Thanks to visit codestin.com
Credit goes to www.scribd.com

0% found this document useful (0 votes)
27 views58 pages

Dynamic Active Subspaces AData Driven Approach To Computing Tim

Uploaded by

yangdw1998
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
27 views58 pages

Dynamic Active Subspaces AData Driven Approach To Computing Tim

Uploaded by

yangdw1998
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 58

DYNAMIC ACTIVE SUBSPACES: A DATA-DRIVEN

APPROACH TO COMPUTING TIME-DEPENDENT


ACTIVE SUBSPACES IN DYNAMICAL SYSTEMS

by

IZABEL PIRIMAI AGUIAR

B.S., Colorado School of Mines, 2017

A thesis submitted to the

Faculty of the Graduate School of the

University of Colorado in partial fulfillment

of the requirements for the degree of

Master of Science

Department of Computer Science

2018
This thesis entitled:
DYNAMIC ACTIVE SUBSPACES: A DATA-DRIVEN APPROACH TO COMPUTING
TIME-DEPENDENT ACTIVE SUBSPACES IN DYNAMICAL SYSTEMS
written by Izabel Pirimai Aguiar
has been approved for the Department of Computer Science

Dr. Paul Constantine

Dr. Elizabeth Bradley

Dr. Jim Curry

Date

The final copy of this thesis has been examined by the signatories, and we find that both
the content and the form meet acceptable presentation standards of scholarly work in the
above mentioned discipline.
iii

Aguiar, Izabel Pirimai (M.S. Computer Science)

Dynamic active subspaces: a data-driven approach to computing time-dependent active

subspaces in dynamical systems

Thesis directed by Dr. Paul Constantine

Computational models are aiding in the advancement of science – from biological, to

engineering, to social systems. To trust the predictions of computational models, however,

we must understand how the errors in the models’ inputs (i.e., through measurement error)

affect the output of the systems: we must quantify the uncertainty that results from these

input errors. Uncertainty quantification (UQ) becomes computationally complex when there

are many parameters in the model. In such cases it is useful to reduce the dimension of the

problem by identifying unimportant parameters and disregarding them for UQ studies. This

makes an otherwise intractable UQ problem tractable. Active subspaces extend this idea

by identifying important linear combinations of parameters, enabling more powerful and

effective dimension reduction. Although active subspaces give model insight and computa-

tional tractability for scalar-valued functions, it is not enough. This analysis does not extend

to time-dependent systems. In this thesis we discuss time-dependent, dynamic active sub-

spaces. We develop a methodology by which to compute and approximate dynamic active

subspaces, and introduce the analytical form of dynamic active subspaces for two cases. To

highlight these methods we find dynamic active subspaces for a linear harmonic oscillator

and a nonlinear enzyme kinetics system.


Dedication

For Mum and Dad.


v

Acknowledgements

I would primarily like to thank my advisor, Paul Constantine. Paul has shown me,

through the opportunities and support he’s given me, how much he believes in my ability to

succeed. In doing so, he has given me some of the most positive and productive experiences

I’ve ever had. He has continually emboldened me to continue on my own path. Professionally,

no one else has impacted and influenced me as much as he has. I am extremely grateful for

the advocacy he has given me.

I would also like to sincerely thank the members of my research group: Zachary Grey,

Andrew Glaws, and Jeffrey Hokanson, for their engaging discussions, cups of coffee, criti-

cal feedback, and overwhelming kindness. Many thanks to my thesis committee, Elizabeth

Bradley, Gianluca Iaccarino, and Jim Curry for their support as well as to Nathan Kutz,

Steven Brunton, John Butcher, Kyle Niemeyer, and Peter Schmid, for their interest and en-

couragement. I am also extremely grateful for the dozens of kind and enthusiastic conference

attendees who have inspired new questions and driven further research. I am excited and

grateful to be a part of such a community.

Certainly not least, I am thankful to my family: Aristotle Johns, Jojo Clark, Sharon

Aguiar, Laura Aguiar, Matt Aguilar, Christina Whippen, Amanda Evans, Julie Clark, and

Xenia Johns. I would be nowhere without their unparalleled support, and I would know

nothing without their boundless Love.


vi

CONTENTS

CHAPTER

1 1

1.1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 EXISTING TOOLS AND RELATED WORK . . . . . . . . . . . . . . . . . 3

1.3 NOTATION AND VOCABULARY . . . . . . . . . . . . . . . . . . . . . . . 4

1.4 BACKGROUND . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.4.1 ACTIVE SUBSPACES . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.4.2 GRONWALL GRADIENTS . . . . . . . . . . . . . . . . . . . . . . . 11

1.5 DATA-DRIVEN RECOVERY OF DYNAMICAL SYSTEMS . . . . . . . . . 12

1.5.1 DYNAMIC MODE DECOMPOSITION . . . . . . . . . . . . . . . . 13

1.5.2 SPARSE IDENTIFICATION OF NONLINEAR DYNAMICAL SYS-

TEMS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2 21

2.1 DYNAMIC ACTIVE SUBSPACES . . . . . . . . . . . . . . . . . . . . . . . 21

2.1.1 The problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.1.2 The proposed solution . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.2 ANALYTIC DYNAMIC ACTIVE SUBSPACES . . . . . . . . . . . . . . . . 23

2.2.1 DyAS for an homogeneous linear dynamical system . . . . . . . . . . 23

2.2.2 DyAS for an inhomogeneous linear dynamical system . . . . . . . . . 24


vii

2.3 APPROXIMATING DYNAMIC ACTIVE SUBSPACES . . . . . . . . . . . 27

2.3.1 Introduction to the methodology . . . . . . . . . . . . . . . . . . . . 27

2.3.2 Example: Linear Harmonic Oscillator . . . . . . . . . . . . . . . . . . 29

2.3.3 Enzyme Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

2.4 DISCUSSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

2.5 CONCLUSIONS AND FUTURE WORK . . . . . . . . . . . . . . . . . . . . 43

BIBLIOGRAPHY 45
viii

TABLES

Table

2.1 The initial conditions p in (2.26) have a uniform joint density with lower and

upper bounds 20% below and above nominal values. . . . . . . . . . . . . . . 31

2.2 The parameters p have uniform joint density with upper and lower bounds

50% above and below the nominal values. . . . . . . . . . . . . . . . . . . . . 35


ix

FIGURES

Figure

2.1 The trajectory of the x, y, and z components of the linear harmonic oscillator

given by (2.24) with nominal initial conditions. . . . . . . . . . . . . . . . . . 30

2.2 The eigenvalues (left) and first eigenvector (middle) of C at time t = 5. Note

that there is only one nonzero eigenvalue. The shadow plot (right) at this time

shows the value of the linear harmonic oscillator y coordinate as a function of

a linear combination of possible initial conditions, p. . . . . . . . . . . . . . 31

2.3 (Left) the first eigenvector of C computed for the function (2.26) (Right) the
p
first eigenvector after being multiplied by λ0 (t). . . . . . . . . . . . . . . . 32

