus |u(x, t)| (orange) at three different times, and shows that the

strongly modulated state at t = 660 can be interpreted as a situation where energy from the

carrier wave has focused in space to produce localized ‘extreme’ waves. The Fourier spectrum

F [η] of the surface elevation at t = 500 is shown in the left-hand side plot of Figure 1.3

and shows that the Benjamin-Feir instability manifests itself as pairs of growing sideband

9

Page 18: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 1. Review of deep-water wave theory

k0 0.5 1 1.5 2

|F[2

]|

0

5

10

15

20

25

30

35

40

45Fourier transform of 2 at t = 500

t0 500 1000 1500 2000 2500 3000 3500

|F[u

]("

k)|

0

20

40

60

80

100

120Fourier modes of |u|

"k = 0"k = 1"km"k = 2"km"k = 3"km"k = 4"km

"km

2"km

Figure 1.3 – Benjamin-Feir instability and Fermi-Pasta-Ulam recurrence in a slightly mod-ulated uniform wavetrain. Left, Fourier spectrum F [η] of the surface elevation at t = 500,with k = 1 corresponding to the carrier wave and k = 1±∆km to the unstable modulation.Right, time evolution of the amplitudes of the carrier wave, the unstable modulation and itsharmonics obtained through a Fourier transform F [u] of the envelope. This time, ∆k = 0corresponds to the carrier wave and ∆k =∆km is the unstable modulation.

wavenumbers comprising the unstable initial modulation k = 1±∆km and its harmonics

k = 1±2∆km ,1±3∆km , ... Note that the carrier wave corresponds to k = 1 because of the

nondimensionalization. In the right half of Figure 1.3, the amplitudes of the carrier wave and of

the prescribed unstable modulation together with three of its harmonics are obtained through

a Fourier transform F [u] of the envelope, and plotted versus time. It is clearly observed that

the unstable modulation ∆k =∆km is growing at the expense of the carrier wave ∆k = 0. The

harmonics being linearly stable, they only appear as forced oscillations and are phase-locked

to the initially prescribed modulation. In the case where some of the harmonics lie within the

linearly unstable range (1.23), however, Yuen & Ferguson (1978) revealed that the long-time

behavior of the solution is governed by both the initially prescribed unstable mode and its

unstable harmonics, with the unstable mode and harmonics each taking turn dominating a

recurrence cycle.

As a conclusion, we would like to mention that in this chapter we have only provided a brief

review of the classical theory of weakly nonlinear deep-water waves. Our goal was mainly to

introduce the NLS equation to be used in the rest of this thesis, and to familiarize the reader

with the notion of a slowly modulated narrow-band wavetrain and its envelope description.

As a result, we have focused on the description of the idealized Benjamin-Feir instability and

Fermi-Pasta-Ulam recurrence of a uniform wavetrain, and we have deliberately omitted other

important notions such as the evolution of more realistic narrow-band Gaussian spectra of

waves or the nonlinear focusing of wave packets. These notions will be introduced in the later

chapters, along with the presentation of our results.

10

Page 19: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 2

NLS equation under the DO framework

In this chapter, we develop a reduced-order framework for the stochastic modeling of water

waves governed by the nonlinear Schrödinger (NLS) equation (1.17). This framework is based

on the dynamically orthogonal (DO) field equations first introduced by Sapsis & Lermusiaux

(2009), a novel order-reduction method for the solution of stochastic partial differential equa-

tions. In Section 2.1, we start with a brief review of probability spaces, stochastic processes

and the Karhunen-Loève expansion. The DO framework is then presented in Section 2.2

and applied to the NLS equation, resulting in a set of equations for the coupled evolution

of the mean state and the reduced-order stochastic fluctuations. In Section 2.3, we derive

expressions for the rates of energy transfer between the different dynamical components of

the solution. Finally, the numerical implementation of the equations is detailed in Section 2.4.

2.1 Review of probability theory and KL expansion

We start by introducing a few definitions and concepts from probability theory and stochastic

processes. Our goal is not to provide a comprehensive review of the subject (for this we refer

to Sobczyk, 1991), but merely to introduce the usual notation and provide select notions that

will prove useful in the following sections.

2.1.1 Probability spaces and random variables

A central concept in the description of random phenomena is the so-called probability space

(Ω,B,P ). The sample space Ω is the set of all elementary events ω ∈Ω associated with the

random phenomenon under consideration. B is the σ-algebra associated withΩ, that is and

loosely speaking, B is the collection of subsets ofΩ. The elements of B are random events and

consist of combinations of elementary eventsω. Finally, P is a probability measure defined on

B, i.e. P is a countably additive function that associates a non-negative number between 0

and 1 to each random event in B, and that assigns the value 1 to the sample spaceΩ. In loose

11

Page 20: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 2. NLS equation under the DO framework

terms, P can be interpreted as the function that ‘counts’ the number of elementary events ω

contained in a given subset of the sample spaceΩ, divided by the total number of elementary

events inΩ.

Various physical outcomes can occur in a random phenomena. In the simplest case, the

outcome of a given random experiment can be represented by a real number, thus one can

assign a real number X (ω) for each elementary event ω ∈Ω. This function X (ω) that maps

elements in Ω to values in R is called a random variable, and its probabilistic behavior is

completely specified by its probability distribution F (x), defined as

F (x) =P ω | X (ω) < x , x ∈R, ω ∈Ω, (2.1)

which characterizes the probability that X (ω) be less than x. For a continuous random variable,

the distribution function is continuous everywhere and we call probability density f (x) the

function given by the following derivative

f (x) = dF

dx, x ∈R. (2.2)

Let us stress that the probability space (Ω,B,P ) is defined independently of any particular

random phenomena, in the sense that the probabilistic behavior associated to the outcome of

a given random experiment is embedded in the random variable X (ω) itself, and not in the

elementary events ω which by definition happen with equal probability dP (ω). Finally, we

have the following definition for the mean value or expectation of the random variable X (ω)

E [X (ω)] =∫Ω

X (ω)dP (ω). (2.3)

2.1.2 Stochastic processes and random fields

In physical applications, many random phenomena are also time dependent, in which case

the outcome of a random experiment can be represented by a real function of time. This leads

us to the notion of a stochastic process. Denoting t ∈ T the time, a stochastic process X (t ;ω)

is a function that maps elementary events ω ∈ Ω to elements in the space of all finite and

real-valued functions of time. In this case, the function x(t) = X (t ;ω) associated to a fixed

elementary event ω ∈Ω is called a realization of the process.

We are ultimately interested in phenomena that are both time and space dependent. Denoting

x ∈ D ⊂R the one-dimensional space variable, we generalize the notion of a stochastic process

to that of a complex random field u(x, t ;ω), that is, a function that maps elementary events

ω ∈Ω to elements in the space of all finite and complex-valued functions of space and time.

Similarly to the mean value of a random variable, we can define the mean value of the random

12

Page 21: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

2.1. Review of probability theory and KL expansion

field u(x, t ;ω) as the ensemble average over the elementary events ω

u(x, t ) = E [u(x, t ;ω)] =∫Ω

u(x, t ;ω)dP (ω), (2.4)

2.1.3 Karhunen-Loève orthogonal expansion

Consider a complex random field u(x, t ;ω) that is continuous and square integrable, i.e.∫D E [u(x, t ;ω)u∗(x, t ;ω)]dx <∞ for all t ∈ T , where the asterisk denotes the complex conju-

gate. For a fixed time t , u(x, t ;ω) is a random function of x and lives in the infinite-dimensional

Hilbert space L2 of all continuous, square integrable and complex-valued functions defined

on D , and having the following spatial inner product

⟨u1,u2⟩ =∫

Du1(x)u∗

2 (x)dx for all u1,u2 ∈ L2. (2.5)

Since L2 is an infinite-dimensional space, it is spanned by an infinite number of orthogonal

basis functions vi (x)∞i=1. Hence, if we want to write the random field u(x, t ;ω) (at a given time)

as a linear decomposition of deterministic fields multiplied by scalar stochastic coefficients by

projecting each realization u(x, t ;ω) onto the basis vi (x)∞i=1, then we need an infinite series

of the form

u(x, t ;ω) = u(x, t )+∞∑

i=1Yi (t ;ω)vi (x), (2.6)

where u(x, t ) is the mean field and Yi (t ;ω) are zero-mean and complex-valued scalar stochastic

coefficients that carry information on the stochastic fluctuations of the random field u(x, t ;ω)

around the mean u(x, t). The question now is whether one can, for a given time, find a

deterministic basis ui (x, t )∞i=1 that is optimal for u(x, t ;ω) (at that time), in the sense that a

finite-dimensional representation of the form

u(x, t ;ω) ' u(x, t )+s∑

i=1Yi (t ;ω)ui (x, t ), s <∞, (2.7)

would approximate members of u(x, t ;ω) better than representations of the same dimension

in any other basis. This statement can be formalized (see Holmes et al., 1996) as follows

maxui∈L2

E [|⟨u(x, t ;ω),ui (x, t )⟩|2]

⟨ui (x, t ),ui (x, t )⟩ , (2.8)

where | · | denotes the absolute value, and it is looked for an ui (x, t) such that the ensemble

average of the projection of u(x, t ;ω) onto ui (x, t) is maximized. Using variational calculus

techniques, this condition reduces to the following eigenvalue problem∫D

Rt (x, y)ui (x, t )dx =λi (t )ui (y, t ), (2.9)

13

Page 22: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 2. NLS equation under the DO framework

where Rt (x, y) = E [u(x, t ;ω)u∗(y, t ;ω)] is the time-dependent autocorrelation function of

the random field u(x, t ;ω) and the desired optimal basis is given by the orthogonal set of

eigenfunctions ui (x, t ). The zero-mean stochastic coefficients in representation (2.7) are then

obtained by the projection of the stochastic fluctuations to the optimal basis

Yi (t ;ω) = ⟨u(x, t ;ω)− u(x, t ),ui (x, t )⟩, (2.10)

and they verify the following properties, relating to the eigenvalues λi (t )

E [|Yi (t ;ω)|2] =λi (t ), E [Yi (t ;ω)Y ∗j (t ;ω)] = 0 for i 6= j . (2.11)

Hence, the eigenvalues λi (t ) represent the variance (or spread) of the stochastic fluctuations

u(x, t ;ω)− u(x, t ) along each direction ui (x, t ) in phase space. As a consequence of the opti-

mality of the representation, the orthogonal directions ui (x, t ) are aligned with the principal

directions of the variance of u(x, t ;ω) (i.e. they capture the ‘dominant fluctuations’ of u(x, t ;ω)),

and it is often the case that the variance along subsequent directions decreases exponentially

i.e. λi (t) ∼ e−ci for some positive c. It is therefore justified to define a threshold s <∞ and

neglect the directions ui (x, t ) with i > s, along which u(x, t ;ω) has negligible spread. The study

of the random field u(x, t ;ω) is therefore restricted to the s-dimensional subspace spanned

by the eigenmodes associated to the s largest eigenvalues. This idea is at the foundation of

reduced-order modeling. Depending on the phenomenon under study, the value of s can be

very small, in which case we say that the system has a low-dimensional attractor.

Note that since Rt (x, t ) is self-adjoint and positive definite, we are assured that equation (2.9)

will always admit a countable infinity of positive eigenvalues and orthogonal eigenfunctions.

The linear orthogonal decomposition (2.7) with the optimal basis functions given by the

eigenvalue problem (2.9) is known as the Karhunen-Loève (KL) expansion (Loève, 1945). It has

found important applications, notably in fluid mechanics for the reduced-order modeling and

analysis of statistically stationary turbulent flows, where it bears the name Proper Orthogonal

Decomposition (POD) (Berkooz et al., 1993; Holmes et al., 1996). In this context, the ensemble

averages can be replaced by time averages over a single experimental run, and the time

dependence of the mean and the modes in the decomposition (2.7) is removed. The procedure

then involves two steps. First, a relevant set of optimal basis functions is found by capturing

experimental or numerical snapshots of a flow at different times, constructing the resulting

‘empirical’ autocorrelation function and solving the resulting eigenvalue problem (2.9). Then,

the low-dimensional deterministic dynamics are obtained by projecting the governing (Navier-

Stokes) equations on the previously obtained finite-dimensional basis, resulting in a set of

coupled equations for the time evolution of the coefficients Yi (t ;ω). Although the POD has

proved successful in identifying coherent structures and their dynamics in turbulent flows

(Aubry et al., 1988), it has two main limitations: (i) it uses time-independent basis functions

and is therefore not suitable for the description of transient phenomena, and (ii) it relies on

experiments or numerical simulations to derive the set of optimal basis functions.

14

Page 23: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

2.2. Dynamically orthogonal NLS equation

To overcome the aforementioned limitations of the POD and other methods in the context of

highly transient stochastic systems, Sapsis & Lermusiaux (2009) introduced the dynamically

orthogonal (DO) equations, a novel and time-adaptive reduced-order framework for the

solution of systems governed by generic stochastic partial differential equations (SPDEs).

Since transient phenomena and intermittent instabilities are often observed in deep-water

gravity waves (Cousins & Sapsis, 2015b,a), we opt to use the DO equations for the derivation

in the next section of a reduced-order framework for the modeling of stochastic water waves.

2.2 Dynamically orthogonal NLS equation

Let’s consider a complex random field u(x, t ;ω) representing the (nondimensional) complex

envelope of a deep-water weakly nonlinear narrow-band wavetrain, as defined in Chapter 1.

We assume that u(x, t ;ω) is continuous and square integrable (which makes sense from a

physical viewpoint). While this complex envelope is governed by the deterministic NLS

equation (1.17), we are interested in studying its evolution under random initial conditions

that follow some given probability distribution.

Stochasticity therefore enters the problem through the initial condition, which is specified in

the form of a large ensemble of initial realizations u(x, t0;ω) that follow a given probability

distribution. We then want to simultaneously evolve in time all realizations, so that at each

time instant we can recover the full stochastic solution u(x, t ;ω) and its associated statistics.

This can be readily done with a Monte-Carlo approach but will result in a high computational

cost, so we instead opt to use the DO reduced-order framework introduced in Sapsis & Lermu-

siaux (2009) to do this in an efficient way. Our choice of the DO framework instead of other

order-reduction methods follows from both its adaptivity and the fact that no prior knowledge

on the form of the basis functions is required.

2.2.1 Dynamically orthogonal expansion

The basis idea behind the DO method goes as follows. We first suppose that the random initial

condition can be accurately represented by a truncated KL expansion (2.7) at initial time t0

u(x, t0;ω) = u(x, t0)+s∑

i=1Yi (t0;ω)ui (x, t0), s <∞, (2.12)

where u(x, t0) is the initial mean, Yi (t0;ω) are the zero-mean initial stochastic coefficients

and ui (x, t0) are the initial set of orthonormal basis functions or modes. Note that s defines

the dimensionality of the subspace containing the initial stochastic fluctuations. Next, we

assume that the system (in this case governed by the NLS equation) retains a low-dimensional

attractor as time evolves, i.e. its stochastic solution u(x, t ;ω) at time t can still be accurately

represented by a truncated KL expansion of similar dimension. The DO method then provides

a set of coupled equations, directly derived from the system governing equation, for the time

15

Page 24: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 2. NLS equation under the DO framework

evolution of all quantities involved in (2.12), in such a way that the approximate full stochastic

solution at time t can be written in the form of the following DO expansion

u(x, t ;ω) = u(x, t )+s∑

i=1Yi (t ;ω)ui (x, t ), (2.13)

where u(x, t) is the time-dependent mean, ui (x, t) are the time-dependent deterministic

modes describing the main directions of stochastic fluctuations at time t and Yi (t ;ω) are the

time-dependent zero-mean stochastic coefficients. The DO solution given by (2.13) aims at

being close enough to an s-truncated KL expansion of the exact stochastic solution at time t ,

as would be obtained from a direct Monte-Carlo simulation on the system governing equation

with initial realizations u(x, t0;ω). The discrepancy between the two is caused by the effects

that dynamics along the neglected directions i > s may have on the resolved directions i ≤ s

and the mean. Although these effects can be large in turbulent systems (Sapsis & Majda, 2013),

they are negligible in a number of other cases (see Mantic-Lugo, Arratia & Gallaire, 2014, for a

dramatic example in the case of the flow behind a cylinder, where as a further approximation

the single mode that is used is computed as a quasilinear approximation around the mean).