2.4 The maximum absolute error of the dynamic active subspace approximated

by (left) DMD and (right) SINDy. The data matrices for each algorithm have

been created with N samples from p proportion of the total time. . . . . . . 34

2.5 The convergence of the approximation to the integral in (1.1) for the number

of nodes per dimension in the tensor product Gauss-Legendre quadrature

scheme. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

2.6 (Left) the average absolute error over time between Gronwall Gradients and

the second-order finite difference approximation for the gradient of f with

respect to parameters. (Right) the absolute error over time between Gronwall

Gradients and the second-order finite difference approximation for the gradient

of f with respect to parameters for h = 10−3 . . . . . . . . . . . . . . . . . . 38


x

2.7 The wallclock time in computing M second-order finite difference approxi-

mations to the gradient versus integrating the augmented system in (2.35),

averaged over 15 runs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

2.8 The influence of varying α on the (left) cost and (right) absolute error of the

approximation of the dynamic active subspace with SINDy. . . . . . . . . . . 39

2.9 The maximum absolute error of the dynamic active subspace approximated

by (left) DMD and (right) SINDy. The data matrices for each algorithm have

been created with N samples from p proportion of the total time. . . . . . . 40

2.10 The (left) approximated dynamic active subspace and its (right) absolute error

using (top) DMD and (bottom) SINDy. Solutions have been approximated

with 10 samples from the first 50% of time. . . . . . . . . . . . . . . . . . . . 41


CHAPTER 1

1.1 INTRODUCTION

Computational models are aiding in the advancement of science – from biological, to

engineering, to social systems. To trust the predictions of computational models, however,

we must understand how the errors in the models’ inputs (i.e., through measurement error)

affect the output of the systems: we must quantify the uncertainty that results from these

input errors. Uncertainty quantification (UQ) becomes computationally expensive when

there are many parameters in the model. In such cases it is useful to reduce the dimension

of the model by identifying unimportant parameters and disregarding them for UQ studies.

This makes an otherwise intractable UQ problem tractable. For time-dependent systems,

these analyses are limited to local metrics that do not fully explore the parameter space,

or global metrics that are either limited to a static point in time or to the sensitivity of

individual parameters.

Active subspaces [11] identify linear combinations of the parameters that change the

quantity of interest the most on average. The active subspace of a system can be exploited

to identify coordinate-based global sensitivity metrics called activity scores [13] that provide

comparable sensitivity analysis to existing metrics. The existing approach for extending

active subspaces to time-dependent systems involves recomputing the active subspace at each

time step of interest [33, 14]. This analysis requires sampling across the parameter space,

approximating gradients, approximating an expectation, and eigendecomposing a matrix at


2

each time step. For complex time-dependent systems or for those with many parameters

this can become computationally expensive.

Extending active subspaces to time-dependent systems will enable uncertainty quantifi-

cation, sensitivity analysis, and parameter estimation for computational models that have

explicit dependence on time. Understanding how linear combinations of the parameters

change the quantity of interest over time could inform important discipline-specific insights.

Such analysis would identify which linear combinations change the quantity of interest the

most on average, and how those linear combinations change in time.

The contributions of this work. In this thesis we present a novel approach for

uncertainty quantification and sensitivity analysis in time-dependent systems. We prove the

existence of a dynamical system describing the dynamic active subspace in the cases of three

linear dynamical systems.

For other time-dependent quantities of interest, we present and develop a methodol-

ogy for approximating the dynamic active subspace with “data” in the form of the one-

dimensional active subspace of a system at different points in time. We approximate the

dynamic active subspace using this methodology for (i) a linear harmonic oscillator and (ii)

the enzyme kinetics system. Our methodology provides a means by which to identify the

linear combination of parameters that will change the function the most on average, and

how the linear combination changes in time.

The concept of using data-driven recovery techniques to approximate a time-dependent

sensitivity metric given snapshots of the sensitivity metric is, to the best of our knowledge,

novel. Furthermore, the methodology presented in this work can be extended to develop

time-dependent insight for other global sensitivity metrics. This intellectual and technical

contribution of the novel ideas and methodology developed in this work lend to the advance-

ment of uncertainty quantification and sensitivity analysis for time-dependent functions,

with applications ranging from epidemiological studies, to chemical kinetics, to biological

systems.
3

The remainder of this thesis proceeds as follows. We begin by reviewing ex-

isting global sensitivity analysis tools and related work and follow by defining the notation

and vocabulary used in the paper. In Section 1.4 we detail the background information on

active subspaces, dynamic mode decomposition (DMD), and sparse identification for nonlin-

ear dynamical systems (SINDy). In Section 2.2 we prove the existence of the analytic form

of three dynamic active subspaces. In Section 2.3 we propose the method of approximating

dynamic active subspaces with DMD and SINDy. We discuss the benefits and drawbacks

of such methods and conduct numerical experiments on the dynamic active subspace for an

output of a linear and nonlinear dynamical system. We finish by drawing conclusions on the

preceding work and detailing directions for future work.

A note to the reader The pronoun we will be used from the perspective of the

authors. This choice is intended to directly include the reader in the instruction, analysis,

and discussion of the following ideas.

1.2 EXISTING TOOLS AND RELATED WORK

Sensitivity analysis methods can be separated into local and global metrics. As dis-

cussed in [34], local sensitivity analysis is appropriate when the parameters of a model are

known with little uncertainty. In such cases we can evaluate the partial derivative of the

quantity of interest (QoI) with respect to its parameters to evaluate the relative change

in the QoI with respect to each parameter. However, in many biological and engineering

systems the input parameters are associated with a range of uncertainty or possibility; e.g.

glycosis modeling [46], molecular systems [43, 3], battery chemistries [37], and HIV modeling

[38]. In such cases a global sensitivity metric is necessary to understand the sensitivity of

the system to its range of inputs. There are many existing global sensitivity analysis tools,

including “One-At-A-Time” screening methods [23] such as the Morris method [36], the full

factorial method [18], the analysis of variance (ANOVA) decomposition [29], Fourier Ampli-

tude Sensitivity Test (FAST) [17], and active subspaces [13]. For a more detailed review of
4

global sensitivity analysis see [23] and [41].

Local and global metrics exist for sensitivity analysis of dynamical systems, as well.

For a system of ordinary differential equations (ODEs), time-dependent local sensitivity met-

rics are given by the augmented ODE with additional terms for the partials of the states

with respect to each parameter [21]. Similarly, [47] and [7] develop a time-dependent local

sensitivity metric for evaluating the parameter dependence of dynamic population models.

These metrics, along with others reviewed in [51], are local and only assess small perturba-

tions from the nominal parameter values. Global sensitivity analyses of dynamical systems

are often limited to analyzing an aspect of the system that is not time-dependent, such as

the steady state [1] or known stages [28] of the system. For time-dependent analyses, the

Fourier Amplitude Sensitivity Test (FAST) computes the time dependent variances of each

state of a system. In [34], the Pearson Rank Correlation Coefficient [39] is computed at

multiple points in time. These time-dependent analyses are limited to assessing the vari-

ance in or correlation between the QoI and each parameter in the system. In the case of a

high-dimensional parameter space, these individual variances become difficult to assess.

1.3 NOTATION AND VOCABULARY

In this section we define the notation and vocabulary used throughout the remainder

of the paper to create a means by which definitions and notations can easily be referenced.

Although we define these terms below, they will also be explicitly detailed in the following

sections.
5

p ∈ Rm The m real-valued parameters of a scalar-valued quantity of in-

terest. The upper and lower bounds of these parameters are given

by p` ∈ Rm and pu ∈ Rm , respectively.

f (t, p) f : Rm → R The real-valued scalar quantity of interest dependent

on m parameters p and time. In the discussions of static active

subspaces, this quantity of interest will not be dependent on time.

ρ(p) The joint probability density for the parameter space of f . For ex-

ample, in the below computations we assume that the parameters

are distributed uniformly between [p` , pu ].

∇f ∈ Rm The gradient of f (p, t) with respect to parameters p.


R
C ∈ Rm×m The matrix given by ∇f ∇f T ρ(p)dp.

W ΛW T The eigendecomposition of C with eigenvectors wi and eigenval-

ues λi for i = 1, 2, . . . , m.

W 1 ∈ Rm×k The k first eigenvectors of C corresponding to the k largest eigen-

values. This is also referred to as the active subspace of f (t, p).

DMD Shorthand for dynamic mode decomposition and broadly used to

refer to the method presented in [44] for approximating dynamics

of dynamical systems from data.

u ∈ Rq The q real-valued states of a dynamical system.

u̇ The q real-valued time derivatives of the states of a dynamical

system.

g(u) g : Rq → Rq The governing equations describing the time depen-

dence of the states of the dynamical system

XDM D ∈ Rq×(n−1) Data matrix used in DMD consisting of observations of u ∈ Rq

from time 0 to time n − 1. Note that in this data matrix, the

observations are columns.


6

0 q×(n−1)
XDM D ∈ R Data matrix used in DMD consisting of observations of u from

time 1 to time n. Note that in this data matrix, the observations

are columns.

SINDy Shorthand for sparse identification for nonlinear dynamical sys-

tems and broadly used to refer to the method presented in [4] for

recovering the governing system of dynamical systems from data.

XSIN Dy ∈ Rn×q Data matrix used in SINDy consisting of observations of u ∈ Rq at

n points in time. Note that in this data matrix, the observations

are rows.

ẊSIN Dy ∈ Rn×q Data matrix used in SINDy consisting of time derivatives of u ∈

Rq at n points in time.

Θ ∈ Rn×p Library matrix used in SINDy consisting of p transformations of

the columns of XSIN Dy ∈ Rn×q according to candidate functions

for the governing dynamical system.

X Pn An nth order monomial of the columns of XSIN Dy . For exam-

ple, for a system of three states, u ∈ R3 = [x y z]T , X P2 =

[x2 xy xz y 2 yz z 2 ].

Ξ ∈ Rp×q Coefficient matrix used in SINDy defining the contributing weight

of functions in Θ to the governing system g(u).

DyAS Shorthand for dynamic active subspaces and broadly used to refer

to the active subspace of a time-dependent quantity of interest.

w(t) ∈ Rm Used to refer to the one-dimensional dynamic active subspace of

f (t, p), corresponding to the first eigenvector of matrix C in time.


7

1.4 BACKGROUND

1.4.1 ACTIVE SUBSPACES

Consider a function f : Rm → R that maps the inputs of a system, p ∈ Rm , to a scalar

output of interest (e.g., cell count, drug concentration, temperature). We assume that the

parameters p have joint probability density ρ(p) : Rm 7→ R+ . This density represents the

uncertainty in the parameters. For example, if it is equally probable that the parameters

take any value in the range of its lower and upper bounds, p` and pu , respectively, then

p ∼ U(p` , pu ).

We assume that f is continuously differentiable almost everywhere in the support of ρ,

and has square integrable derivatives. Active subspaces [11] are eigenspaces of the matrix,
Z
C= ∇f (p)∇f (p)T ρ(p)dp = W ΛW T . (1.1)

The matrix C is the expectation of the outer product of the gradient of f , and is symmet-

ric and positive semidefinite. Thus the eigendecomposition of C yields real non-negative

eigenvalues, λ1 ≥ λ2 ≥ · · · ≥ λm , whose relationship to the function f is described by the

following lemma.

Lemma 1.4.1 (Lemma 3.1, Constantine, 2015 [11]). The mean-squared directional derivative

of f with respect to the eigenvector wi is equal to the corresponding eigenvalue,


Z
2
(∇f )T wi ρ(p)dp = λi (1.2)

Assuming the largest eigenvalue λ1 is unique, it indicates that w1 is the direction in the

parameter space along which f , changes the most on average. Let the eigendecomposition

be partitioned by
 
Λ1
 
Λ= , W = W1 W2 , (1.3)

Λ2
8

where Λ1 contains the k < m largest eigenvalues and W 1 the corresponding k eigenvectors.

We define new parameters y and z

y = W T1 p ∈ Rk , z = W T2 p ∈ Rm−k . (1.4)

The new parameters y and z are linear combinations of the original parameters p. The

parameters y are those that are most important in the following sense.

Lemma 1.4.2 (Constantine, 2015). The mean-squared gradients of f with respect to y and

z satisfy
Z
(∇y f )T (∇y f )ρ(p)dp = λ1 + . . . + λk (1.5)
Z
(∇z f )T (∇z f )ρ(p)dp = λk+1 + . . . + λm (1.6)

Thus the parameters defined by y are those for which the mean squared gradients of

f are largest in the 2-norm. Consider, for example, the case where λk+1 = . . . = λm = 0. In

this case, the only directions in which the function changes are those defined by W 1 .

Although finding the active subspace and principal component analysis (PCA) [24]

both involve the process of eigendecomposing a matrix, it is important to note that the two

analyses are not the same. The matrix to be eigendecomposed in PCA is the covariance of

the parameters p, whereas the matrix used to find the active subspace is that defined by

(1.1). The most critical distinction between these two analyses is that PCA gives insight

into the relationship between one parameter and another ; active subspaces give insight into

the relationship between the parameters and the function’s output.

1.4.1.1 Practicalities

In this section we discuss the practicalities of finding and estimating the active sub-

space. We address the convention of centering and normalizing the parameters p and when

to take the log-transform of the parameters. Additionally, we discuss the methods by which
9

we estimate (i) the gradient of f with respect to p and (ii) the integral (1.1) defining the

active subspace.

Normalizing the parameters In the analysis we assume that p ∈ [−1, 1]m . To satisfy

this assumption in practice we often need to consider an initial map that shifts and scales

the model’s natural parameter space to the normalized space. For parameters p ∈ Rm with

upper and lower bounds pu , p` respectively, the appropriate shifted and scaled parameters

are given by

diag(pu − pl )p + (pu + p` )
p∗ = , p∗ ∈ [−1, 1]m . (1.7)
2

Log-transform of parameter space When the inputs of f have units, as in the enzyme-

kinetics system we study in Section 2.3.3, we take a log-transform of the parameter space.

Such a decision is made from experience and insight with other dimensionalized systems.

There is an interesting connection between the classical process of finding non-dimensional

parameters for a physical system and identifying important linear combinations of parame-

ters. Since the non-dimensional parameters are products of powers of the original parameters,

a log-transform produces a linear combination of the logs of the original parameters. This im-

plies that, for systems where the original parameters are amenable to non-dimensionalization,

the log-transformed system depends only on a few linear combinations. Such structure is

closely related to active subspaces; see [12] for a detailed discussion.

Estimating gradients In practice we do not have the analytic form of ∇f (p) and thus

must approximate it. If this is the case one may approximate the gradient using first order

forward finite differences,

∂f f (pi + h) − f (pi )
≈ , i = 1, 2, ..., m. (1.8)
∂pi h

For dynamical systems with an analytical form ∇f (t, p) can be approximated by augmenting

the original ODE system with the time derivative of the states’ partial derivatives. These
10

will be referred to as Gronwall Gradients [21] and are detailed in Section 1.4.2.