The restriction of the stochastic dynamics to the subspace Vs = spanui (x, t)si=1 containing

the dominant fluctuations can thus be a good approximation, and results in a much better

computational efficiency than a full Monte-Carlo simulation.

In deriving explicit equations for all unknown quantities in the DO expansion (2.13), the

redundancy stemming from the allowed time variation of both the coefficients Yi (t ;ω) and

the modes ui (x, t ) needs to be overcome. Sapsis & Lermusiaux (2009) showed that this can be

achieved by imposing the following dynamical orthogonality (DO) condition

dVs

dt⊥Vs ⇔

⟨∂ui (·, t )

∂t,u j (·, t )

⟩= 0, i , j = 1, ..., s, (2.14)

that restricts the time variation of the subspace Vs where stochasticity lives to be orthonormal

to itself. In other words, since variations of the stochastic fluctuations within Vs can be entirely

described by variations of the stochastic coefficients Yi (t ;ω), the DO condition imposes the

natural constraint that modes only move when the fluctuations evolve to new directions

not already included in Vs . The DO condition also implies the preservation of the initial

orthonormality of the modes ui (x, t ) since

∂t⟨ui (·, t ),u j (·, t )⟩ =

⟨∂ui (·, t )

∂t,u j (·, t )

⟩+

⟨∂u j (·, t )

∂t,ui (·, t )

⟩= 0, i , j = 1, ..., s. (2.15)

In the next subsections, it will be shown how the insertion of the DO expansion (2.13) together

with the DO condition (2.14) in the NLS governing equation can lead to a closed and exact set

of coupled equations for the mean u(x, t ), the modes ui (x, t ) and the stochastic coefficients

Yi (t ;ω). We emphasize that the time-dependent basis functions ui (x, t ) are not chosen a priori

and are able to dynamically evolve to adapt to temporal changes in the dominant stochastic

fluctuations, thereby remedying two shortcomings of the POD.

16

Page 25: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

2.2. Dynamically orthogonal NLS equation

2.2.2 On the choice of the inner product

Before proceeding to the derivation of equations for the mean, modes and stochastic coef-

ficients, we remark that the choice of the inner product has important implications on the

quantities involved in the DO expansion (2.13) of the solution. Indeed, the inner product

defines the way the basis functions ui (x, t ) are orthogonal to each others, as well as the value

of the scalar coefficients Yi (t ;ω), for they result from the projection of the solution u(x, t ;ω)

onto the basis ui (x, t )si=1.

Since u(x, t ;ω) is complex-valued, it appears logical to consider the standard complex-valued

spatial inner product defined in equation (2.5) as

⟨u1,u2⟩ =∫

Du1(x, t ;ω)u∗

2 (x, t ;ω)dx. (2.16)

This complex-valued inner product implies that the stochastic coefficients in the DO ex-

pansion (2.13) will also be complex-valued. Note that previous DO schemes have always

concerned real-valued systems where all quantities are real (Choi et al., 2013; Sapsis et al.,

2013) and this is, to our knowledge, the first time that an attempt at using complex coefficients

and complex fields within the DO framework is carried out. While we managed to derive

the DO equations with complex coefficients (see Appendix A), complex statistics have to be

dealt with when analyzing the resulting solution, and this is very much still an area of active

research (for an overview of some of the complications involved, see Eriksson & Koivunen,

2006; Adali et al., 2011; Cheong Took et al., 2012).

Therefore we opted to go the safer route by using the following real-valued inner product

⟨u1,u2⟩ = Re

∫D

u1(x, t ;ω)u∗2 (x, t ;ω)dx

, (2.17)

so that the stochastic coefficients in the DO expansion (2.13) will be real-valued. To better

understand the implications of using a real-valued inner product in a space of complex-valued

functions, consider two basis functions u1(x, t) and u2(x, t) = i u1(x, t) that are orthogonal

under inner product (2.17), since ⟨u1,u2⟩ = Re−i = 0. Together they span the subspace

spanu1,u2 = Y1 u1(x, t )+Y2u2(x, t ) | Y1,Y2 ∈R

= (Y1 + i Y2)u1(x, t ) | Y1,Y2 ∈R

= Z1u1(x, t ) | Z1 ∈C . (2.18)

It is therefore observed that u1(x, t) and u2(x, t) span the same subspace as would u1(x, t)

alone with a complex coefficient. Note that under the complex-valued inner product (2.16) we

would indeed have ⟨u1,u2⟩ = −i 6= 0, confirming that the two directions are not orthogonal

when considering complex coefficients. Therefore the use of the real-valued inner product

(2.17) with real stochastic coefficients, while implying no loss of generality, results in a higher

number of basis functions required to span a given subspace for the stochastic fluctuations.

17

Page 26: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 2. NLS equation under the DO framework

As a side note, when using the real-valued inner product (2.17) with real coefficients, we can

make an analogy between the space of complex functions u(x, t ;ω) and that of real 2D vector

fields defined as (Reu(x, t ;ω), Imu(x, t ;ω)). Indeed, in this case it is easily seen that the real

inner product (2.17) is equal to the standard inner product on the space of real-valued 2D

vector fields. The two formulations are therefore completely equivalent. On the other hand,

the use of the complex-valued inner product (2.16) results in complex coefficients that have

the ability to swap the real and imaginary parts of the complex fields they are multiplying,

something that real coefficients cannot do hence there is no analogy with real 2D vector fields

in this case.

2.2.3 Dynamically orthogonal equations

In this section, we follow the steps in Sapsis & Lermusiaux (2009) to derive the DO equations

that govern the evolution of all unknown quantities in the DO expansion (2.13), i.e. the

mean u(x, t ), the deterministic modes ui (x, t ) and the stochastic coefficients Yi (t ;ω) for i =1, ..., s. As discussed in the previous section, we consider the real-valued inner product (2.17),

which implies that the stochastic coefficients Yi (t ;ω) are real. Recall that the nondimensional

complex envelope u(x, t ;ω) is governed by the deterministic NLS equation (1.17)

∂u(x, t ;ω)

∂t=− i

8

∂2u(x, t ;ω)

∂x2 − i

2|u(x, t ;ω)|2u(x, t ;ω), x ∈ D, t ∈ T, ω ∈Ω, (2.19)

subject to periodic boundary conditions, and to the following random initial conditions with

known probability distribution

u(x, t0;ω) = u0(x;ω), x ∈ D, ω ∈Ω. (2.20)

We begin by inserting the DO expansion (2.13) in the NLS equation (2.19), leading to the

following governing equation for all unknown quantities

∂u

∂t+ dYi

dtui +Yi

∂ui

∂t= F0 +Yi Fi +Yi Y j Fi j +Yi Y j Yk Fi j k , (2.21)

where repeated indices indicate summation from 1 to s, and F0, Fi , Fi j and Fi j k are complex

deterministic fields defined by

F0 =− i

8

∂2u

∂x2 − i

2|u|2u, Fi =− i

8

∂2ui

∂x2 − i u Reuu∗i − i

2|u|2ui ,

Fi j =− i

2Reui u∗

j u − i Reuu∗i u j , Fi j k =− i

2Reui u∗

j uk ,

(2.22)

where i , j ,k = 1, ..., s. The deterministic PDE governing the evolution of the mean field u(x, t )

is obtained by taking the ensemble average of the governing equation (2.21)

∂u

∂t= F0 +CYi Y j Fi j +MYi Y j Yk Fi j k , (2.23)

18

Page 27: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

2.3. Stochastic energy transfers

where CYi Y j (t ) = E [Yi Y j ] is the time-dependent covariance matrix and MYi Y j Yk (t ) = E [Yi Y j Yk ]

is the time-dependent matrix of third-order moments of the stochastic coefficients. Next, we

project the governing equation (2.21) onto each of the modes ui (x, t ). Using the DO condition,

the orthonormality of the modes and the zero-mean property of the coefficients, we get a set

of s coupled stochastic differential equations (SDEs) for the stochastic coefficients Yi (t ;ω)

dYi

dt= Ym ⟨Fm ,ui ⟩+ (YmYn −CYm Yn )⟨Fmn ,ui ⟩+ (YmYnYl −MYm Yn Yl )⟨Fmnl ,ui ⟩. (2.24)

Finally, we multiply the governing equation (2.21) with each of the stochastic coefficients

Yi (t ;ω), we apply the ensemble average operator and we use the SDEs for the stochastic

coefficients to obtain a set of s coupled deterministic PDEs governing the evolution of the DO

modes ui (x, t )∂ui

∂t= Hi −⟨Hi ,u j ⟩u j , (2.25)

where Hi is a complex deterministic field defined by Hi = E [L [u]Yk ]C−1Yi Yk

with L [u] the

right-hand side of the governing equation (2.21), and is expressed as

Hi = Fi +MYm Yn Yk C−1Yi Yk

Fmn +MYm Yn Yl Yk C−1Yi Yk

Fmnl , (2.26)

with MYi Y j Yk Yl (t) = E [Yi Y j Yk Yl ] the time-dependent matrix of fourth-order moments. The

mean u(x, t), the modes ui (x, t) and the stochastic coefficients Yi (t ;ω) are initialized at t0

through a KL expansion (2.12) of the initial condition u0(x;ω). We have thus derived an exact

set of fully coupled evolution equations for these quantities, in the sense that no approximation

other than the truncation at finite size of the DO expansion has been used.

2.3 Stochastic energy transfers

The spread of the stochastic fluctuations along each directions ui (x, t ) of the stochastic sub-

space Vs is varying with time, owing to flows of energy (variance) between modes and the

mean. In this section, we derive expressions for these rates of stochastic energy transfers

between a given DO mode, the mean and the other modes. Recall from equation (1.8) that the

potential and kinetic energies of the wavetrain are equal to leading order, so that the stochastic

total energy over the domain D = [0,L] can be expressed as

H (t ;ω) =∫ L

0η(x, t ;ω)2dx =

∫ L

0Reu(x, t ;ω)e i (x−t )2dx ' 1

2

∫ L

0|u(x, t ;ω)|2dx (2.27)

where the energy has been made nondimensional with ρg /k30 , and the last equality follows

from the slow space variation of u(x, t ;ω). The last term can be expressed in terms of the inner

product, so that we can make use of the DO expansion (2.13) to decompose the average energy

in the solution as contributions from the mean and the modes

E (t ) = E [H ] = 1

2E [⟨u,u⟩] = 1

2E [⟨u +Yi ui , u +Yi ui ⟩] = 1

2

(‖u‖2 +E [Yi Yi ])

(2.28)

19

Page 28: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 2. NLS equation under the DO framework

where the last equality follows from the orthonormality of the DO modes, and shows that the

energy contained in the modes is equal to the variance E [Y 2i ] of the stochastic coefficients,

while the term ‖u‖2 = ⟨u, u⟩ represents the energy contained in the mean. In order to study

the flow of energy between the different modes and the mean, we use equation (2.24) for the

stochastic coefficients to write the rate of change of the stochastic energy contained in mode i

εi = 1

2

d

dtE [Y 2

i ] = E

[Yi

dYi

dt

]= Ai i CYi Yi +Bi mn MYi Ym Yn +Ci mnl MYi Ym Yn Yl , (2.29)

(no sum on i ), where we have assumed that the covariance matrix CYi Y j has been diagonalized

at the present time instant (see Section 2.4.3 for the details) so that the stochastic coefficients

are uncorrelated. Ai i , Bi mn and Ci mnl are the deterministic fields appearing in equation (2.24)

Ai i =⟨− i

8

∂2ui

∂x2 − i u Reuu∗i − i

2|u|2ui ,ui

⟩, (2.30)

Bi mn =⟨− i

2Reumu∗

nu − i Reuu∗mun ,ui

⟩, (2.31)

Ci mnl =⟨− i

2Reumu∗

nul ,ui

⟩. (2.32)

The term Ai i can be considerably simplified since⟨− i

8

∂2ui

∂x2 ,ui

⟩= Re

∫ L

0

i

8

∂2ui

∂x2 u∗i dx

= Re

− i

8

∂ui

∂xu∗

i

∣∣∣∣L

0+

∫ L

0

i

8

∂ui

∂x

∂u∗i

∂xdx

= 0, (2.33)

and ⟨− i

2|u|2ui ,ui

⟩= Re

∫ L

0

i

2|u|2ui u∗

i dx

= 0, (2.34)

where we have taken advantage of the periodicity in the boundary conditions and the realness

of the inner product. After some work, we obtain the following expressions

Ai i =−Re

∫ L

0

i

2u2u∗

i2dx

, (2.35)

Bi mn =−Re

∫ L

0

i

2[ uumu∗

n + uu∗mun + u∗umun ]u∗

i dx

, (2.36)

Ci mnl =−Re

∫ L

0

i

2umu∗

nul u∗i dx

. (2.37)

By inspection, we observe that the linear term Ai i represents the rate of energy transfer

between the mean and mode i . The term Bi mn indicates modal energy production due to the

simultaneous interaction of mode i with the mean and two other modes, while Ci mnl involves

the interaction of mode i with three other modes. These nonlinear ‘four-mode interactions’

arise due to the cubic term in the NLS equation and depend on the non-Gaussian statistics of

the system, for they are associated with the high-order moments MYi Ym Yn and MYi Ym Yn Yl (note

that, however, for Gaussian statistics there can still be nonlinear energy transfers localized in

phase space that don’t manifest in the variance). The modal energy production (in terms of

20

Page 29: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

2.4. Numerical implementation

variance) of mode i thus boils down to

εi = εmean→i +εmean,mn→i +εmnl→i , (2.38)

with the following linear and nonlinear contributions

εmean→i = Ai i CYi Yi , εmean,mn→i = Bi mn MYi Ym Yn , εmnl→i =Ci mnl MYi Ym Yn Yl , (2.39)

(no summation on i ). Finally, note that there is no energy dissipation and the sum of the

average energy contained in the mean and the modes is conserved

dE

dt= E

[⟨∂u

∂t,u

⟩]= E

[⟨− i

8

∂2u

∂x2 ,u

⟩]+E

[⟨− i

2|u|2u,u

⟩]= E

[Re

∫ L

0

i

8

∂2u

∂x2 u∗dx

]+E

[Re

∫ L

0

i

2|u|2uu∗dx

]= E

[Re

− i

8

∂u

∂xu∗

∣∣∣∣L

0+

∫ L

0

i

8

∂u

∂x

∂u∗

∂xdx

]+E

[Re

∫ L

0

i

2|u|4dx

]= 0,

(2.40)

which is the well-known energy conservation property of the NLS equation (Zakharov & Shabat,

1972). Therefore, the average variation in total stochastic energy d/dt∑s

i=1 E [Y 2i ]/2 =∑s

i=1 εi

is only caused by interactions between the mean and the modes, and we anticipate the sum

of all nonlinear interactions between the modes to be zero. Indeed we have Ci mnl =−Cmi l n ,

leading to∑s

i=1 εmnl→i = 0.