Estimating the expectation The expectation defined by the integral in (1.1) must be

approximated. For this paper we employ two methods for approximating this integral using

(i) Monte Carlo and (ii) Gauss-Legendre quadrature. Both approximation rules are achieved

by the following algorithm.

Algorithm 1 Estimation of the active subspace.


Given M quadrature weights ωi ∈ R and points pi ∈ Rm :

(1) For each point pi compute the quantity of interest f (pi )

(2) Estimate the gradient ∇p fi = ∇p f (pi ).

(3) Compute the matrix Ĉ and its eigenvalue decomposition,


M
X
Ĉ = ωi ∇p fi ∇p fiT = Ŵ Λ̂Ŵ (1.9)
i=1

The choice of M depends on f and the dimension of the parameter space. With Monte

Carlo sampling, the points pi are randomly drawn according to density ρ(p) with weights

ωi = 1
M
. The estimate for Ĉ converges to C like O(M −1/2 ). When the dimension m of f is

low, and when f has sufficiently smooth derivatives we can achieve a much better approx-

imation using a Gauss quadrature rule to approximate C. Here, M nodes pi and weights

ωi are drawn per dimension according to the tensor product Gauss-Legendre quadrature

scheme. It is important to note that the number of nodes M per dimension m leads to M m

total function evaluations and gradient approximations. The choice of M again depends on

f and in practice one should conduct numerical convergence studies to find an appropriate

M . We implement and focus on these two methods for approximating the active subspace,

although additional approaches can be found in [11].


11

1.4.2 GRONWALL GRADIENTS

Assume that the q states of the dynamical system (1.12) are dependent upon at least

one parameter p. In [21], Gronwall derives a system of ordinary differential equations that
∂ui
includes the dynamics of both the original state variables and of their partial derivatives, .
∂p
Numerically integrating the system provides an approximation for the partial derivatives of

the state with respect to a parameter, p as a function of time. In the case where there is

more than one parameter, the system is easily extended to include the dynamics of these

partial derivatives. This approach for approximating partial derivatives was developed as a

local sensitivity metric, but can be used as an efficient way to approximate the gradient of

a function.

1.4.2.1 Description of variables

Let the following system of q differential equations describe the dynamics of the states

ui over time for i = 1, . . . , q with parameters pk for k = 1, . . . , m:

dui
= gi (t; u1 , . . . , uq ; p1 , . . . pm ). (1.10)
dt
∂gi
Let the initial conditions ui = u0i for t = t0 and assume that for j = 1, . . . , q the partials ∂uj
∂gi
and ∂pk
are continuous [21].

1.4.2.2 Extended dynamical system with partial derivatives

∂ui
The partial derivative of state ui with respect to parameter pk is written ∂pk
. The

differential equation that describes the dynamics of these partials is as follows.


  X
d ∂ui ∂gi ∂uj ∂gi
= (t; u1 , . . . , uq ; p1 , . . . pm ) + (t; u1 , . . . , uq ; p1 , . . . pm ) (1.11)
dt ∂pk j
∂u j ∂p k ∂p k

As defined, the initial conditions for all of the partial derivatives are zero. The dynamical

system resulting from augmenting (1.10) with (1.11) has q+(m×q) states. The integration of

this system will result in the solutions of the states in time and the solutions of the partials
12

in time. For our purposes we approximate the integration of the system using python’s

scipy.integrate.odeint function with parameter values defined by the integration nodes

used to approximate C.

1.5 DATA-DRIVEN RECOVERY OF DYNAMICAL SYSTEMS

A parameterized dynamical system is a system in which the time dependence of the

states is described by governing equations dependent on states and parameters. The vector

form of (1.10) is written,

u̇ = g(u, p), u ∈ Rq , p ∈ Rm , g : Rq+m −→ Rq . (1.12)

In this form, u̇ denotes the time derivative of the states u, and g denotes a continuous

function of the states describing the governing time-dependence. The governing equations of

dynamical systems enables in-depth analysis and computational simulations of the physical

systems they model. However, in many systems the governing equations are not known.

To address such situations, methods exist to recover the underlying dynamical system and

predict future states from observations of the system with a local linear approximation [44];

recovering governing equations by solving a constrained optimization problem [49]; recovering

governing equations through a sparse least squares problem [4]; nonlinear regression analysis

[52]; iterative symbolic regression [45]; nonlinear adaptive projection onto a sparse basis

[42]; and with embedding dimension and global equations of motion analysis [16]. Below

we present the methods of: (i) dynamic mode decomposition (DMD) developed by Schmid

[44] and further analyzed by Kutz, et al. in [30]; and (ii) sparse identification of nonlinear

dynamical systems (SINDy) developed by Brunton et al. in [4]. We review these data-driven

recovery methods to motivate their use for recovering and approximating the dynamics of

time-dependent active subspaces.


13

1.5.1 DYNAMIC MODE DECOMPOSITION

Assume that we can observe a dynamic process with q state variables from which we

can collect n snapshots of data. We further assume that the dynamical system (1.12) has

corresponding discrete-time map given by

uk+1 = F(uk ). (1.13)

for k = 1, 2, 3, . . . n − 1 time steps. DMD approximates this discrete-time map by

uk+1 ≈ Auk (1.14)

where A minimizes ||uk+1 − Auk ||2 over the data snapshots. Although DMD can be used

to provide insight into the dynamical system, we implement the algorithm to exploit only

its ability to estimate states in between data snapshots and predict future states using the

approximate discrete-time map given by (1.14).

The definition of the dynamic mode decomposition as given by [50] is:

Suppose we have a dynamical system (1.12) and two sets of data,


 

X = u(t1 ) u(t2 ) . . . u(tn−1 ) , (1.15)


 

X 0 = u0 (t2 ) u0 (t3 ) . . . u0 (tn ) (1.16)

so that u0 k = F(uk ), where F is the map in (1.13) corresponding to the evo-


lution of (1.12) for time ∆t. DMD computes the leading eigendecomposition
of the best-fit linear operator A relating the data X 0 ≈ AX :
A = X 0X †. (1.17)
The DMD nodes, also called dynamic modes, are the eigenvectors of A, and
each DMD node corresponds to a particular eigenvalue of A.

1.5.1.1 The method

We implement a simple approach to the DMD algorithm presented in [30]. The algo-

rithm as-is in [30] consists of using a low-rank projection of A onto its proper orthogonal
14

decomposition (POD) modes. For our application of DMD we want the full-dimensional

representation of the dynamics and thus avoid this step, however it should be noted that for

other applications Algorithm 1.1 in [30] can be used exactly.


15
Algorithm 2 Dynamic mode decomposition
Given matrices of snapshots X and X 0 :

(1) Take the singular value decomposition (SVD) of X:

X = U ΣV T , (1.18)

(2) Construct matrix A from (1.17) with the pseudoinverse of X given by the above

SVD:

X † = V Σ−1 U T , (1.19)

A = X 0 V Σ−1 U T . (1.20)

(3) Compute the eigendecomposition of A:

AW = W Λ, (1.21)

where the eigenvectors of A are the columns of W and the eigenvalues λi are the

diagonal entries of Λ.
(4) Construct

Φ = X 0 V Σ−1 W . (1.22)

(5) The future states of the system are approximated by

x(t) ≈ Σrk=1 φk exp (ωk t)bk = Φexp(Ωt)b (1.23)

where ωk = ln (λk )/∆t, and b = Φ† x1 where b is the least-squares solution to

x1 = Φb.