2.4 Numerical implementation

The DO equations for the mean (2.23), the modes (2.25) and the coefficients (2.24) are imple-

mented numerically in the MATLAB software and solved in a coupled fashion. To increase

the speed efficiency, we implemented specific parts of the code in the C language through the

MEX interface. The various details and hurdles associated with the numerical implementation

are given in the subsections that follow, and an overview of the steps followed by the code is

provided in the last subsection.

2.4.1 Numerical schemes

The deterministic PDEs for the mean and the modes are discretized on a grid of size 1024 points

and a semi-implicit Euler scheme is used for time advancement of the solution. Indeed, the

diffusion operator that appears in the ‘forcing’ fields in equation (2.22) is built with a second-

order finite-difference scheme and is treated implicitly when it appears in the linear terms in

equations (2.23) and (2.25), while the nonlinear terms are treated explicitly. The set of SDEs

21

Page 30: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 2. NLS equation under the DO framework

for the stochastic coefficients is solved using a Monte-Carlo method with 103 to 104 particles

and is advanced in time with a 4th-order Runge-Kutta scheme, using a nondimensional time

step of 0.01. Finally, it should be mentioned that while the DO condition (2.14) implies the

preservation of the orthonormality of the modes, numerical rounding errors lead to a deviation

from that state. Therefore, orthonormality is enforced by applying a stabilized Gram-Schmidt

process to the modes at each time step, and adjusting the stochastic coefficients for the new

basis so that the solution itself remains the same.

Since the stochastic coefficients are expressed as a large ensemble of realizations, the compu-

tation of various statistical quantities such as the variance or the joint probability distribution

of the coefficients is straightforward. In addition, this allows for the direct recovery from

the DO expansion (2.13) of the complex envelope solution u(x, t ;ω) corresponding to any

ensemble member, enabling the study of individual envelope realizations or statistics such as

the probability density function of the surface elevation.

Note that provided s is low enough, the DO method can lead to significantly increased com-

putational efficiency compared to a Monte-Carlo simulation of the governing NLS equation

for all realizations, as the mean and the modes only require the solution of s +1 expensive

PDEs, while the stochastic coefficients that need to be evolved for all realizations are given by

a simpler s-dimensional ODE.

2.4.2 Initial condition formulation

It was seen in equation (2.12) that in general the initial condition is formulated in terms of a

truncated KL expansion at time t0

u(x, t0;ω) = u(x, t0)+s∑

i=1Yi (t0;ω)ui (x, t0), s <∞, (2.41)

where u(x, t0) is the initial mean, Yi (t0;ω) are the zero-mean initial stochastic coefficients and

ui (x, t0) are the initial set of orthonormal modes. In practice, instead of computing the initial

modes from the eigenvalue problem (2.9) for a given autocorrelation function Rt0 (x, y), the

computation is initialized by directly assigning a shape to the modes and realizations of a

given probability distribution to the random coefficients. In general, we formulate the initial

condition in terms of a Fourier series with coefficients having random modulus and phase

u(x, t0;ω) = u(x, t0)+N∑

n=1An(ω)e iθn (ω)e i∆kn x , N <∞, (2.42)

where the modulus An(ω) and phase θn(ω) follow a desired probability distribution and N is

the finite number of Fourier modes that are present in the initial condition. However, since

we use the real inner product (2.17), we can only assign real values to the initial stochastic

22

Page 31: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

2.4. Numerical implementation

coefficients Yi (t0;ω). This issue can be overcome by expanding (2.42) as

u(x, t0;ω) = u(x, t0)+N∑

n=1An(ω)cosθn(ω)e i∆kn x +

N∑n=1

An(ω)sinθn(ω)e i (∆kn x+π/2), (2.43)

resulting in an expansion similar to the DO initial condition (2.41), where the DO modes and

the real-valued stochastic coefficients are given by

Y2n−1(t0;ω) = An(ω)cosθn(ω), u2n−1(x, t0) = e i∆kn x ,

Y2n(t0;ω) = An(ω)sinθn(ω), u2n(x, t0) = e i (∆kn x+π/2),(2.44)

and a number of modes s = 2N is required because of the realness of the stochastic coefficients

(as was thoroughly discussed in Section 2.2.2, enabling the use of complex coefficients would

eliminate this drawback). Note that we indeed have ⟨e i∆kn x ,e i (∆kn x+π/2)⟩ = 0, confirming the

fact that a given complex Fourier mode is spanned by two orthogonal directions under the

real-valued inner product. The values given in (2.44) will be used to initiate the quantities in

the DO solution for initial conditions of the type (2.42).

2.4.3 Diagonalization of the covariance matrix

In the KL expansion (2.7), the stochastic coefficients are uncorrelated i.e. E [Yi (t ;ω)Y j (t ;ω)] = 0

for i 6= j , which is a consequence of the fact that the directions ui (x, t) are aligned with

the principal directions of variance of u(x, t ;ω). On the other hand, in the DO expansion

(2.13) the stochastic coefficients are not constrained to remain uncorrelated, so that even

when the simulation is initiated with uncorrelated coefficients, over a finite time they will

develop some correlation and off-diagonal terms will appear in the covariance matrix CYi Y j =E [Yi (t ;ω)Y j (t ;ω)]. As a result, the modes ui (x, t) are no longer aligned with the principal

directions of variance of the DO solution u(x, t ;ω). This problem can nonetheless be easily

overcome by diagonalizing the covariance matrix CYi Y j , which corresponds to applying a

rotation to the modes ui (x, t) such that they become aligned with the principal variance

directions and the coefficients Yi (t ;ω) become uncorrelated, while the full solution u(x, t ;ω)

remains intact.

Let us describe the reasoning behind the procedure. Suppose that we have modes ui and that

the covariance matrix C =CYi Y j has off-diagonal elements at a given time instant. Since it is

real symmetric, we are assured of the existence of the following diagonal decomposition

C =V DV T (2.45)

where T denotes the transpose, V is formed by the eigenvectors of C hence is orthogonal, and

D is a diagonal matrix containing the eigenvalues of C . Since V is orthogonal, it can be used

as a rotation matrix and we define a new basis with u′i = umVmi . Note that we have

⟨ui ,u′j ⟩ = ⟨ui ,umVm j ⟩ =Vm j ⟨ui ,um⟩ =Vi j , (2.46)

23

Page 32: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 2. NLS equation under the DO framework

We first show that the new basis u′i is orthonormal

⟨u′i ,u′

j ⟩ = ⟨umVmi ,unVn j ⟩ =Vmi Vn j ⟨um ,un⟩ =Vmi Vm j =V Tj mVmi = δi j . (2.47)

Then, the coefficients Y ′i in the new basis are given by the following projection

Y ′i = ⟨Y j u j ,u′

i ⟩ = Y j ⟨u j ,u′i ⟩ = Y j V j i , (2.48)

and they are uncorrelated since

CY ′i Y ′

j= E [Y ′

i Y ′j ] = E [YmVmi YnVn j ] =V T

i mCmnVn j = Di j = δi jλi , (2.49)

where λi are the eigenvalues of C and the diagonal elements of D . We have therefore shown

that by diagonal decomposition of the covariance matrix CYi Y j , we can always define a new

rotated basis such that the stochastic coefficients in the new basis become uncorrelated.

The new directions u′i (x, t ) correspond to the directions of principal variance of the solution

u(x, t ;ω). We apply this procedure every time the solution is plotted or saved.

2.4.4 Overview of the code structure

Here we provide an overview of the structure of the code. The DO solution is first initialized by

assigning a value to the mean field u(x, t ), the modes ui (x, t ) and the stochastic coefficients

Yi (t ;ω), for example by following (2.44). We then apply a Gram-Schmidt process to the initial

modes to make sure they are orthonormal. The code then enters a loop where the solution is

advanced in time, and that consists of the following steps, in order:

1. The zero-mean property of the stochastic coefficients is enforced to avoid deviations

due to rounding errors.

2. The covariance matrix CYi Y j and higher-order moments matrices MYi Y j Yk and MYi Y j Yk Yl

of the stochastic coefficients are calculated. The calculation is implemented in C

(through the MEX interface) and takes advantage of the symmetries in the moments.

3. The deterministic ‘forcing’ fields in equation (2.22) are calculated.

4. The stochastic coefficients are advanced in time using a 4th-order Runge-Kutta scheme,

where the calculation of the right-hand side of equation (2.24) is implemented in C

(through the MEX interface).

5. The mean field is advanced in time through equation (2.23) and a semi-implicit Euler

scheme, where the diffusion term is treated implicitly in the linear term only.

6. The deterministic ‘forcing’ fields in equation (2.26) are calculated.

7. The modes are advanced in time through equation (2.25) and the same semi-implicit

Euler scheme as for the mean.

24

Page 33: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

2.4. Numerical implementation

8. The modes are orthonormalized using a stabilized Gram-Schmidt process and the

adjusted stochastic coefficients are calculated.

9. At some of the time steps, the solution is plotted and/or saved after the modes have

been rotated following the procedure from Section 2.4.3 and leading to uncorrelated

stochastic coefficients.

25

Page 34: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework
Page 35: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 3

Preliminary results and validation

In this chapter, we illustrate the use of the DO reduced-order model introduced in the previous

chapter, by presenting simulation results for situations that are well documented in the

literature. These situations will also provide us with a way to benchmark our results and

validate the accuracy our DO reduced-order equations. Specifically, we simulate in Section 3.1

the evolution of a uniform wavetrain undergoing ‘semi-stochastic’ Benjamin-Feir instability

and Fermi-Pasta-Ulam recurrence. In Section 3.2, we compute the stochastic evolution of a

random Gaussian spectrum of waves and investigate the properties of the resulting solution.

3.1 Idealized Benjamin-Feir instability

Recall from Section 1.3 that a uniform wavetrain, represented by a spatially constant enve-

lope u(x, t0) = a0, is linearly unstable to small Fourier mode perturbations with modulation

wavenumber in the range 0 < ∆k < ∆kc = 2p

2a0. If there is only one prescribed unstable

modulation wavenumber ∆k and if its harmonics fall outside the unstable regime (that is

∆k >∆kc /2), then the long time evolution of this perturbation is very well understood. After

initially undergoing Benjamin-Feir (BF) instability, the unstable modulation will grow and

decay repeatedly in a Fermi-Pasta-Ulam (FPU) recurrence cycle that is illustrated in Figures

1.2 and 1.3. Since this behavior is the same regardless of the phase difference between the

modulation and the carrier wave, this situation therefore provides us with a simple frame-

work for the illustration and validation of our DO stochastic equations. Indeed, we can then

consider random initial conditions consisting of a constant envelope a0 perturbed by Fourier

modes with deterministic small amplitude but random phase

u(x, t0;ω) = a0 +N∑

n=1Ane iθn (ω)e i∆kn x , N <∞, (3.1)

and we expect that all realizations will evolve similarly and according to the deterministic

results. The stochastic fluctuations introduced by the phase randomness should persist and

27

Page 36: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 3. Preliminary results and validationM

ean

-0.1

0

0.1

t = 0

Mod

e 1

-0.1

-0.05

0

0.05

0.1

Mod

e 2

-0.1

-0.05

0

0.05

0.1

Mod

e 3

-0.1

-0.05

0

0.05

0.1

x0 10 20 30

Mod

e 4

-0.1

-0.05

0

0.05

0.1

-0.1

0

0.1

t = 800

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

x0 10 20 30

-0.1

-0.05

0

0.05

0.1

-0.1

0

0.1

t = 1400

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

x0 10 20 30

-0.1

-0.05

0

0.05

0.1

Figure 3.1 – Mean and DO modes at various times for the stochastic BF instability and FPUrecurrence with random phase in the initial modulation. Both the complex envelope modulus(blue) and real part (orange) are represented. Note that while the domain size is L = 50 ·2π, weonly plot the solution over a portion of the domain corresponding to the wavelength of thelinearly most unstable wavenumber ∆km = 2a0 = 0.2.

grow, in such a way that the modes should reveal the dominant components of the dynamics.

We therefore consider initial condition (3.1) with a0 = 0.1. We choose to use N = 2 Fourier

modes of wavenumbers ∆k1 = 0.2 and ∆k2 = 0.4, in such a way that the first Fourier mode cor-

responds to the linearly most unstable wavenumber (given by ∆km = 2a0) while its harmonics

and the second Fourier mode are stable. The random phases θn(ω) are drawn from indepen-

dent and uniform distributions on [0,2π], and the deterministic amplitudes An are assigned

the infinitesimal value 0.0036. For the initial DO expansion (2.12), the mean is assigned the

uniform wave component u(x, t0) = a0, while the modes ui (x, t0) and stochastic coefficients

Yi (t0;ω) carry the stochastic perturbations and are initialized with the relations (2.44). Note

that each Fourier mode has to be represented with two DO modes because of the real-valued

coefficients, resulting in a total of 4 DO modes. A periodic domain of size L = 50 ·2π is used

but its size doesn’t affect the solution since the latter is periodic and non-localized.

The solution for the mean and the modes at various times is plotted in Figure 3.1, where

both the modulus (blue) and real part (orange) are represented. The smallest wavenumber

∆k1 = 0.2 present in the initial condition implies that the solution will remain periodic with

period 2π/∆k1, therefore we only show a portion of the total domain corresponding to this

28

Page 37: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

3.1. Idealized Benjamin-Feir instability

t0 200 400 600 800 1000 1200 1400 1600 1800 2000

E[Y

i2]

10 -3

10 -2

10 -1

100

101

MeanMode 1Mode 2Mode 3Mode 4Deterministic

Figure 3.2 – Energy of the mean ⟨u, u⟩ and the modes E [Y 2i ] for the stochastic BF instability

and FPU recurrence with random phase in the initial modulation. The dashed lines showthe corresponding energies for a deterministic simulation of an equivalent initial conditionu(x,0) = 0.1+ 0.0036cos∆km x with ∆km = 0.2, with the energies obtained as the normal-ized squared modulus of the Fourier coefficients of wavenumber ∆k = 0 (carrier wave), 0.2(unstable modulation) and 0.4 (stable harmonic) (similarly to Figure 1.3).

wavelength. In Figure 3.2, we show the energy present in the mean ⟨u, u⟩ together with the

stochastic energy present in the modes E [Y 2i ]. Focusing on the modes at t = 800 in Figure

3.1, we observe that the modes keep their initial pairing indicative of a randomness in the

phase of the associated structure. Modes 1 and 2 represent the same linearly most unstable

modulation ∆k1 = 0.2 as was assigned in the initial condition, while modes 3 and 4 represent

the same stable modulation ∆k2 = 0.4. The spatially constant mean still represents a uniform

carrier wave. From Figure 3.2, it is observed that the stochastic energy in the linearly unstable

modes 1 and 2 grows, saturates then decays. The stable modes 3 and 4 are slaved to the

unstable modulation and experience the same process from t ∼ 400, albeit to a lesser degree.

Meanwhile, the energy of the mean follows the inverse tendency since the modes are growing

at its expense. Note that the modulations grow as stochastic fluctuations (since they are

represented in the modes) because of the phase randomness that they have been assigned in

the initial condition.

These observations are in perfect accordance with the deterministic BF instability and sub-

sequent FPU recurrence shown in Figures 1.2 and 1.3, where the unstable modulations

are growing at the expense of the uniform carrier wave, before decaying. For a quantita-

tive comparison, we computed the solution to a deterministic initial condition of the form

u(x,0) = 0.1+0.0036cos∆km x where ∆km = 0.2, i.e. similar in structure to one realization of

the random DO initial condition (3.1) (only without the phase randomness of the modulation),

and we retrieve the energy of the carrier wave and the modulations by means of a Fourier trans-

form of the envelope. The normalized resulting squared modulus of the Fourier coefficients

for ∆k = 0 (carrier wave), 0.2 (unstable modulation) and 0.4 (stable harmonic) are shown as

dashed lines in Figure 3.2. The agreement between the stochastic DO computation and the

deterministic calculation is remarkable, particularly the maximum energy of the unstable

modulation ∆k = 0.2 (contained in the DO modes 1 and 2), and the time at which it occurs.