Thus, given snapshots of a dynamical system in time, we can approximate the dynamics
16

and future states with a locally linear approach given by dynamic mode decomposition.

1.5.2 SPARSE IDENTIFICATION OF NONLINEAR DYNAMICAL SYS-

TEMS

The method of sparse identification of nonlinear dynamical systems (SINDy) uses snap-

shots of a dynamical system to recover the governing equations of the system. We assume

that we can observe and measure a time-dependent process of one or more states. For ex-

ample, for the movement of an oscillator in three dimensional space, the state vector u ∈ R3

would have three components x, y, z. We aim to recover the governing system of equations

for the dynamical system of interest.

1.5.2.1 Setting up the overdetermined least squares problem

We form the data matrix X ∈ Rn×q by collecting observations of u ∈ Rq at n points

in time. We approximate the time derivative of these states in time using a differentiation

scheme to form Ẋ. In this paper we implement a first order forward difference scheme,
u(ti + h) − u(ti )
u̇i ≈ (1.24)
h
where ui are the observations at time ti and h is the time-step used to compute derivatives.

The authors of [4] suggest that when the snapshots are collected from a system with noise,

a total variation regularized derivative scheme [8] should be used to compute Ẋ.

We next construct the library matrix Θ ∈ Rn×p . This library matrix is formed by

taking p functional transformations of the columns of X and represents the possibilities of

the functions which might be present in the true governing system g. Thus we have
   
T T
u (t1 ) u̇ (t1 )
   
   
 uT (t2 )   u̇T (t2 ) 
X =  .  Ẋ =  .  (1.25)
   
 .   . 
 .   . 
   
T T
u (tn ) u̇ (tn )
17
 
Θ(X) = 1 X X P2 X P3 · · · sin(X) cos(X) · · · . (1.26)

The notation X Pi indicates monomials of the states up to order i. For example, for the

states x, y, z of a system, X P2 = [x2 , y 2 , z 2 , xy, xz, yz].

We now seek a coefficient matrix Ξ ∈ Rp×q ,


 
Ξ = ξ1 ξ1 . . . ξq , ξi ∈ Rp (1.27)

such that

Ẋ ≈ Θ(X)Ξ (1.28)

The coefficients of Ξ will thus indicate which, and how much, of the functions of Θ are in the

true governing equations, g(u). Given the various approximations present in the formulation

of (1.28) (i.e., data observation errors, approximation of Ẋ, and errors in Θ(X)), an exact

solution is unlikely or impossible. Thus we seek a solution to the minimization problem given

by

min ||ẋi − Θ(X)ξi ||22 + α||ξ||1 , i = 1, 2, · · · , q.



(1.29)
ξi

A note on the choice of functions in Θ The transformations chosen can be informed

by intuition of the underlying physical system or of the observed dynamics. The authors of

[4], however, argue that the library matrix can contain as many functional transformations

as the user chooses. This argument is based on the experience that unnecessary functions

present in Θ will result in a corresponding 0 coefficient in Ξ. If, however, a function present

in the true governing equations is not included in the library matrix, it will be impossible

to recover with this algorithm. If we know nothing about the true governing equations of

the system from which we collect data, then we have no way to verify whether or not we are

actually recovering the same system.


18

1.5.2.2 Sequential thresholded least squares

The method of sequential thresholded least squares was developed in [4] to solve the

problem given by (1.29). At the heart of this algorithm is an assumption of sparsity in the

space of all functions. The thresholding aspect of the method imposes this sparsity and aids

in the recovery of the underlying dynamical system. As an example of how this sparsity
x2 x3
may be imposed, consider the function ex ≈ 1 + x + 2
+ 6
.
If we observed data from this

function and formed the library matrix as Θ(X) = eX 1 X X X 2 3 A reasonable

solution to (1.29) could either be Ξ = [1 0 0 0 0]T or Ξ = [0 1 1 21 16 ]T . The solution vector

identifying only ex , however, is most sparse. The following method has been developed for

such examples and has been used to recover the governing dynamics of many systems [4].

Algorithm 3 Sequential thresholded least squares [4]

Given Ẋ, Θ(X), and parameters λ and k:

(1) Let Ξ0 be the least squares solution to the problem,

min ||Ẋ − Θ(X)Ξ||22 . (1.30)


Ξ

(2) Repeat k times:

(a) For all entries ξi,j of matrix Ξ, if ξi,j < λ, update such that ξi,j = 0.

(b) for each column ξj of Ξ and column , u̇j of Ẋ:

(i) Let r1 , . . . r` be the rows of ξj whose entries are not equal to zero. Let ξ ∗

be the column vector whose rows are the corresponding nonzero entries of

ξj .

(ii) Let Θ∗ be the matrix whose columns are the columns r1 , . . . r` of Θ.

(iii) Update ξ ∗ = leastsquares(u̇j , Θ∗ ).


19

The coefficient matrix, Ξ returned from Algorithm 3 indicates which functions of the

library matrix are in the true governing equations, g(x). For example, in a three state system

where  
0 3 22  
 
Ξ = 6 0 9  , Θ = sin x xy z ,
  3
 
0 0.8 2
the governing equations recovered by SINDy would be

ẋ = 6xy

ẏ = 3 sin x + 0.8z 3

ż = 22 sin x + 9xy + z 3

Practicalities

Assumption of sparsity The choice of using this method to solve the overdetermined

least squares problem is based on a few assumptions that are necessary to address. The first

assumption is that the functions composing the governing system g(x) are within the library

matrix Θ. The second assumption is that the true governing system is sparse in the space

of all functions. This means that of all possible functions, the governing equations of the

dynamical system will only be composed of a few. The validity of this assumption for the

case in which we will employ the algorithm will be discussed in Section 1.5.2.3.

Choice of tuning parameter λ The sequential thresholded least squares algorithm

for solving the `1 minimization problem given by (1.29) is extremely and unpredictably sensi-

tive to the choice of λ. Because the values of the coefficient matrix are thresholded according

to this parameter at each step, the choice of λ will drastically affect the recovered dynam-

ical system. This sensitivity and unpredictability of the choice of λ is also what makes

the complexity analysis of this algorithm difficult. The cost of each least squares computa-

tion at each iteration in the sequential thresholded least squares approach to the problem is

determined by the size of the matrix and vector. These sizes are dependent on the value of λ.
20

1.5.2.3 Least absolute shrinkage and selection operator (LASSO)

Although sequential thresholded least squares has been empirically shown to recover the

governing dynamical system, the practicalities noted above make it difficult to understand

how the method of SINDy can be translated to new problems. For example, to actually

recover a dynamical system about which one has no intuition (stay tuned!), choosing an

appropriate value for λ can be difficult. Alternatively, if there is no system governing the

observed dynamics, one may want to model the dynamics instead. In such a case, this notion

of sparsity may be incorrect. Thus we attempt to find a solution to the `1 minimization

least squares problem (1.29) by taking advantage of the well-developed regression analysis

methods developed to solve the problem of the least absolute shrinkage and selection operator

(LASSO) [48] [22], defined as (1.29). For our purposes we use the Lasso function in python’s

scikit-learn package. This method for solving (1.29) involves the tuning of the parameter

α which controls the relative importance of the 1-norm of the solution. The effect of this

parameter on the solution is studied in Section 2.3.3.