29

Page 38: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 3. Preliminary results and validation

Figure 3.3 – Stochastic attractor at various times of the stochastic BF instability and FPUrecurrence with random phase in the initial modulation. The attractor is represented in termsof the 3D scatter plot of the first three stochastic coefficients for all realizations. In addition,each realization is assigned a color indicative of the maximum value of its correspondingsurface elevation η(x, t ;ω) = Reu(x, t ;ω)e i (x−t/2).

However, from time t ∼ 1100, the DO solution is observed in Figure 3.2 to deviate from its

deterministic counterpart, which can also be seen in Figure 3.1 in the shapes of the modes

at t = 1400 that lose any kind of regular structure. Moreover, from t ∼ 1900, the energy in the

modes begin oscillating, a sign that the numerical solution is collapsing. We tried refining

the mesh, using a smaller time step or a larger number of Monte-Carlo realizations, but to

no avail. The source of the problem thus remains unknown and could be related either to

the numerics or to the DO equations themselves. It should be noted that the NLS equation

has been found to cause numerical instabilities in similar situations (Herbst & Ablowitz, 1989;

Ablowitz & Herbst, 1990), while numerical instabilities have also been encountered in other

systems using reduced-order models based on the KL expansion (Kirby & Armbruster, 1992).

Since the stochastic coefficients are expressed as a large ensemble of realizations, it is possible

to visualize the time-dependent structure in phase space of the ensemble solution. In Figure

3.3, we therefore display the stochastic attractor of the solution, represented as the 3D scatter

plot of the first three stochastic coefficients for all realizations. The coefficients associated

to modes 1 and 2 (containing the linearly unstable modulation) are observed to be of the

form Y1(t ;ω) = A(t)cosθ(ω) and Y2(t ;ω) = A(t)sinθ(ω), with amplitude A(t) approximately

equal for all realizations and undergoing growth then decay. This shows that the unstable

modulation contained in the first two modes is growing and decaying at a similar rate for all

realizations, but with a random phase shift θ(ω), as can be expected since all realizations have

been assigned the same initial perturbation amplitude in the initial condition (3.1).

In addition, each realization in Figure 3.3 is assigned a color indicative of the maximum value

of the corresponding surface elevation, obtained as η(x, t ;ω) = Reu(x, t ;ω)e i (x−t/2) where the

envelope u(x, t ;ω) can be reconstructed from the DO expansion (2.13). Since the amplitude

of the modulation is the same between all members of the ensemble, differences in surface

elevation only arise from the stochastic phase difference between the modulation and the

30

Page 39: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

3.1. Idealized Benjamin-Feir instability

mem

ber

with

max

imum

2

-0.2-0.1

00.10.2

t = 0

x0 10 20 30

mem

ber

with

min

imum

2

-0.2-0.1

00.10.2

-0.2-0.1

00.10.2

t = 800

x0 10 20 30

-0.2-0.1

00.10.2

-0.2-0.1

00.10.2

t = 1100

x0 10 20 30

-0.2-0.1

00.10.2

Figure 3.4 – Two realizations at various times of the stochastic BF instability and FPU recur-rence with random phase in the initial modulation. Both the surface elevation (blue) andcomplex envelope modulus (orange) are represented. The top row shows the realization withmaximum surface elevation η(x, tω) at t = 800, while the bottom row shows the one withminimum surface elevation at that time. Note that we only plot a portion of the domaincorresponding to the wavelength of the linearly most unstable wavenumber ∆km = 2a0 = 0.2.

carrier wave. To illustrate this, we plot in Figure 3.4 the surface elevation (blue) and envelope

modulus (orange) at various times of the two realizations that have the maximum (top row)

and minimum (bottom row) surface elevation at t = 800. Their envelope is indeed observed

to be modulated to a similar degree, with differences in surface elevation resulting from the

phase shift between complex envelope and carrier wave.

Finally, in Figure 3.5 we illustrate the flows of stochastic energy between modes and the

mean. Using the expressions derived in Section 2.3, we calculate the energy production in

every mode due to (i) its linear interaction with the mean (first column), (ii) its nonlinear

simultaneous interaction with the mean and two other modes (second column) and (iii) its

nonlinear interaction with three other modes (third column). In the first column it is seen that

DO modes 1 and 2 (corresponding to the linearly unstable modulation) are from the onset

receiving energy from the mean in a linear fashion, before giving it back after saturation of

the modulation. The same happens for DO modes 3 and 4 (corresponding to the linearly

stable modulation), although in this case the transfer of energy is delayed since they are

initially linearly stable. The third column shows that there are nonlinear exchanges of energy

occurring between the unstable and stable modulations, although these are one to two orders

of magnitude smaller than the exchanges of energy with the mean. Sapsis (2013) proved

that linear transfers of energy from the mean to the modes result in the uniform increase or

decrease of the attractor in phase space, while only nonlinear energy exchanges are able to

cause local changes in its shape. This is indeed the case here, since the attractor is mainly

expanding and contracting uniformly in phase space and energy transfers mostly occur in a

linear fashion.

31

Page 40: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 3. Preliminary results and validation

Mod

e 1

#10 -3

-2

-1

0

1

2"mean!i

Mod

e 2

#10 -3

-2

-1

0

1

2

Mod

e 3

#10 -4

-2

-1

0

1

2

t200 400 600 800 1000 1200

Mod

e 4

#10 -4

-2

-1

0

1

2

#10 -4

-5

0

5"mean;mn!i

#10 -4

-5

0

5

#10 -4

-4

-2

0

2

4

t200 400 600 800 1000 1200

#10 -4

-4

-2

0

2

4

#10 -5

-1

-0.5

0

0.5

1"mnl!i

#10 -5

-2

-1

0

1

#10 -5

-1

0

1

2

t200 400 600 800 1000 1200

#10 -5

-1

-0.5

0

0.5

1

Figure 3.5 – Modal energy production in every mode due to (i) linear interaction with themean flow (first column), (ii) nonlinear simultaneous interaction with the mean and twoother modes (second column) and (iii) nonlinear interaction with three other modes (thirdcolumn) for the stochastic BF instability and FPU recurrence with random phase in the initialmodulation. Recall that the solution is only valid up to t ∼ 1100.

3.2 Random Gaussian wavenumber spectrum

The Benjamin-Feir instability provides an idealized framework for the investigation of instabil-

ities of water waves. In practice, however, the ocean surface is not merely a uniform wavetrain

and energy is rather distributed over a range of wavenumbers around the carrier wave. There-

fore, a more realistic setup would be to consider the evolution of a narrow-band wave field

consisting of a Gaussian random distribution of waves around the carrier wavenumber. The

resulting non-uniform wavetrain can be written in terms of the complex envelope as

u(x, t0;ω) =N /2∑

n=−N /2+1

√2∆k1F (∆kn)e iθn (ω)e i∆kn x , (3.2)

where the Fourier coefficients have random uncorrelated phases θn(ω) drawn from a uniform

distribution on [0,2π], and deterministic amplitude proportional to the square root of the

following Gaussian wavenumber spectrum

F (∆kn) = ε2

σp

2πexp

(−∆k2

n

2σ2

), (3.3)

where ∆kn = n∆k1 is the modulation wavenumber with ∆k1 the wavenumber discretization,

ε is the average wave steepness and σ is the relative bandwidth of the spectrum. The av-

erage steepness is defined as the standard deviation of the surface elevation, and it can be

32

Page 41: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

3.2. Random Gaussian wavenumber spectrum

verified that we indeed have ε=√

E [η(x, t0;ω)2] at any x location. Note that the wave field

defined in (3.2) leads to a Gaussian distribution of the surface elevation since the Fourier wave

components are uncorrelated.

Based on the NLS equation, Alber (1978) investigated the stability of the random wave field

(3.2) under a narrow-bandwidth and near-Gaussian statistics approximation, and found that

the ensemble-averaged Gaussian spectrum (3.3) is stable when the ratio of the wave steepness

to the relative bandwidth, defined as the Benjamin-Feir index

BFI = 2p

σ, (3.4)

is less than 1. In the opposite case, numerical simulations of the NLS equation (Janssen, 2003;

Dysthe et al., 2003) and experiments (Onorato et al., 2005) have shown that when the initial

BFI > 1, the ensemble-averaged spectrum relaxes on a time scale of O (ε−2) to a wider stable

spectrum (characterized by a final time BFI ∼ 1), while in the meantime the so-called ‘random

version’ of the Benjamin-Feir instability occurs. The latter manifests as the focusing of localized

wave packets in the irregular wavetrain due to the increased effect of nonlinearities, resulting

in large coherent structures (similar to those observed in the deterministic Benjamin-Feir

instability in Figure 1.2) and creating heavy-tailed statistics for the surface elevation.

The random Gaussian spectrum of waves (3.3) thus provides us with a full stochastic setting in

which we can benchmark our DO equations. We assign the stochastic initial condition (3.2) to

our initial DO expansion (2.12), and with a single DO simulation we study the properties of

the resulting ensemble solution and compare them with results obtained from Monte-Carlo

simulations in the literature.

Specifically, we consider initial condition (3.2) in the case of N = 10 Fourier modes with a

wavenumber discretization ∆k1 = 0.06. The spectrum is assigned a fixed initial width σ= 0.1,

and we consider seven different values for the wave steepness ε ranging from 0.025 to 0.1,

giving a range of values for the initial BFI from 0.72 to 2.87 (in this regard note that most

sea states have a BFI < 1, Dysthe et al., 2008). The resulting discrete Gaussian spectrum

(3.3) and its continuous equivalent are shown in the left-hand side plot of Figure 3.6 for the

case BFI = 1.43. The importance of having random and uncorrelated phases is shown in the

right-hand side plot, where we compare two realizations of the surface elevation (blue) and

envelope modulus (orange) corresponding to this Gaussian spectrum, either when all phases

are equal to zero (top) or with random uncorrelated phases (bottom). The randomness of

the phases θn(ω) creates mixing between the different wavenumbers and leads to the desired

irregularity for the surface elevation. For the initial DO expansion (2.12), each Fourier mode

in (3.2) is assigned to two DO modes ui (x, t0) as per the relations (2.44), since two real-valued

stochastic coefficients Yi (t0;ω) are necessary to reproduce the phase randomness of the

Fourier coefficients. Without loss of generality, we decide to assign a deterministic phase equal

to zero to the modulation wavenumber zero so that we can assign it to the mean, resulting in a

total number of 18 DO modes. As before, we use a periodic domain of size L = 50 ·2π. Note

33

Page 42: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 3. Preliminary results and validation

"k-1 -0.5 0 0.5 1

F("

k)

0

0.002

0.004

0.006

0.008

0.01

0.012Gaussian spectrum with BFI = 1.43

discretecontinuous

x0 20 40 60 80 100

rand

om p

hase

s

-0.2

-0.1

0

0.1

0.2

zero

pha

ses

-0.2

-0.1

0

0.1

0.2|u| (orange) and 2 (blue)

Figure 3.6 – Initial condition for the DO simulation of a Gaussian random wave field forthe case BFI = 1.43. Left, discretized Gaussian wavenumber spectrum and its continuousequivalent. Right, two corresponding realizations of the surface elevation η(x, t0;ω) (blue) andenvelope modulus |u(x, t0;ω)| (orange), either when all phases are equal to zero (top) or withrandom uncorrelated phases (bottom).

k0 1 2

E[|F

[2]|2

]

0

20

40

60

80

100t0, BFI = 0.72

k0 1 2

tf

k0 1 2

E[|F

[2]|2

]

0

200

400

600

800

1000

1200

1400t0, BFI = 2.87

k0 1 2

tf

Figure 3.7 – Initial and final time wavenumber spectra E [|F [η]|2] for initial BFI = 0.72 (left)and BFI = 2.87 (right). The final time is taken as t f = 2/ε2.

that the minimum wavenumber allowed by the domain is ∆kmi n = 2π/L = 0.02, thus equal

to one third the wavenumber discretization ∆k1 = 0.06 that we assign to our initial Gaussian

spectrum. We could have used a finer wavenumber discretization ∆k1, but spanning the same

spectral width σ= 0.1 would have resulted in an unreasonably high number of DO modes and

prohibitive computational cost.

3.2.1 Statistical properties of the DO solution

We begin by investigating the stability properties of the ensemble-averaged spectrum with

respect to the BFI. In Figure 3.7, we plot the initial and final time ensemble-averaged wavenum-

ber spectra given by E [|F [η]|2] for two different initial values of BFI = 0.72 (left) and BFI = 2.87

(right). Since spectral changes for unstable values of the BFI are supposed to occur on a

timescale of O (ε−2) (Dysthe et al., 2003), results are shown for a final time defined as t f = 2/ε2.

In accordance with results in the literature, it is observed that the case BFI = 0.72 < 1 is stable

while the case BFI = 2.87 > 1 is unstable and relaxes to a new spectrum that appears to be

34

Page 43: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

3.2. Random Gaussian wavenumber spectrum

Initial time BFI0 0.5 1 1.5 2 2.5 3

Fin

al ti

me

BF

I

0

0.5

1

1.5

DO simulationJanssen (2003)Dysthe et al. (2003)

Figure 3.8 – Final versus initial BFI, where the final value is calculated using equations (3.4)and (3.5). Results from Monte-Carlo simulations of the NLS equation from Janssen (2003) andDysthe et al. (2003) are also shown.

stable. Note that the spikes reflect the fact that the wavenumber resolution ∆kmi n of the

domain is one third the wavenumber discretization ∆k1 used in the initial condition. These

spectrum relaxation results are shown in a more quantitative way and for all considered initial

values of the BFI in Figure 3.8, where we plot the final time BFI as a function of the initial one.

The final time BFI, indicative of the broadening of the spectrum, is calculated with equation

(3.4) where the final time steepness and spectral width are given by

ε(t f ) =√

E [η(x, t f ;ω)2] and σ(t f ) =∫∆k2E [|F [η(x, t f ;ω)]|2]d∆k∫

E [|F [η(x, t f ;ω)]|2]d∆k. (3.5)

While no clear bifurcation is found, it is indeed observed that values of the BFI initially larger

than 1 lead to a broadening of the spectrum in such a way that the final time BFI approaches 1.

The results from the DO simulations are also compared with Monte-Carlo simulations of the

NLS equation from the literature (Janssen, 2003; Dysthe et al., 2003), and show a reasonable

agreement with these authors.

Apart from the relaxation of the spectrum, a value of the BFI larger than 1 also leads to

increased focusing of wave packets due to the increased effect of nonlinearities, resulting

in a frequent occurrence of extreme waves. In Figure 3.9, we plot the ensemble-averaged

probability density function of the surface elevation normalized by ε =√

E [η(x, t ;ω)2], at

initial and final times and for initial BFI = 0.72 (left) and initial BFI = 1.43 (right). For reference,

the corresponding Gaussian distribution in each case is also shown in orange dotted line. The

deviation from normality of the final time pdf for to BFI = 1.43 is indicative of strong nonlinear

effects resulting in a large occurrence of focusing wave groups, in accordance with results

from the literature. We can get a more quantitative picture of the probability of extreme waves

with the kurtosis, defined as

C = E [η(x, t ;ω)4]

3E [η(x, t ;ω)2]−1, (3.6)

which is a measure of the deviation of the pdf of the surface elevation from the Gaussian

distribution. A value of 0 corresponds to a Gaussian surface, while a value greater than 0

35

Page 44: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 3. Preliminary results and validation

2/0-5 0 5

pdf

10 -6

10 -4

10 -2

100t0, BFI = 0.72

2/0-5 0 5

tf

2/0-5 0 5

pdf

10 -6

10 -4

10 -2

100t0, BFI = 1.43

2/0-5 0 5

tf

Figure 3.9 – Probability density function at initial and final times of the surface elevationnormalized by ε=

√E [η(x, t ;ω)2], for initial BFI = 0.72 (left) and BFI = 1.43 (right). The final

time is taken as t f = 2/ε2. For reference, the corresponding Gaussian distribution is alsoshown in orange dotted line in each case.