Ill-conditioning The Θ may be ill-conditioned. This ill conditioning affects the

ability to recover a solution with least squares. To assist in the solution we impose the `1

regularization condition defined by the α term in (1.29).

Now that we’ve presented the necessary background information on active subspaces

and the data-driven recovery of dynamical systems techniques of DMD and SINDy, we

proceed to introduce the idea of dynamic active subspaces.


CHAPTER 2

2.1 DYNAMIC ACTIVE SUBSPACES

Active subspaces enable dimension reduction and sensitivity analysis for computational

models. As is, this analysis does not explicitly extend to functions that are dependent on both

parameter inputs and time. Extending active subspaces to time-dependent systems enables

uncertainty quantification, sensitivity analysis, and parameter estimation for computational

models that have explicit dependence on time.

2.1.1 The problem

To begin, consider a system of parameterized ordinary differential equations with q

states u and m parameters p,

u̇ = g(u, p), u ∈ Rq , p ∈ Rm . (2.1)

Let us consider a specific time-dependent quantity of interest (QoI) of these states, dependent

on states u ∈ Rq and parameters p ∈ Rm . This QoI is defined as

f (t, p) = F (u(t, p)) : Rm → R. (2.2)

Note that f can be a nonlinear function of the states, and that these states can be nonlinearly

dependent on the parameters. In practice the solution u(t, p) is numerically computed.


22

2.1.2 The proposed solution

The main idea Suppose we want to identify the linear combinations of the pa-

rameters that change the QoI the most on average, and we want to know how these linear

combinations change in time. Doing so would identify the dynamic active subspaces (DyAS)

and would give quantifiable insight into the parameter sensitivity of time-dependent sys-

tems. We propose and discuss the following three approaches to identifying dynamic active

subspaces.

Computing DyAS at independent times One option to identify these time-

dependent active subspaces is to independently compute them at each step in time. Specif-

ically, we could find the QoI at time t0 , compute the active subspace, find the QoI at time

t1 , compute the active subspace, find the QoI at time t2 , compute the active subspace, and

repeat this until we reach final time tn . Such analysis and identification has been done in

[33] and [14]. This approach to finding time-dependent active subspaces, however, is ex-

tremely expensive, especially in systems with many parameters. This computation can be

reduced by only computing active subspaces at “interesting” points in time (i.e., stages of

a disease, transient dynamics, final time, time of maximum concentration). However, what

is considered an “interesting” point in time for nominal parameter values might not be for

other values in the parameter space.

Analytically defining DyAS Rather than computing the active subspace inde-

pendently throughout time, consider the possibility of defining a time-dependent C(t) with

eigenvalues λi (t) and a governing dynamical system for wi (t). The computation involved in

computing the active subspace at each time step could be bypassed by simply integrating the

dynamical system for wi (t). Classical analysis for dynamical systems could then give further

insight into these dynamic active subspaces and, in turn, the physical systems from which

they are derived. In Section 2.2 we present three cases for which we can use this approach

to define DyAS.
23

Approximating the DyAS In cases for which there does not exist an analytical

form for the DyAS or for which there is no analytical form for the dynamical system, we

can consider approximating the dynamics of the active subspace. Below we propose the use

of DMD and SINDy (see Section 1.5 for details) to approximate the one-dimensional DyAS

from snapshots of the first eigenvector of C in time.

2.2 ANALYTIC DYNAMIC ACTIVE SUBSPACES

Just as we can describe the dynamics of the state space of a physical system with

an ordinary differential equation, we propose similarly describing the dynamics of the state

space of an eigenvector. An ODE describing the first eigenvector of C over time would give

an analytical form of the time derivative for the one-dimensional dynamic active subspace.

Below we prove the existence of two such dynamical systems for dynamic active subspaces.

2.2.1 DyAS for an homogeneous linear dynamical system

Consider a linear dynamical system,

u̇ = Au, u(0) = p, u(t, p), p ∈ Rq , A ∈ Rq×q . (2.3)

The solution to this dynamical system is,

u(t, p) = eAt p. (2.4)

Consider a time-dependent scalar output of the dynamical system, f (t, p) = φT u(t, p), for

φ ∈ Rq . For example, where f (t, p) is the ith state of the dynamical system, φ is the ith

column of the identity matrix. Assume that the initial conditions of the dynamical system

(2.3) have joint density ρ(p). Thus the time-dependent scalar output of the parameterized

dynamical system (2.3) is

f (t, p) = φT eAt p. (2.5)


24

The exact time-dependent gradient of f is

∇f (t, p) = (eAt )T φ, (2.6)

and thus the time-dependent matrix C(t) is,


Z
T
(eAt )T φ (eAt )T φ ρ(p)dp

C(t) = (2.7)
Z
(eAt )T φ φT eAt ρ(p)dp
 
= (2.8)
Z
At T T At
= (e ) φφ e ρ(p)dp (2.9)

= (eAt )T φφT eAt . (2.10)

Note that Rank(φφT ) = 1 and thus Rank(C(t)) = 1. Therefore C(t) has only one

nonzero eigenvalue with corresponding normalized eigenvector,

(eAt )T φ
v1 (t) = . (2.11)
||(eAt )T φ||2

This eigenvector is only unique up to a constant. Thus we seek a governing dynamical system

for the unnormalized first eigenvector.

Let w(t) = (eAt )T φ. Then the dynamical system for the unnormalized first eigenvector

of C(t) is described by the linear dynamical system,

ẇ(t) = AT w(t), w(0) = φ. (2.12)

2.2.2 DyAS for an inhomogeneous linear dynamical system

Consider the inhomogeneous linear dynamical system,

u̇ = Au + p, u(0) = η, u(t, p), η, p ∈ Rq , A ∈ Rq×q . (2.13)

For invertible A the solution to this dynamical system is given by Duhamel’s Principle [32],

u(t, p) = eAt η + A−1 (eAt − I)p. (2.14)


25

Consider a time-dependent scalar output of the dynamical system, f (t, p) = φT u(t, p), for

φ ∈ Rn . Assume that p has a joint density ρ(p). Thus the time-dependent scalar output of

the parameterized dynamical system (2.13) is

f (t, p) = φT eAt η + φT A−1 (eAt − I)p. (2.15)

The exact time-dependent gradients of f are

∇f (t, p) = ((eAt )T − I)A−T φ, (2.16)

and thus the time-dependent matrix C(t) is given by


Z
T
((eAt )T − I)A−T φ ((eAt )T − I)A−T φ ρ(p)dp

C(t) = (2.17)
Z
= ((eAt )T − I)A−T φφT A−1 (eAt − I)ρ(p)dp (2.18)
Z
At T −T T −1 At
= ((e ) − I)A φφ A (e − I) ρ(p)dp (2.19)

= ((eAt )T − I)A−T φφT A−1 (eAt − I). (2.20)

Note that Rank(φφT ) = 1 and thus Rank(C(t)) = 1. Therefore C(t) has only one

nonzero eigenvalue, with corresponding normalized eigenvector,

((eAt )T − I)A−T φ
v1 (t) = . (2.21)
||((eAt )T − I)A−T φ||2

This eigenvector is only unique up to a constant. Thus we seek a governing dynamical system

for the unnormalized first eigenvector.