Initial time BFI0 0.5 1 1.5 2 2.5 3

Fin

al ti

me

kurt

osis

-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

Figure 3.10 – Final time kurtosis versus initial BFI, where the kurtosis is calculated usingequation (3.6) and is representative of the deviation from normality of the distribution of thesurface elevation.

is representative of a higher probability of extreme waves due to predominant nonlinear

effects. The final time kurtosis is plotted in Figure 3.10 versus the initial BFI, and is observed

to increase with the value of the BFI, corroborating results from the literature.

We have therefore shown that the DO solution for the Gaussian wavenumber spectrum gives

accurate statistical (i.e. ensemble-averaged) results compared with Monte-Carlo simulations

from the literature. It should however be noted that solving the DO equations for this situation

of a random Gaussian spectrum is computationally very costly, because of the large number

of DO modes required to represent the large number of random wavenumbers needed in the

initial condition (18 in our case). For computations directed towards the obtention of purely

statistical results of the kind that were shown in this section, it appears preferable to use a

traditional Monte-Carlo method. Nevertheless, we observed Section 3.1 that a strength of the

DO method lies in its ability to reveal the time-dependent dominant directions of stochastic

fluctuations, plus the structure in phase space of the stochastic solution. Therefore, we

investigate in the following section the structure of our DO solutions of random Gaussian wave

fields, similarly to what was done in Section 3.1 for the idealized Benjamin-Feir instability.

36

Page 45: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

3.2. Random Gaussian wavenumber spectrum

Mea

n

-0.1

-0.05

0

0.05

0.1BFI = 0.72

Mod

e 1

-0.1

-0.05

0

0.05

0.1

Mod

e 2

-0.1

-0.05

0

0.05

0.1

Mod

e 3

-0.1

-0.05

0

0.05

0.1

x0 50 100

Mod

e 4

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1BFI = 1.00

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

x0 50 100

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1BFI = 1.43

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

x0 50 100

-0.1

-0.05

0

0.05

0.1

Figure 3.11 – Mean and first four DO modes at final time t f = 2/ε2, for different initial valuesof the BFI. Both the complex envelope modulus (blue) and real part (orange) are represented.Note that while the domain size is L = 50 ·2π, we only plot the solution over one third of thewhole domain because of the effective periodicity induced by the wavenumber discretization∆k1 = 0.06 of the initial condition.

3.2.2 Structure of the DO solution

First, we show in Figure 3.11 the final time shape of the mean and first four DO modes, for three

different values of the initial BFI, where again the final time is defined as t f = 2/ε2. Note that

only one third of the whole domain size is represented, since the wavenumber discretization

∆k1 = 3∆kmi n of the initial condition leads to a solution that has a triple periodicity within the

whole domain. The modes appear more regular when the initial BFI < 1 (corresponding to a

stable ensemble-averaged spectrum), and we note that the DO modes 1 and 2 for BFI = 0.72

correspond to sinusoidal modulations of wavenumber 0.06, equal to the Benjamin-Feir linearly

most unstable wavenumber ∆km = 2a0 for a uniform wave of amplitude a0 = 0.03, a value

close to the actual r.m.s amplitude E [a] =√

2E [η2] = 0.035 corresponding to this irregular

wave field. This being said, for higher BFI the situation is less clear and the modes are difficult

to interpret.

In Figure 3.12, we show the evolution from initial to final time of the energy present in the mean

and the modes, for the same three values of the initial BFI. It is observed that for BFI = 0.72,

the energy of the DO modes 1 and 2 (that have a regular sinusoidal shape) is set apart from the

others, confirming their ‘special status’. For higher BFI, the modes converge to an energy level

37

Page 46: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 3. Preliminary results and validation

t0 1000 2000

E[Y

i2]

0

0.02

0.04

0.06

0.08

0.1BFI = 0.72

MeanMode 1Mode 2etc

t0 500 1000 1500

0

0.05

0.1

0.15

0.2BFI = 1.00

MeanMode 1Mode 2etc

t0 200 400 600 800

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4BFI = 1.43

MeanMode 1Mode 2etc

Figure 3.12 – Evolution of the energy in the mean and the modes for different initial values ofthe BFI. Recall that the mean is initially assigned the modulation wavenumber 0.

Figure 3.13 – Stochastic attractor at final time of the solution for different initial values of theBFI. The attractor is represented in terms of the 3D scatter plot of the first three stochasticcoefficients for all realizations. In addition, each realization is assigned a color indicative ofthe maximum value of its corresponding surface elevation η(x, t ;ω) = Reu(x, t ;ω)e i (x−t/2).

that is closer with the others, in such a way that there is not really a dominant direction in the

stochastic fluctuations. This corroborates observations from Figure 3.11, where no definite

trend was observed in the shape of the modes.

Recall that the stochastic coefficients are expressed as a large ensemble of realizations, al-

lowing for the visualization in phase space of the time-dependent structure of the ensem-

ble solution. In Figure 3.13, we display the stochastic attractor of the final time solution

for the same three values of the initial BFI, represented as the 3D scatter plot of the first

three stochastic coefficients for all realizations. In addition, each realization is assigned a

color indicative of the maximum value of the corresponding surface elevation, obtained as

η(x, t ;ω) = Reu(x, t ;ω)e i (x−t/2) where the stochastic envelope u(x, t ;ω) can be reconstructed

from the DO expansion (2.13). As can be expected from the similar energy levels present in

all the modes, no particular structure can be observed and there doesn’t appear to be a clear

correlation between the maximum of the surface elevation for a given realization and the

value of the corresponding stochastic coefficients.

38

Page 47: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

3.2. Random Gaussian wavenumber spectrum

mem

ber

with

max

imum

2

-0.2

-0.1

0

0.1

0.2BFI = 0.72

x0 50 100

mem

ber

with

min

imum

2

-0.1

-0.05

0

0.05

0.1-0.2

-0.1

0

0.1

0.2BFI = 1.00

x0 50 100

-0.1

-0.05

0

0.05

0.1-0.4

-0.2

0

0.2

0.4BFI = 1.43

x0 50 100

-0.1

-0.05

0

0.05

0.1

Figure 3.14 – Two realizations of the stochastic solution at final time for different initial valuesof the BFI. The top row shows the realization with maximum surface elevation η(x, t ;ω) whilethe bottom row shows that with minimum surface elevation. Note that we only plot thesolution over one third of the whole domain because of the effective periodicity induced bythe wavenumber discretization ∆k1 = 0.06 of the initial condition.

Finally, for each of the same three values of the initial BFI, we plot in Figure 3.14 the surface

elevation (blue) and envelope modulus (orange) of the two realizations that have the maximum

(top row) and minimum (bottom row) surface elevation at the final time. The realizations

with maximum surface elevation for BFI = 1.00 and 1.43 show localized wave groups that

have focused, sucking energy from the nearby wave field. These focusing wave groups are

known to appear for BFI > 1 (Ruban, 2013) and are responsible for the heavy-tailed statistics

of the surface elevation in these higher-energy wave fields. Being able to observe them in

individual realizations of the reduced-order DO stochastic solution therefore constitutes a

check of the validity of our DO framework. However, it appears difficult to establish a precise

relationship between the shape of these wave groups and that of the modes from Figure 3.11.

This is because the focusing wave groups observed in Figure 3.14 are events of a local nature

that appear at random locations in the domain, therefore they cannot be captured as such by

the DO modes. The absence of a clear relationship between the shape of the realizations and

that of the modes also confirms that the dynamics of these irregular wave fields don’t possess

any dominant component on a global scale, as could be inferred from the similar levels of

energy present in all the modes in Figure 3.12.

As a conclusion, the advantages brought by the DO framework in the context of the stochastic

solution to a Gaussian random spectrum of waves are limited. There are mainly two inter-

connected reasons for that. First, by assigning all the stochastic components of the initial

condition into the DO modes, a high number of them is required, resulting in a high com-

putational cost as the latter mostly scales to the cube of the number of modes. Second, the

Gaussian stochastic wave field is such that energy is spread over a large number of modes and

there is no clear dominant component in the stochastic dynamics on the global scale. This

means that (i) the high number of modes required to initialize the solution is still needed as

time evolves, and (ii) the modes don’t reveal any global dominant tendency in the stochastic

fluctuations and the solution does not possess a clear structure in phase space.

39

Page 48: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 3. Preliminary results and validation

Based on these observations, it looks like the DO framework is better suited for computing and

analyzing the nonlinear evolution of a given deterministic wave field that would be assigned

to the mean u(x, t), and perturbed by stochastic perturbations that can be contained in a

reasonable number of DO modes. This is in essence what was done in Section 3.1, where in

order to study the stochastic evolution of the idealized Benjamin-Feir instability, we assigned a

uniform wavetrain to the mean and perturbed it with small stochastic perturbations contained

in the modes. At the same time, we observed in Figure 3.14 that while there are no dominant

components in the global dynamics of these irregular wave fields, local events constituting

of the focusing of localized wave groups are nevertheless present. These wave groups have

recently attracted attention in the literature (Adcock & Taylor, 2009; Cousins & Sapsis, 2015b,a)

due to the extreme waves that can result from them. Therefore, in the following chapter

we focus our attention to these spatially localized wave groups, and study their nonlinear

evolution when subjected to small stochastic perturbations.

40

Page 49: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 4

Dynamics of an extreme wave

In this chapter, we exploit the benefits of the DO framework to study the nonlinear evolution

of deterministic wave fields under small initial stochastic perturbations. Specifically, we

concentrate on the focusing behavior of spatially localized wave groups induced by nonlinear

effects of the NLS equation, resulting in extreme waves. This known phenomenon is presented

in Section 4.1. Then, in Section 4.2 we investigate the ability of the DO modes to track the

emergence of these extreme waves out of a Gaussian spectrum of background waves. Finally,

in Section 4.3 we study the structure in phase space of an idealized extreme wave (i.e. without

any background wave field) subject to small initial perturbations. The results presented in this

chapter are still a work in progress.

4.1 Nonlinear focusing of localized wave packets

In Section 3.2, we mentioned that in a random Gaussian spectrum of waves (3.2) characterized

by a BFI > 1, increased effects of nonlinearities lead to the focusing of spatially localized

wave packets of moderate amplitude, resulting in extreme waves and heavy-tailed statistics

for the distribution of the surface elevation (Ruban, 2013). An example of such an extreme

event formation in shown in Figure 4.1, where the initial wave field is generated through

a Gaussian spectrum (3.3) with random phases and BFI = 1.43. A spatially localized wave

packet around x = 140 at time 0 (top row) is observed to focus in an extreme wave and reaches

a maximum amplitude of 0.37 at t = 270 (middle row), before eventually fading away at a

larger time (bottom row). This focusing of the localized wave group is due to the nonlinear

term of the NLS equation, which counterbalances the inverse effect of the dispersive term.

More specifically, it has recently been shown (Cousins & Sapsis, 2015b,a) that there exists a

critical length scale and amplitude for the wave packet above which nonlinear effects become

predominant in such a way that it will be likely to focus, giving rise to an extreme wave.

We are now interested in taking advantage of the DO framework to investigate the nonlinear

evolution of these focusing wave groups when they are subject to small stochastic initial

41

Page 50: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 4. Dynamics of an extreme wave

t = 0

-0.4

-0.2

0

0.2

0.4

t = 2

70

-0.4

-0.2

0

0.2

0.4

x0 50 100 150 200 250 300

t = 4

00

-0.4

-0.2

0

0.2

0.4

Figure 4.1 – Focusing wave packet in a deterministic initial condition generated as a Gaussianspectrum of waves (3.3) with BFI = 1.43.

perturbations, and to analyze the resulting structure of the stochastic solution. As mentioned

in the concluding remarks of Chapter 3, we do this by assigning the wave field of interest

to the mean component of the initial DO expansion (2.12), while the modes and stochastic

coefficients carry small stochastic perturbations of a desired shape.

4.2 Adaptivity of the DO modes

We first investigate the ability of the DO modes to adaptively track the emergence of a localized

extreme wave in the mean. We consider the following random initial condition

u(x, t0;ω) = u(x, t0)+3∑

n=1Ane iθn (ω)e i∆kn x , (4.1)

where the mean u(x, t0) is assigned the initial wave field of Figure 4.1 (which contains the

focusing wave group), and the Fourier modes represent small stochastic perturbations with

wavenumbers ∆kn = 0.2n, deterministic amplitude An = 0.001 and random uncorrelated

phases θn(ω) drawn from a uniform distribution on [0,2π]. As before, the stochastic coeffi-

cients and modes of the initial DO expansion (2.12) are initiated through the relations (2.44),

resulting in 6 DO modes. The solution for the mean and the first four DO modes at vari-

ous times is plotted in Figure 4.2, where both the modulus (blue) and real part (orange) are

represented. While we have assigned global Fourier mode perturbations to the initial DO

modes, at time t = 270 when the extreme wave occurs in the mean the first two DO modes are

observed to converge towards the location of the extreme wave. This reveals that the dominant

42

Page 51: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

4.3. Attractor of an idealized extreme wave

Mea

n

-0.2

-0.1

0

0.1

0.2t = 0

Mod

e 1

-0.1

-0.05

0

0.05

0.1

Mod

e 2

-0.1

-0.05

0

0.05

0.1

Mod

e 3

-0.1

-0.05

0

0.05

0.1

x100 150 200

Mod

e 4

-0.1

-0.05

0

0.05

0.1

-0.2

-0.1

0

0.1

0.2t = 100

-0.1

0

0.1

0.2

-0.1

0

0.1

0.2

-0.1

-0.05

0

0.05

0.1

x100 150 200

-0.1

0

0.1

0.2

-0.2

0

0.2

0.4t = 270

-0.2

0

0.2

0.4

-0.4

-0.2

0

0.2

0.4

-0.1

-0.05

0

0.05

0.1

x100 150 200

-0.1

0

0.1

0.2

Figure 4.2 – Mean and first four DO modes at various times for the initial wave field of Figure4.1 (contained in the mean) subject to small stochastic perturbations (contained in the modes).Both the complex envelope modulus (blue) and real part (orange) are represented.

directions of stochastic fluctuations at that time are concentrated within the extreme wave

itself. This is not surprising considering that the focusing properties of localized wave groups

depend in a sensitive manner on their initial amplitude (Cousins & Sapsis, 2015b), meaning

that the initial stochastic perturbations will be more amplified at the location of the focusing

wave group than elsewhere in the domain. Still, these results show the nice time-adaptive

property of the DO modes, able to track the main directions of stochastic fluctuations even

these are highly time-dependent. Such a property is completely absent from order-reduction

schemes that rely on fixed basis functions such as the POD.