Let

w(t) = ((eAt )T − I)A−T φ. (2.22)

Then the dynamical system of the unnormalized first eigenvector of C(t) is described by the

inhomogeneous linear dynamical system,

ẇ(t) = AT w(t) + φ, w(0) = 0. (2.23)


26

Note that Duhamel’s principle on (2.23) yields

w(t) = (eAt )T w0 + A−T ((eAt )T − I)φ = A−T ((eAt )T − I)φ.

Consider A = W ΛW −1 and recall the definition of the matrix exponential [32], eAt =

W eΛt W −1 . The expression for w(t) given by Duhamel’s principle is equivalent to (2.22),

A−T ((eAt )T − I)φ = A−T (eAt )T φ − A−T φ

= W −T Λ−1 W T W −T eΛt W T φ − A−T φ

= W −T Λ−1 eΛt W T φ − A−T φ

= W −T eΛt Λ−1 W T φ − A−T φ


T
= eA t A−T φ − A−T φ

= ((eAt )T − I)A−T φ.

The analyses of sections 2.2.1 and 2.2.2 can be extended for the case where the pa-

rameters of an inhomogeneous linear dynamical system (2.13) are the initial conditions η.

Thus we prove the existence for the analytic form of the dynamic active subspace for three

cases. For these specific problems, the cost of computing the active subspace at each time

point of interest is avoided by simply evaluating a linear dynamical system which gives ex-

actly the one-dimensional active subspace of the time-dependent QoI. The existence of the

governing equations of the dynamic active subspace for these cases is directly comparable

to the existence of an analytical static active subspace for a QoI with linear dependence on

its parameters. As we saw in the above analyses, when the gradient of f is not parameter-

dependent, the outer product in (1.1) can be pulled outside of the integral, and the resulting

matrix will be rank-one. A rank-one C guarantees a one-dimensional active subspace.

However exciting as this may be (we think it’s pretty cool), we must acknowledge that

these analytic forms exist (so far!) for three very specific cases. We would like to have a

form of the dynamic active subspace for any time dependent, differentiable QoI. Thus in the

following section we propose a method for finding an approximate dynamic active subspace.
27

2.3 APPROXIMATING DYNAMIC ACTIVE SUBSPACES

In this section we propose the use of DMD (see Section 1.5.1) and SINDy (see Section

1.5.2) to approximate the dynamic active subspace.

2.3.1 Introduction to the methodology

The methods of DMD and SINDy to recover and approximate dynamics rely on a

collection of snapshots of the system. In the case of dynamic active subspaces, these snap-

shots will be components of the first eigenvector of C in time. We propose collecting these

snapshots and using DMD and SINDy to approximate the dynamic active subspace between

snapshots and predict future DyAS. Below we present the methodology.

Methodology Approximating the dynamic active subspace

(1) Choose a quantity of interest from the parameterized dynamical system.

(2) Augment the dynamical system with states for the partial derivatives to compute

Gronwall gradients (1.11) (Alternatively, use a finite difference method to approxi-

mate the gradients ∇f (p, t)).

(3) Identify a numerical integration rule for approximating C.

(4) For each point pi in the set of integration nodes, integrate the dynamical system to

compute f (ti , pi ) and ∇f (ti , pi ) for each ti ∈ [t1 , t2 , t3 , · · · , tn−1 , tn ].

(5) For each ti , estimate C(ti ) and compute the first eigenvector w(ti ).

(a) If using SINDy to approximate the dynamics, estimate C(ti ) and w(ti ) at n

additional points in time, [t1 + h, t2 + h, t3 + h, · · · , tn−1 + h, tn + h] for small

h. Note that h is the time step used to approximate the time derivative for the

dynamics of the eigenvector.


28

(6) Construct the data matrices with the first eigenvector w at each ti

For DMD:

(a) Construct XDM D ∈ Rm×(n−1) with the first (n − 1) eigenvectors:


 
 
 
XDM D = w(t1 ) w(t2 ) · · · w(tn−2 ) w(tn−1 )


 

0 m×(n−1)
(b) Construct XDM D ∈ R with the last (n − 1) eigenvectors:
 
 
0
 
XDM D =
 w(t2 ) w(t 3 ) · · · w(tn−1 ) w(t )
n 

 

For SINDy:

(a) Construct XSIN Dy ∈ Rn×m with the first eigenvector w at [t1 , t2 , t3 , · · · , tn−1 , tn ]:
 
T
 w(t1 ) 
 
 w(t )T 
 2 
..
 
XSIN Dy = 
 
 . 

 
w(tn−1 )T 
 
 
w(tn )T

(b) Use a first order finite difference scheme,


w(ti + h) − w(ti )
ẇ(ti ) ≈
h
to construct ẊSIN Dy ∈ Rn×m :
 
T
 ẇ(t1 ) 
 
 ẇ(t )T 
 2 
..
 
ẊSIN Dy =
 
 . 

 
ẇ(tn−1 )T 
 
 
ẇ(tn )T
29

(7) Let T ∈ Rn be the vector of times at which to predict the dynamic active subspace.

Use Algorithm 2 and/or Algorithm 3 to construct an approximation for the dynamic

active subspace for all t ∈ T.

Below we demonstrate this methodology to approximate the dynamic active subspace for

two dynamical systems: (i) a linear harmonic oscillator and (ii) an enzyme kinetics system.

2.3.2 Example: Linear Harmonic Oscillator

Consider a linear harmonic oscillator whose dynamics are given by


 
−0.1 −2 0 
 
u̇ = 
 2 −0.1 0  u, u(t0 ) = p
 (2.24)
 
0 0 −0.3

The initial conditions, p ∈ R3 are uncertain between the lower and upper bounds,

p` = [ 0.96, 4.0, 8.0 ]T , pu = [ 1.44, 6.0, 12.0 ]T .

This dynamical system has an analytic solution given by


 
−0.1 −2 0 
 
u(t) = eAt p, A =  2 −0.1 0 . (2.25)
 
0 0 −0.3

At nominal initial conditions, pnominal = [ 1.2, 5.0, 10.0 ]T , the trajectory of the linear har-

monic oscillator in 3 dimensions is given by Figure 2.1.


30

Figure 2.1: The trajectory of the x, y, and z components of the linear harmonic oscillator

given by (2.24) with nominal initial conditions.

Consider the y component as the quantity of interest. Then for φ = [ 0, 1, 0 ]T , let

f (t, p) = φT u = φT eAt p. (2.26)

Note that for this system with the initial conditions as parameters, the governing system of

the dynamic active subspace is given by (2.12). For the sake of demonstrating the methodol-

ogy for this simple example we ignore this until it comes time to evaluate our approximation.

The gradient of this function with respect to p can be taken exactly from (2.26) for all

time,

T
∇f (t, p) = eA t φ. (2.27)

We will use this evaluation of the gradient to approximate C with Algorithm 1 and Monte

Carlo sampling. Although we would normally conduct a convergence study on the number

of samples M for our Monte Carlo approximation, it is not necessary in this case because we

know that the gradients (2.27) are independent of our samples pi . To have enough points

to construct the shadow plot in Figure 2.2, we let N = 50. We assume that p is uniformly

distributed with upper and lower bounds 20% above and below the nominal values (see Table

2.1).
31
Parameter Lower Bound Nominal Value Upper Bound

x0 0.96 1.2 1.44

y0 4.0 5.0 6.0

z0 8.0 10.0 12.0

Table 2.1: The initial conditions p in (2.26) have a uniform joint density with lower and

upper bounds 20% below and above nominal values.