4.3 Attractor of an idealized extreme wave

We now study the evolution of an isolated wave packet of the idealized form u(x, t0) =A0sech(x/L0), without any background spectrum of waves. This idealized shape has been

extensively studied in the literature (Yuen & Lake, 1975; Peregrine, 1983; Dysthe & Trulsen,

1999; Cousins & Sapsis, 2015b) as a prototype model for the focusing wave groups observed

in irregular wave fields such as the one in Figure 4.1. Indeed, such a wave packet will either

focus and grow in amplitude when its initial amplitude A0 is larger than the critical value

A0,c = 1/(p

2L0), leading to an extreme wave, or broaden and decay otherwise. Here, we

consider a length scale L0 = 7 and amplitude A0 = 0.15 giving roughly similar properties to

43

Page 52: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 4. Dynamics of an extreme waveM

ean

0

0.05

0.1

0.15t = 0

Mod

e 1

-0.4

-0.2

0

0.2

0.4

Mod

e 2

-0.1

0

0.1

0.2

0.3

Mod

e 3

-0.2

-0.1

0

0.1

0.2

x100 150 200

Mod

e 4

-0.1

0

0.1

0.2

-0.4

-0.2

0

0.2

0.4t = 300

-0.2

0

0.2

0.4

0.6

-0.4

-0.2

0

0.2

0.4

-0.2

0

0.2

0.4

0.6

x100 150 200

-0.2

0

0.2

0.4

-0.05

0

0.05

0.1

0.15t = 500

-0.4

-0.2

0

0.2

0.4

-0.2

0

0.2

0.4

-0.2

0

0.2

0.4

x100 150 200

-0.4

-0.2

0

0.2

0.4

Figure 4.3 – Mean and first four DO modes at various times for a localized wave group ofthe form A0sech(x/L0) initially subject to small localized stochastic perturbations. Both thecomplex envelope modulus (blue) and real part (orange) are represented. Note that t = 300would be the time of maximum focusing for the unperturbed wave group.

those of the focusing localized wave group initially observed in Figure 4.1. A deterministic

simulation reveals that these values lead to a maximum focusing of 0.24 around t = 300.

We now study the evolution of this idealized localized wave group subject to small stochastic

perturbations. We consider the following random initial condition

u(x, t0;ω) = A0sech(x/L0)+ sech(x/L0)3∑

n=1Ane iθn (ω)e i∆kn x , (4.2)

where A0 = 0.15, L0 = 7 and the Fourier modes are mollified with the same sech function as

assigned to the mean. Hence they represent localized small stochastic perturbations, with

wavenumbers ∆kn = 0.02n, deterministic amplitude An = 0.002 and random uncorrelated

phases θn(ω) drawn from a uniform distribution on [0,2π]. The solution for the mean and the

modes at various times is plotted in Figure 4.3, where both the modulus (blue) and real part

(orange) are represented. The initial perturbation implies that every realization corresponds

to a slightly different initial shape for the localized wave group, in such a way that the DO

modes evolve in a time-dependent local optimal basis to describe the dominant directions in

the fluctuations of the realizations around the mean. In Figure 4.4, we show the energy present

in the mean together with the stochastic energy present in the modes. We observe that the

44

Page 53: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

4.3. Attractor of an idealized extreme wave

t0 50 100 150 200 250 300 350 400 450 500

E[Y

i2]

10 -4

10 -3

10 -2

10 -1

100

MeanMode 1Mode 2Mode 3Mode 4Mode 5Mode 6

Figure 4.4 – Energy of the mean ⟨u, u⟩ and the modes E [Y 2i ] for a localized wave group of the

form A0sech(x/L0) subject to small localized stochastic perturbations.

Figure 4.5 – Stochastic attractor at various times for the solution to a localized wave groupof the form A0sech(x/L0) initially subject to small localized stochastic perturbations. Theattractor is represented in terms of the 3D scatter plot of the first three stochastic coefficientsfor all realizations. In addition, each realization is assigned a color indicative of the maximumvalue of its corresponding envelope modulus |u(x, t ;ω)|.

mean is transferring energy to the modes as time evolves, showing that the initial stochastic

fluctuations of the realizations around the mean become amplified over time. In Figure 4.5,

we display the stochastic attractor of the solution, represented as the 3D scatter plot of the

first three stochastic coefficients for all realizations. It is observed that the solution falls on

an attractor of low effective dimensionality. Indeed, the attractor for the first three stochastic

coefficients appears as a locally two-dimensional structure in the three-dimensional space

of all possible realizations. We have also colored each realization according to the maximum

value of its corresponding envelope modulus, and we note that at t = 300 a clear correlation

emerges between the maximum of the envelope modulus for a given realization (i.e. the

degree of focusing of the wave group) and the associated stochastic coefficients. In Figure

4.6 we illustrate the flows of stochastic energy between modes and the mean. Specifically, we

represent the energy production in every mode due to (i) its linear interaction with the mean

(first column), (ii) its nonlinear simultaneous interaction with the mean and two other modes

(second column) and (iii) its nonlinear interaction with three other modes (third column).

45

Page 54: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 4. Dynamics of an extreme wave

Mod

e 1

#10 -4

-1

0

1

2

3"mean!i

Mod

e 2

#10 -5

-4

-2

0

2

Mod

e 3

#10 -5

-3

-2

-1

0

1

t100 200 300 400 500

Mod

e 4

#10 -6

-4

-2

0

2

4

#10 -5

-5

0

5

10

15"mean;mn!i

#10 -5

-5

0

5

#10 -5

-1

0

1

2

3

t100 200 300 400 500

#10 -6

-5

0

5

10

#10 -5

-10

-5

0

5"mnl!i

#10 -5

-5

0

5

10

#10 -5

-2

0

2

4

t100 200 300 400 500

#10 -6

-4

-2

0

2

4

Figure 4.6 – Modal energy production in the first four modes due to (i) linear interaction withthe mean flow (first column), (ii) nonlinear simultaneous interaction with the mean and twoother modes (second column) and (iii) nonlinear interaction with three other modes (thirdcolumn) for the solution to a localized wave group of the form A0sech(x/L0) initially subjectto small localized stochastic perturbations.

mem

ber

with

max

imum

|u|

-0.2

0

0.2

t = 0

x100 150 200

mem

ber

with

min

imum

|u|

-0.2

0

0.2

-0.2

0

0.2

t = 300

x100 150 200

-0.2

0

0.2

-0.2

0

0.2

t = 500

x100 150 200

-0.2

0

0.2

Figure 4.7 – Two realizations at various times of the solution to a localized wave group ofthe form A0sech(x/L0) initially subject to small localized stochastic perturbations. Both thesurface elevation (blue) and complex envelope modulus (orange) are represented. The toprow shows the realization with maximum envelope modulus |u(x, t ;ω)| at t = 300, while thebottom row shows that with minimum envelope modulus.

Energy transfers appear to be dominated by the first mode, as could be inferred from Figure 4.4.

The spikes in the energy transfers associated with DO mode 4 around t ∼ 200 are an artifact

due to its crossing with another mode (see Figure 4.4). Finally, we plot in Figure 4.7 the surface

elevation (blue) and envelope modulus (orange) at various times of the two realizations that

have the maximum (top row) and minimum (bottom row) envelope modulus at t = 300. No

big differences in the degree of focusing at t = 300 are observed between the two.

46

Page 55: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 5

Results with the MNLS equation

In this chapter, we present results from a higher-order version of the NLS equation, the

modified nonlinear Schrödinger (MNLS) equation that we briefly review in Section 5.1, along

with the corresponding DO reduced-order equations. Similarly to what was done in Section 3.1

for the NLS equation, in Section 5.2 we validate our new DO equations for the MNLS equation

through the simulation of a uniform wavetrain undergoing ‘semi-stochastic’ Benjamin-Feir

instability and Fermi-Pasta-Ulam recurrence. Finally, in Section 5.3 we investigate the behavior

under the MNLS equation of an idealized extreme wave subject to small stochastic initial

perturbations and we compare the results to the corresponding ones for the NLS equation

from Section 4.3.

5.1 Modified nonlinear Schrödinger and DO equations

The modified nonlinear Schrödinger (MNLS) equation was introduced by Dysthe (1979)

with the aim of relaxing the steepness limitation of the original NLS equation. Taking the

perturbation expansion in the wave steepness ε= k0a leading to the NLS equation one step

further to O (ε4), the MNLS equation is obtained. In addition to being more accurate for values

of the wave steepness larger than 0.15, this new equation corrects an important shortcoming of

the NLS equation in two dimensions. Indeed, the instability region of a uniform wave subject

to two-dimensional perturbations is infinite in extent (in the perturbation wavenumber plane)

for the NLS equation (Yuen & Lake, 1980; Janssen, 2004), which is a highly unphysical result.

Meanwhile, the same two-dimensional instability does have a high-wavenumber cutoff under

the MNLS equation (Trulsen & Dysthe, 1996). Finally, the MNLS equation has been shown to

compare favorably to experiments and simulations of the governing fully nonlinear equations

(Lo & Mei, 1985; Goullet & Choi, 2011), up to a time scale t ∼ 10ε−2, one order of magnitude

larger than that for the NLS equation (Trulsen et al., 2001).

The MNLS equation is presented in Appendix B, together with its corresponding DO equations.

Except for the deterministic forcing fields F0, Fi , Fi j and Fi j k appearing in the DO evolution

47

Page 56: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 5. Results with the MNLS equation

equation (2.21) and which can be found in the Appendix, the DO equations for MNLS are

exactly the same as those presented in Section 2.2 for the NLS equation. Hence the expressions

for energy transfers can be directly derived from the results of Section 2.3, and the numerical

implementation procedure described in Section 2.4 is still valid. Note however that, this time,

all equations (i.e. for the mean, the modes and the stochastic coefficients) are advanced in

time using an explicit 4th-order Runge-Kutta scheme with a nondimensional time step of

0.01, and the spatial derivatives are calculated with a spectral method. As is usually done wen

numerically solving the MNLS equation, modulation wavenumbers ∆k greater than 1 in the

mean and the modes are deleted at each time step (Dysthe et al., 2003).

5.2 Idealized Benjamin-Feir instability

In order to validate the DO equations for MNLS and their numerical implementation, we

compute the stochastic evolution of a uniform plane wave subject to small Fourier mode per-

turbations with deterministic amplitude but random phase. This situation is expected to lead

to Benjamin-Feir instability and Fermi-Pasta-Ulam recurrence, and was already considered in

Section 3.1 in the context of the NLS equation. Here, we use the same initial condition (3.1)

under the same parameter values as in Section 3.1, that is, we consider a spatially constant en-

velope of amplitude a0 = 0.1 that is perturbed with one unstable Fourier mode of modulation

wavenumber ∆k1 = 0.2 and one stable Fourier mode of wavenumber ∆k2 = 0.4. The Fourier

modes are assigned independent random phases and deterministic small amplitude equal to

0.0036, resulting in a total of 4 DO modes.

The solution for the mean and the modes at various times is plotted in Figure 5.1, where both

the modulus (blue) and real part (orange) are represented. In Figure 5.2, we show the energy

present in the mean ⟨u, u⟩ together with the stochastic energy present in the modes E [Y 2i ]. The

observations from Section 3.1 are still valid here. In a nutshell, the DO modes 1 and 2 at t = 800

in Figure 5.1 still represent the linearly unstable modulation ∆k1 = 0.2 that was assigned in

the initial condition, while modes 3 and 4 represent the stable modulation ∆k2 = 0.4, and

the spatially constant mean still contains the uniform carrier wave. Figure 5.2 shows that

the stochastic energy in the linearly unstable modes 1 and 2 grows then decays while the

contrary happens to the energy of the mean (i.e. the uniform wave), in perfect accordance

with Benjamin-Feir instability and Fermi-Pasta-Ulam recurrence.

For a quantitative comparison with deterministic theory and similarly to what we did in Section

3.1, we computed the MNLS solution to a deterministic initial condition of the form u(x,0) =0.1+0.0036cos∆km x where∆km = 0.2, i.e. similar in structure to one realization of the random

DO initial condition, and we retrieve the energy of the carrier wave and the modulations by

means of a Fourier transform of the envelope. The normalized resulting squared modulus of

the Fourier coefficients for ∆k = 0 (carrier wave), 0.2 (unstable modulation) and 0.4 (stable

harmonic) are shown as dashed lines in Figure 5.2. This time, the agreement between the

stochastic DO computation and the deterministic calculation is not as good as what we

48

Page 57: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

5.2. Idealized Benjamin-Feir instability

Mea

n

-0.1

0

0.1

t = 0M

ode

1

-0.1

-0.05

0

0.05

0.1

Mod

e 2

-0.1

-0.05

0

0.05

0.1

Mod

e 3

-0.1

-0.05

0

0.05

0.1

x0 10 20 30

Mod

e 4

-0.1

-0.05

0

0.05

0.1

-0.1

0

0.1

t = 800

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

x0 10 20 30

-0.1

-0.05

0

0.05

0.1

-0.1

0

0.1

t = 1400

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

-0.1

-0.05

0

0.05

0.1

x0 10 20 30

-0.1

-0.05

0

0.05

0.1

Figure 5.1 – Mean and DO modes at various times for the stochastic BF instability and FPUrecurrence with random phase in the initial modulation, under the MNLS equation. Both thecomplex envelope modulus (blue) and real part (orange) are represented. Note that while thedomain size is L = 50·2π, we only plot the solution over a portion of the domain correspondingto the wavelength of the linearly most unstable wavenumber ∆km = 2a0 = 0.2.

t0 200 400 600 800 1000 1200 1400 1600 1800 2000

E[Y

i2]

10 -3

10 -2

10 -1

100

101

MeanMode 1Mode 2Mode 3Mode 4Deterministic

Figure 5.2 – Energy of the mean ⟨u, u⟩ and the modes E [Y 2i ] for the stochastic BF instability and

FPU recurrence with random phase in the initial modulation, under the MNLS equation. Thedashed lines show the corresponding energies for a deterministic simulation of an equivalentinitial condition u(x,0) = 0.1+0.0036cos∆km x with ∆km = 0.2, with the energies obtained asthe normalized squared modulus of the Fourier coefficients of wavenumber ∆k = 0 (carrierwave), 0.2 (unstable modulation) and 0.4 (stable harmonic) (similarly to Figure 1.3).

obtained for the NLS equation, and the reason for this discrepancy has yet to be understood.

Note also that the collapse of the numerical solution observed in Section 3.1 for the NLS

49

Page 58: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 5. Results with the MNLS equation

equation is also happening here, as can be seen from the shape of the modes at t = 1400 which

deviate from deterministic theory, or the oscillations in the energy levels from t ∼ 1900 in

Figure 5.2.

As a conclusion, we have shown that our DO equations for MNLS produce results in accor-

dance with deterministic Benjamin-Feir and Fermi-Pasta-Ulam recurrence theory. While the

quantitative agreement between our DO simulation and deterministic results is not as excel-

lent as what was obtained for the NLS equation, we nevertheless proceed with the investigation

of extreme waves in the following section.

5.3 Attractor of an idealized extreme wave

Here we investigate the behavior under the MNLS equation of an idealized extreme wave

subject to small stochastic initial perturbations. We consider the exact same case as that

studied in Section 4.3 in the context of the NLS equation. Specifically, we study the evolution

of an idealized wave packet of the form u(x, t0) = A0sech(x/L0) with L0 = 7 and A0 = 0.15. A

deterministic simulation of the MNLS equation reveals that these values lead to a maximum

focusing amplitude of 0.21 at t = 410. The lower growth rate and maximum focusing amplitude

for MNLS as compared to NLS have both been previously reported in the literature (Dysthe,

1979; Cousins & Sapsis, 2015b).

We now study the behavior of this idealized localized wave group subject to small stochastic

initial perturbations. We consider the same random initial condition (4.2) as in Section 4.3,

we compute its evolution under the MNLS equation in the DO framework and we compare

the obtained stochastic solution with the NLS results from Section 4.3. The solution for the

mean and the modes at various times is plotted in Figure 5.3, where both the modulus (blue)

and real part (orange) are represented. While the shape of the modes is overall similar to the

NLS results of Figure 4.3, the asymmetric profile of the mean in the case of MNLS is a major

difference with NLS and reflects the observed appearance of a tail during the evolution of

certain envelope solitons (Yuen & Lake, 1975). In Figure 5.4, we show the energy present in

the mean together with the stochastic energy present in the modes, and they are very similar

to those observed for NLS in Figure 5.4. In Figure 5.5, we display the stochastic attractor of

the solution, represented as the 3D scatter plot of the first three stochastic coefficients for all

realizations. Additionally, each realization is colored according to the maximum value of its

corresponding envelope modulus. Comparing this attractor to the NLS one of Figure 5.5 shows

that both attractors are roughly similar at the time of maximum focusing (t = 300 for NLS,

t = 410 for MNLS), while differences arise between NLS and MNLS at a later time. In Figure

5.6 we illustrate the flows of stochastic energy between modes and the mean. Specifically, we

represent the energy production in every mode due to (i) its linear interaction with the mean

(first column), (ii) its nonlinear simultaneous interaction with the mean and two other modes

(second column) and (iii) its nonlinear interaction with three other modes (third column).

These energy transfer plots are qualitatively similar to the NLS results of Figure 4.6. Finally, we

50

Page 59: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

5.3. Attractor of an idealized extreme wave

Mea

n

0

0.05

0.1

0.15t = 0

Mod

e 1

-0.4

-0.2

0

0.2

0.4

Mod

e 2

-0.1

0

0.1

0.2

0.3

Mod

e 3

-0.2

-0.1

0

0.1

0.2

x100 150 200

Mod

e 4

-0.1

0

0.1

0.2

-0.2

-0.1

0

0.1

0.2t = 410

-0.2

0

0.2

0.4

-0.2

0

0.2

0.4

-0.4

-0.2

0

0.2

0.4

x100 150 200

-0.4

-0.2

0

0.2

0.4

-0.05

0

0.05

0.1

0.15t = 680

-0.4

-0.2

0

0.2

0.4

-0.4

-0.2

0

0.2

0.4

-0.2

0

0.2

0.4

x100 150 200

-0.2

0

0.2

0.4

Figure 5.3 – Mean and first four DO modes at various times for a localized wave group ofthe form A0sech(x/L0) initially subject to small localized stochastic perturbations, underthe MNLS equation. Both the complex envelope modulus (blue) and real part (orange) arerepresented. Note that t = 410 would be the time of maximum focusing for the unperturbedwave group.

t0 100 200 300 400 500 600

E[Y

i2]

10 -4

10 -3

10 -2

10 -1

100

MeanMode 1Mode 2Mode 3Mode 4Mode 5Mode 6

Figure 5.4 – Energy of the mean ⟨u, u⟩ and the modes E [Y 2i ] for a localized wave group of

the form A0sech(x/L0) subject to small localized stochastic perturbations, under the MNLSequation.

plot in Figure 5.7 the surface elevation (blue) and envelope modulus (orange) at various times

of the two realizations that have the maximum (top row) and minimum (bottom row) envelope

modulus at t = 410. The tail observed in the mean in Figure 5.3 is reflected in these individual

realizations. Note that such dail does not appear in the NLS realizations of Figure 4.7 that

maintain a regular symmetric shape. As was observed in the case of NLS, no big differences in

51

Page 60: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 5. Results with the MNLS equation

Figure 5.5 – Stochastic attractor at various times for the MNLS solution to a localized wavegroup of the form A0sech(x/L0) initially subject to small localized stochastic perturbations.The attractor is represented in terms of the 3D scatter plot of the first three stochastic coef-ficients for all realizations. In addition, each realization is assigned a color indicative of themaximum value of its corresponding envelope modulus |u(x, t ;ω)|.

Mod

e 1

#10 -4

0

0.5

1

1.5

2"mean!i

Mod

e 2

#10 -5

-2

-1

0

1

Mod

e 3

#10 -6

-15

-10

-5

0

5

t200 400 600

Mod

e 4

#10 -6

-2

-1

0

1

2

#10 -5

-5

0

5

10

15"mean;mn!i

#10 -5

-5

0

5

#10 -5

-1

0

1

2

t200 400 600

#10 -6

-4

-2

0

2

4

#10 -5

-10

-5

0

5"mnl!i

#10 -5

-5

0

5

10

#10 -6

-5

0

5

10

t200 400 600

#10 -6

-2

0

2

4

6

Figure 5.6 – Modal energy production in the first four modes due to (i) linear interaction withthe mean flow (first column), (ii) nonlinear simultaneous interaction with the mean and twoother modes (second column) and (iii) nonlinear interaction with three other modes (thirdcolumn) for the MNLS solution to a localized wave group of the form A0sech(x/L0) initiallysubject to small localized stochastic perturbations.

the degree of focusing at t = 410 are observed between the two realizations since the initial

perturbations are small. However, here a difference in shape is observed between the two

realizations, the top one having a more pronounced tail at t = 410 than the bottom one, a

situation that reverses at t = 680.

52

Page 61: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

5.3. Attractor of an idealized extreme wave

mem

ber

with

max

imum

|u|

-0.2

0

0.2

t = 0

x100 150 200

mem

ber

with

min

imum

|u|

-0.2

0

0.2

-0.2

0

0.2

t = 410

x100 150 200

-0.2

0

0.2

-0.2

0

0.2

t = 680

x100 150 200

-0.2

0

0.2

Figure 5.7 – Two realizations at various times of the MNLS solution to a localized wave groupof the form A0sech(x/L0) initially subject to small localized stochastic perturbations. Both thesurface elevation (blue) and complex envelope modulus (orange) are represented. The toprow shows the realization with maximum envelope modulus |u(x, t ;ω)| at t = 410, while thebottom row shows that with minimum envelope modulus.

53

Page 62: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework
Page 63: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Conclusions and perspectives

In this thesis, we have implemented a reduced-order stochastic framework based on the

dynamically orthogonal (DO) equations (introduced in Sapsis & Lermusiaux, 2009), for the

stochastic evolution of water waves governed by the nonlinear Schrödinger (NLS) equation

and subject to random initial conditions. Using a generalized time-dependent truncated

Karhunen-Loève expansion, we decomposed the stochastic solution in a mean state and

stochastic fluctuations, the latter being described by a finite number of deterministic modes

and associated stochastic coefficients. A set of explicit and coupled equations was then derived

for the time evolution of these quantities, allowing for the efficient computation of the full

stochastic solution, since only the stochastic coefficients are solved in a Monte-Carlo fashion.

Additionally, expressions were derived to quantify the transfers of energy (in a variance sense)

between modes and the mean state.

We benchmarked our DO equations against two well-known cases, at the same time illustrating

their use. First, we considered a uniform wavetrain perturbed by unstable modulations with

small deterministic amplitude but random phases. In perfect accordance with deterministic

results, the stochastic solution was observed to undergo Benjamin-Feir instability followed by

Fermi-Pasta-Ulam recurrence. The DO modes reproduced exactly the shape of the linearly

unstable modulation and its harmonics, while the variance of the stochastic coefficients was

observed to increase and decrease at the exact same rate as the unstable modulation of an

equivalent deterministic simulation. Energy transfers showed the flow of energy from the

mean (containing the uniform wave) to the modes then back to the mean, in accordance

with Fermi-Pasta-Ulam recurrence. The stochastic solution was observed to possess a low

dimensional attractor in phase space.

Second, we considered the case of a Gaussian spectrum of waves with random phases, leading

to an irregular wave field for each realization. Relevant statistical properties of the DO solution

were compared with ensemble-averaged results from full Monte-Carlo simulations and a

reasonable agreement between the two was obtained. However, the need to reproduce a large

number of wavenumbers with independent random phases resulted in a high number of DO

modes and a high computational cost. Moreover, the modes were observed to all converge to

similar levels of energy, meaning that no dominant direction in the stochastic dynamics of

these irregular wave fields could be perceived on a global scale. However, we observed in some

55

Page 64: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Chapter 5. Results with the MNLS equation

individual realizations of the higher-energy wave fields that concentration of energy in the

form of localized wave packets resulted in local extreme waves, phenomenon that is known in

literature (Ruban, 2013). Due to their appearance at random locations in the domain and their

presence in only some of the realizations, these extreme waves could not be captured as such

by the DO modes. This, combined with the issue of computational cost, indicated that the DO

framework was not suited for the solution of a stochastic wave field with Gaussian spectrum

and random phases.

Instead, we turned our attention to a single one of these extreme waves, this time using

the DO framework as a means to investigate its nonlinear evolution under small stochastic

perturbations of its initial shape. We first showed that in this case the DO modes are able to

adaptively track the emergence of the extreme wave out of a Gaussian spectrum background

of waves. We then focused on an isolated wave packet of idealized shape as a prototype model

for these extreme waves and investigated its nonlinear evolution when initially subjected to

small localized stochastic perturbations. The resulting stochastic solution was observed to fall

on a low-dimensional attractor in phase space, and the modes revealed the effect of initial

perturbations on the nonlinear evolution of the wave group.

As an immediate next step, we want to consider a higher-order version of the NLS equation,

the modified nonlinear Schrödinger (MNLS) equation introduced by Dysthe (1979). Since it is

accurate to a higher order in amplitude, its use would be more appropriate for the investigation

of the nonlinear evolution of extreme waves. The DO equations for the MNLS equation can be

found in Appendix B and initial results for the case of Section 3.1 have already been obtained,

validating the equations and their implementation. Another possible future direction concerns

the use of a complex-valued inner product hence complex-valued stochastic coefficients in

the DO expansion (2.13). As has been discussed in Section 2.2.2, this choice would lead

to a reduction in the number of required modes, but dealing with the resulting solution is

not straightforward as was seen in Appendix A. A third possible line of work would be the

integration of real-life measurements with our reduced-order model, following a Kalman

filtering algorithm. Finally, recall that a problem of the DO modes resides in their inability to

track localized events that appear at different random locations between the realizations. A

possible way around this problem could be to use the following expansion instead of (2.13)

u(x, t ;ω) = u(x + c(t ;ω), t ;ω) = ¯u(x + c(t ;ω), t )+s∑

i=1Yi (t ;ω)ui (x + c(t ;ω), t ), (5.1)

where the time-dependent stochastic shift c(t ;ω) would allow the modes to track dynamics

appearing at different locations between the realizations. Additional equations are then

needed for c(t ;ω) in order to close the problem. Similar issues have been investigated in the

context of the POD (Rowley & Marsden, 2000), but these concerned statistically stationary

systems so our problem presents a greater challenge.

56

Page 65: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Appendix A

Complex coefficients in the DOframework

Here, we consider the complex-valued inner product (2.5) and we derive the resulting DO

equations for the NLS equation. As was thoroughly discussed in Section 2.2.2, this complex

inner product implies that the stochastic coefficients in the DO expansion (2.13) are also

complex-valued, hence we hereafter refer to them as Zi (t ;ω). First, we derive the DO equations

in Section A.1. We then show issues associated with the removal of the correlations between

the complex coefficients in Section A.2.

A.1 Dynamically orthogonal equations with complex coefficients

As was done in Section 2.2.3, we begin by inserting the DO expansion (2.13) with complex

stochastic coefficients Zi (t ;ω) in the NLS equation (2.19), leading to the following governing

equation for all unknown quantities

∂u

∂t= F0 +Zi Fi +Z∗

i Gi +Zi Z j Fi j +Zi Z∗j Gi j +Zi Z j Z∗

k Fi j k , (A.1)

where the asterisk denotes the complex conjugate, and the deterministic fields on the right-

hand side are defined as

F0 =− i

8

∂2u

∂x2 − i

2|u|2u, Fi =− i

8

∂2ui

∂x2 − i |u|2ui , Gi =− i

2u2u∗

i ,

Fi j =− i

2u∗ui u j , Gi j =−i uui u∗

j , Fi j k =− i

2ui u j u∗

k .

(A.2)

Following the procedure of Section 2.2.3, the equation for the mean field writes

∂u

∂t= F0 +CZi Z j Fi j +CZi Z∗

jGi j +MZi Z j Z∗

kFi j k , (A.3)

57

Page 66: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Appendix A. Complex coefficients in the DO framework

where CZi Z j = E [Zi Z j ] and CZi Z∗j= E [Zi Z∗

j ] are respectively the complex pseudo-covariance

and covariance matrices, and MZi Z j Z∗k= E [Zi Z j Z∗

k ] is the matrix of third-order moments. The

stochastic coefficients obey the following evolution equation

dZi

dt= Zm ⟨Fm ,ui ⟩+Z∗

m ⟨Gm ,ui ⟩+ (Zm Zn −CZm Zn )⟨Fmn ,ui ⟩+ (Zm Z∗

n −CZm Z∗n

)⟨Gmn ,ui ⟩+ (Zm Zn Z∗l −MZm Zn Z∗

l)⟨Fmnl ,ui ⟩.

(A.4)

Finally, the evolution of the basis functions is given by

∂ui

∂t= Hi −⟨Hi ,u j ⟩u j , (A.5)

where Hi is defined by E [L [u]Zk ]C−1Zk Z∗

i= Hi and has the following expression

Hi = Fi +Gm CZ∗m Z∗

kC−1

Zk Z∗i+Fmn MZm Zn Z∗

kC−1

Zk Z∗i

+Gmn MZm Z∗n Z∗

kC−1

Zk Z∗i+Fmnl MZm Zn Z∗

l Z∗k

C−1Zk Z∗

i,

(A.6)

with MZm Zn Z∗l Z∗

k= E [Zm Zn Z∗

l Z∗k ] the matrix of fourth-order moments. While we were able to

derive these equations in a straightforward manner, we will see some of the complications

introduced by complex coefficients in the next section.

A.2 Diagonalization of the complex covariance matrix

Consider the DO solution at a given time. Similarly to what is done in Section 2.4.3 for real-

valued coefficients, we want to remove the second-order correlations between the complex-

valued stochastic coefficients. The second-order statistics are described by a complex covari-

ance matrix Ci j = E [Zi Z∗j ] and a complex pseudo-covariance matrix Ci j = E [Zi Z j ], and both

matrices may have non-zero off-diagonal elements. The covariance matrix is Hermitian and

can therefore be decomposed as

C =V DV † (A.7)

where the columns of V contain the eigenvectors of C and V satisfies V V † =V †V = I with †

denoting the complex conjugate, and D is a diagonal matrix containing the (real) eigenvalues

of C . We now define a new basis with u′i = umVmi . Note that

⟨ui ,u′j ⟩ = ⟨ui ,umVm j ⟩ =V ∗

m j ⟨ui ,um⟩ =V ∗i j ⇒ Vi j = ⟨u′

j ,ui ⟩. (A.8)

We first show that the new basis u′i is orthonormal

⟨u′i ,u′

j ⟩ = ⟨umVmi ,unVn j ⟩ =Vmi V ∗n j ⟨um ,un⟩ =Vmi V ∗

m j =V †j mVmi = δi j . (A.9)

58

Page 67: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

A.2. Diagonalization of the complex covariance matrix

The coefficients Z ′i in the new basis are given by

Z ′i = ⟨Z j u j ,u′

i ⟩ = Z j ⟨u j ,u′i ⟩ = Z j V ∗

j i , (A.10)

and they are uncorrelated with respect to the new covariance matrix

C ′i j = E [Z ′

i Z ′∗j ] = E [ZmV ∗

mi Z∗n Vn j ] =V †

i mCmnVn j = Di j = δi jλi , (A.11)

where λi are the eigenvalues of C and the diagonal elements of D . Note that, however, the new

pseudo-covariance matrix remains correlated. Indeed, we have

C ′i j = E [Z ′

i Z ′j ] = E [ZmV ∗

mi ZnV ∗n j ] =V †

i mCmnV ∗n j (A.12)

which is not necessarily diagonal. To better understand the implications of this, we consider

what we would have if we were to represent the same solution with twice the number of

real-valued coefficients. Following the discussion in Section 2.2.2, we write Zi = Xi + i Yi and

Z ′i = X ′

i + i Y ′i , where Xi , Yi , X ′

i , Y ′i would be the equivalent real-valued stochastic coefficients.

If we were to directly work with these real-valued coefficients, there would be one single

real-valued covariance matrix four times the size of the equivalent complex covariance or

pseudo-covariance matrices (since there would be twice the number of coefficients), but we

would be able to diagonalize it, resulting in completely uncorrelated coefficients i.e. E [X ′i X ′

j ] =E [Y ′

i Y ′j ] = E [X ′

i Y ′j ] = E [Y ′

i X ′j ] = 0 for i 6= j . However, in the case of the complex formulation

for which we only managed to diagonalize the complex covariance matrix C ′i j , we have

C ′i j = E [X ′

i X ′j ]+E [Y ′

i Y ′j ]− i E [X ′

i Y ′j ]+ i E [Y ′

i X ′j ] = δi jλi , (A.13)

C ′i j = E [X ′

i X ′j ]−E [Y ′

i Y ′j ]+ i E [X ′

i Y ′j ]+ i E [Y ′

i X ′j ] 6= δi jλi , (A.14)

from which we see that in order to eliminate the correlation between all the X ′i and Y ′

i , we

would need both the covariance and pseudo-covariance matrices C ′ and C ′ to be diagonal.

Since this cannot be achieved with our diagonalization procedure, there will still be correla-

tion between the real and imaginary parts X ′i and Y ′

i of the complex coefficients, while that

correlation could have been completely removed in the equivalent formulation involving only

real-valued coefficients and one single real-valued covariance matrix.

As a side note, remark that in the original KL expansion (2.7) with complex coefficients, the

coefficients only verify E [Zi Z∗j ] = 0 for i 6= j , meaning that only their covariance matrix is

diagonalized, and not their pseudo-covariance matrix. This therefore corresponds to what we

are able to do, but doesn’t explain the difference with the equivalent real-valued formulation

for which we are moreover able to remove the correlations between the real and imaginary

parts of the equivalent complex coefficients. In any case, much more understanding of the

subject must be done before we can feel confident to proceed with using the complex-valued

inner product and coefficients. This issue of diagonalization of complex statistics is indeed an

area of active research (Eriksson & Koivunen, 2006; Adali et al., 2011; Cheong Took et al., 2012).

59

Page 68: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework
Page 69: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Appendix B

Dynamically orthogonal MNLSequation

The modified nonlinear Schrödinger equation (MNLS) was obtained by Dysthe (1979) through

a similar multiple-scales expansion (1.12) as for the NLS equation, but taking the perturbation

expansion one step further to fourth order in wave steepness ε= k0a. In a reference frame

moving at the group velocity of the carrier wave, the resulting equation for the complex

envelope of the first harmonic writes

∂u

∂t=− i

8

∂2u

∂x2 + 1

16

∂3u

∂x3 − i

2|u|2u − 3

2|u|2 ∂u

∂x− 1

4u2 ∂u∗

∂x− i u

∂φ

∂x

∣∣∣∣z=0

(B.1)

where time has been nondimensionalized withω0 and space with k0, the asterisk ∗ denotes the

complex conjugate and φ is the zeroth harmonic velocity potential, which can be expressed

in terms of u by solving Laplace’s equation with appropriate boundary conditions (given in

Dysthe, 1979; Trulsen & Dysthe, 1996) to give

∂φ

∂x

∣∣∣∣z=0

=−1

2F−1|k|F |u|2, (B.2)

where F denotes the Fourier transform. The corresponding reduced-order DO equations are

derived in a similar fashion as those for the NLS equation. Using the real-valued inner product

(2.17), the equations for the quantities involved in the DO expansion (2.13) are exactly the

same, except for the deterministic fields F0, Fi , Fi j and Fi j k which become in this case

F0 =− i

8

∂2u

∂x2 + 1

16

∂3u

∂x3 − i

2|u|2u − 3

2|u|2 ∂u

∂x− 1

4u2 ∂u∗

∂x+ i

2F−1|k|F |u|2u, (B.3)

Fi =− i

8

∂2ui

∂x2 + 1

16

∂3ui

∂x3 − i

2|u|2ui − i u Reuu∗

i − 3

2|u|2 ∂ui

∂x−3

∂u

∂xReuu∗

i

− 1

4u2 ∂u∗

i

∂x− 1

2u∂u∗

∂xui + iF−1|k|F Reuu∗

i u + i

2F−1|k|F |u|2ui , (B.4)

61

Page 70: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Appendix B. Dynamically orthogonal MNLS equation

Fi j =− i

2u Reui u∗

j − i Reuu∗i u j − 3

2

∂u

∂xReui u∗

j −3Reuu∗i ∂u j

∂x− 1

4

∂u∗

∂xui u j

− 1

2uui

∂u∗j

∂x+ i

2F−1|k|F Reui u∗

j u + iF−1|k|F Reuu∗i u j , (B.5)

Fi j k =− i

2Reui u∗

j uk −3

2Reui u∗

j ∂uk

∂x− 1

4ui u j

∂u∗k

∂x+ i

2F−1|k|F Reui u∗

j uk , (B.6)

where i , j ,k = 1, ..., s.

62

Page 71: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Bibliography

ABLOWITZ, M.J. & HERBST, B.M. 1990 On homoclinic structure and numerically induced

chaos for the nonlinear Schrödinger equation. SIAM Journal on Applied Mathematics 50 (2),

339–351.

ADALI, T., SCHREIER, P.J. & SCHARF, L.L. 2011 Complex-valued signal processing: The proper

way to deal with impropriety. IEEE Transactions on Signal Processing 59 (11), 5101–5125.

ADCOCK, T.A.A. & TAYLOR, P.H. 2009 Focusing of unidirectional wave groups on deep water:

an approximate nonlinear Schrödinger equation-based model. Proceedings of the Royal

Society of London A: Mathematical, Physical and Engineering Sciences 465 (2110), 3083–3102.

ALBER, I.E. 1978 The effects of randomness on the stability of two-dimensional surface

wavetrains. Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences

363 (1715), 525–546.

AUBRY, N., HOLMES, P., LUMLEY, J.L. & STONE, E. 1988 The dynamics of coherent structures

in the wall region of a turbulent boundary layer. Journal of Fluid Mechanics 192, 115–173.

BENJAMIN, T.B. & FEIR, J.E. 1967 The disintegration of wave trains on deep water. Part 1.

Theory. Journal of Fluid Mechanics 27 (03), 417–430.

BENNEY, D.J. & NEWELL, A.C. 1967 Propagation of nonlinear wave envelopes. Journal of

Mathematics and Physics 46 (2), 133.

BERKOOZ, G., HOLMES, P. & LUMLEY, J.L. 1993 The proper orthogonal decomposition in the

analysis of turbulent flows. Annual Review of Fluid Mechanics 25 (1), 539–575.

CHEONG TOOK, C., DOUGLAS, S.C. & MANDIC, D.P. 2012 On approximate diagonalization

of correlation matrices in widely linear signal processing. IEEE Transactions on Signal

Processing 60 (3), 1469–1473.

CHOI, M., SAPSIS, T.P. & KARNIADAKIS, G.E. 2013 A convergence study for spdes using com-

bined Polynomial Chaos and Dynamically-Orthogonal schemes. Journal of Computational

Physics 245, 281–301.

COUSINS, W. & SAPSIS, T.P. 2015a Reduced order prediction of rare events in unidirectional

nonlinear water waves. Submitted.

63

Page 72: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Bibliography

COUSINS, W. & SAPSIS, T.P. 2015b The unsteady evolution of localized unidirectional deep

water wave groups. Submitted.

DAVEY, A. 1972 The propagation of a weak nonlinear wave. Journal of Fluid Mechanics 53 (04),

769–781.

DYSTHE, K.B. 1979 Note on a modification to the nonlinear Schrödinger equation for applica-

tion to deep water waves. Proceedings of the Royal Society of London. A. Mathematical and

Physical Sciences 369 (1736), 105–114.

DYSTHE, K., KROGSTAD, H.E. & MÜLLER, P. 2008 Oceanic rogue waves. Annual Review of Fluid

Mechanics 40, 287–310.

DYSTHE, K.B. & TRULSEN, K. 1999 Note on breather type solutions of the NLS as models for

freak-waves. Physica Scripta 1999 (T82), 48.

DYSTHE, K.B., TRULSEN, K., KROGSTAD, H.E. & SOCQUET-JUGLARD, H. 2003 Evolution of a

narrow-band spectrum of random surface gravity waves. Journal of Fluid Mechanics 478,

1–10.

ERIKSSON, J. & KOIVUNEN, V. 2006 Complex random vectors and ICA models: Identifiability,

uniqueness, and separability. IEEE Transactions on Information Theory 52 (3), 1017–1029.

FERMI, E., PASTA, J. & ULAM, S. 1955 Studies of nonlinear problems. Los Alamos Report

LA-1940 .

GOULLET, A. & CHOI, W. 2011 A numerical and experimental study on the nonlinear evolution

of long-crested irregular waves. Physics of Fluids (1994-present) 23 (1), 016601.

HASIMOTO, H. & ONO, H. 1972 Nonlinear modulation of gravity waves. Journal of the Physical

Society of Japan 33 (3), 805–811.

HERBST, B.M. & ABLOWITZ, M.J. 1989 Numerically induced chaos in the nonlinear

Schrödinger equation. Physical Review Letters 62 (18), 2065.

HOLMES, P., LUMLEY, J.L. & BERKOOZ, G. 1996 Turbulence, coherent structures, dynamical

systems and symmetry. Cambridge University Press.

JANSSEN, P.A.E.M. 2003 Nonlinear four-wave interactions and freak waves. Journal of Physical

Oceanography 33 (4), 863–884.

JANSSEN, P.A.E.M. 2004 The interaction of ocean waves and wind. Cambridge University Press.

KIRBY, M. & ARMBRUSTER, D. 1992 Reconstructing phase space from PDE simulations.

Zeitschrift für Angewandte Mathematik und Physik 43 (6), 999–1022.

LAKE, B.M., YUEN, H.C., RUNGALDIER, H. & FERGUSON, W.E. 1977 Nonlinear deep-water

waves: theory and experiment. Part 2. Evolution of a continuous wave train. Journal of Fluid

Mechanics 83 (01), 49–74.

64

Page 73: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Bibliography

LIGHTHILL, M.J. 1965 Contributions to the theory of waves in non-linear dispersive systems.

IMA Journal of Applied Mathematics 1 (3), 269–306.

LIU, P.C. 2007 A chronology of freaque wave encounters. Geofizika 24 (1), 57–70.

LO, E. & MEI, C.C. 1985 A numerical study of water-wave modulation based on a higher-order

nonlinear Schrödinger equation. Journal of Fluid Mechanics 150, 395–416.

LOÈVE, M. 1945 Fonctions aléatoires de second ordre. Comptes Rendus de l’Académie des

Sciences 220.

MANTIC-LUGO, V., ARRATIA, C. & GALLAIRE, F. 2014 Self-consistent mean flow description of

the nonlinear saturation of the vortex shedding in the cylinder wake. Physical Review Letters

113 (8), 084501.

ONORATO, M., OSBORNE, A.R., SERIO, M. & BERTONE, S. 2001 Freak waves in random oceanic

sea states. Physical Review Letters 86 (25), 5831.

ONORATO, M., OSBORNE, A.R., SERIO, M. & CAVALERI, L. 2005 Modulational instability and

non-gaussian statistics in experimental random water-wave trains. Physics of Fluids (1994-

present) 17 (7), 078101.

PEREGRINE, D.H. 1983 Water waves, nonlinear Schrödinger equations and their solutions.

The Journal of the Australian Mathematical Society. Series B. Applied Mathematics 25 (01),

16–43.

ROWLEY, C.W. & MARSDEN, J.E. 2000 Reconstruction equations and the Karhunen–Loève

expansion for systems with symmetry. Physica D 142 (1), 1–19.

RUBAN, V.P. 2013 Rogue waves at low Benjamin-Feir indices: Numerical study of the role of

nonlinearity. JETP Letters 97 (12), 686–689.

SAPSIS, T.P. 2013 Attractor local dimensionality, nonlinear energy transfers and finite-time

instabilities in unstable dynamical systems with applications to two-dimensional fluid flows.

Proceedings of the Royal Society A 469 (2153), 20120550.

SAPSIS, T.P. & LERMUSIAUX, P.F.J. 2009 Dynamically orthogonal field equations for continuous

stochastic dynamical systems. Physica D 238 (23), 2347–2360.

SAPSIS, T.P. & MAJDA, A.J. 2013 Blended reduced subspace algorithms for uncertainty quan-

tification of quadratic systems with a stable mean state. Physica D 258, 61–76.

SAPSIS, T.P., UECKERMANN, M.P. & LERMUSIAUX, P.F.J. 2013 Global analysis of Navier–Stokes

and Boussinesq stochastic flows using dynamical orthogonality. Journal of Fluid Mechanics

734, 83–113.

SOBCZYK, K. 1991 Stochastic differential equations, with applications to physics and engineer-

ing. Kluwer Academic Publishers.

65

Page 74: DYNAMICALLY ORTHOGONAL REDUCED-ORDER MODELING …sandlab.mit.edu/Papers/Student_Theses/2015_Mowlavi_MS Thesis.pdf · Abstract In this thesis, we implement a reduced-order framework

Bibliography

STOKES, G.G. 1847 On the theory of oscillatory waves. Transactions of the Cambridge Philo-

sophical Society 8, 441–473.

TRULSEN, K. & DYSTHE, K.B. 1996 A modified nonlinear Schrödinger equation for broader

bandwidth gravity waves on deep water. Wave motion 24 (3), 281–289.

TRULSEN, K., STANSBERG, C.T. et al. 2001 Spatial evolution of water surface waves: Numerical

simulation and experiment of bichromatic waves. In Proceedings of the Eleventh Interna-

tional Offshore and Polar Engineering Conference, , vol. 3, pp. 71–77.

WHITHAM, G.B. 1965 A general approach to linear and non-linear dispersive waves using a

Lagrangian. Journal of Fluid Mechanics 22 (02), 273–283.

YUEN, H.C. & FERGUSON, W.E. 1978 Relationship between Benjamin–Feir instability and

recurrence in the nonlinear Schrödinger equation. Physics of Fluids (1958-1988) 21 (8),

1275–1278.

YUEN, H.C. & LAKE, B.M. 1975 Nonlinear deep water waves: Theory and experiment. Physics

of Fluids (1958-1988) 18 (8), 956–960.

YUEN, H.C. & LAKE, B.M. 1980 Instabilities of waves on deep water. Annual Review of Fluid

Mechanics 12 (1), 303–334.

YUEN, H.C. & LAKE, B.M. 1982 Nonlinear dynamics of deep-water gravity waves. Advances in

Applied Mechanics 22 (67), 229.

ZAKHAROV, V.E. 1968 Stability of periodic waves of finite amplitude on the surface of a deep

fluid. Journal of Applied Mechanics and Technical Physics 9 (2), 190–194.

ZAKHAROV, V.E. & SHABAT, A.B. 1972 Exact theory of two-dimensional self-focusing and

one-dimensional self-modulation of waves in nonlinear media. Soviet Physics JETP 34 (1),

62.

66


Recommended