To examine and explain the different aspects of the active subspace, we first compute

C at time t = 5.

Figure 2.2: The eigenvalues (left) and first eigenvector (middle) of C at time t = 5. Note

that there is only one nonzero eigenvalue. The shadow plot (right) at this time shows the

value of the linear harmonic oscillator y coordinate as a function of a linear combination of

possible initial conditions, p.

The rightmost plot of Figure 2.2 shows the shadow plot for the active subspace at time

t = 5. The shadow plot reveals the low-dimensional structure in f (t = 5, p) by viewing the

function evaluations in the direction defined by w.

Given the methodology from Section 2.3.1, we seek to form data matrices using snap-

shots of the eigenvector w at n points in time. We compute the first eigenvector from

t0 = 0.001 to tF = 10.0 at 8000 points in time by independently computing the active sub-

space at each time step. The time-series of this eigenvector will be used as the “truth” to
32

which we will compare all future approximations. Figure 2.3 shows an interesting aspect

of the computation. Using numpy’s eig function, the eigenvectors of C are normalized at

each time step. Using snapshots from the left of Figure 2.3 will be problematic if aiming

to use a linear approximation method (DMD) or a derivative-based method (SINDy) in re-

covering the dynamics. Following analysis from Section 2.2 we observe that multiplying the

computed time series by the square root of the first computed eigenvalue of C yields the

same dynamics given by the dynamical system (2.12). Thus in the following experiments we

use this unnormalized form of the eigenvector to construct the data matrices for DMD and

SINDy.

Important to note is that the sign of an eigenvector is not unique (an eigenvector is

still an eigenvector if multiplied by negative one). Thus when computing eigenvectors for

multiple points in time, we must normalize the sign of the eigenvector at each time step. We

do so by checking,

||w(ti ) + w(ti+1 )||2 > ||w(ti ) − w(ti+1 )||2 . (2.28)

If (2.28) is true, w(ti+1 ) is updated such that w(ti+1 ) = −w(ti+1 ). This ensures that any

discontinuity observed in the time-dependent eigenvector is not due to a difference in signs.

Figure 2.3: (Left) the first eigenvector of C computed for the function (2.26) (Right) the
p
first eigenvector after being multiplied by λ0 (t).
33

2.3.2.1 Numerical Experiments

The methods of DMD and SINDy depend on data matrices formed with snapshots

of the dynamical system in time. We test the accuracy of these methods in recovering the

dynamic active subspace as a function of (i) the proportion of time the snapshots represent,

i.e. snapshots from the first two out of ten seconds of a system would represent 0.2 of the

data, and (ii) the number of samples in the data matrices.

In the following experiments we construct the data matrices XDM D and XSIN Dy with

snapshots from proportions of the data from 0.1 to 1.0. For each proportion we also construct

the data matrices with 10, 20, 30, 40, 50, 75, 100, 200, 500, 1000, 2000, and 5000 snapshots.

In the implementation of SINDy we use scikit-learn’s Lasso function with α = 10−5 .

In order to construct a data matrix XSIN Dy with N snapshots at times t0 , t1 , · · · , tN , we

compute an additional N eigenvectors at t0 + h, t1 + h, · · · , tN + h for h = 10−6 with which

to approximate Ẋ. The library matrix Θ is constructed with monomials of up to degree

5, trigonometric combinations of up to degree 3, and exponential combinations of up to

degree 3. In Figure 2.4, the error for each experiment is computed as the maximum absolute

error between the approximated dynamic active subspace, and the “true” active subspace

computed at all 8000 time steps.


34

Figure 2.4: The maximum absolute error of the dynamic active subspace approximated

by (left) DMD and (right) SINDy. The data matrices for each algorithm have been created

with N samples from p proportion of the total time.

One of the reasons DMD and SINDy poorly approximate the dynamic active subspace

when given few snapshots of data is due to the non-uniqueness of eigenvectors discussed

above. For a small time step between computed eigenvectors (i.e., there are many snapshots),

the normalization of signs by checking (2.28) ensures that the nearby eigenvectors have

the same sign. However, when the time step between computed eigenvectors is large, the

trajectory may have evolved significantly enough that (2.28) will not properly indicate what

the sign “should” be. In such cases, the DMD and SINDy data matrices will be given

snapshots that do not match the “true” state of the eigenvector.

2.3.3 Enzyme Kinetics

In this section we extend the methodology presented in Section 2.3.1 to a nonlinear

dynamical system. The simple enzyme kinetics system given by (2.29) describes the chemical

reaction of the catalysis of enzymes [26], and the reaction is written as,

k+ k c
−−
S+E )−*− ES −− → P + E. (2.29)
− k
35

The reaction (2.29) is described by the system of nonlinear ordinary differential equations

in (2.33).

d[S]
= −k + [E][S] + k − [ES] (2.30)
dt
d[E]
= −k + [E][S] + (k − + k c )[ES] (2.31)
dt
d[P ]
= k c [ES] (2.32)
dt
d[ES]
= k + [E][S] − (k − + k c )[ES] (2.33)
dt

Let us consider the concentration of [ES] in time our quantity of interest,

f (t, p) = [ES](t, p) (2.34)

We assume that the parameters, p = [k + , k − , k c ]T have uniform joint density with lower and

upper bounds 50% below and above the nominal values (see Table 2.2).

Parameter Lower Bound Nominal Value Upper Bound

k+ 1.0 2.0 3.0

k− 0.1 0.2 0.3

kc 0.05 0.1 0.15

Table 2.2: The parameters p have uniform joint density with upper and lower bounds 50%

above and below the nominal values.

Approximation of C Given that our quantity of interest only has 3 parameters,

we approximate the integral in (1.1) with the tensor product Gauss-Legendre quadrature

scheme. We conduct a numerical experiment to choose the number of quadrature points per

dimension, M , in approximating (1.1). We choose to use 7 quadrature points per dimension

for a total number of 73 = 343 sample points (see Figure 2.5).


36

Figure 2.5: The convergence of the approximation to the integral in (1.1) for the number

of nodes per dimension in the tensor product Gauss-Legendre quadrature scheme.


37

Gradients We augment the system of ODEs in (2.33) to include the Gronwall Gra-

dients of each state with respect to each parameter according to (1.11).


   
+ −
[S] −k [E][S] + k [ES]
   
   
+ − c
 [E]  
   −k [E][S] + (k + k )[ES] 

   
 [P ]  
   k + [ES] 

   
+ − c
k [E][S] − (k + k )[ES]
   
[ES]  
   
 ∂[S]  
 +   −k + [E] ∂[S]+ − k + [S] ∂[E] − ∂[ES]

 ∂k   ∂k ∂k+
+ k ∂k+
− [E][S] 

   
 ∂[E]  −k + [E] ∂[S] − k + [S] ∂[E] + (k − + k c ) ∂[ES] − [E][S]
 ∂k+   ∂k+ ∂k+ ∂k+ 
   
 ∂[P ]   c ∂[ES] 
 ∂k+   k ∂k+ 
   
 ∂[ES]   ∂[S] ∂[E] c ∂[ES]

+ + −
d  ∂k+  
   k [E] ∂k+
+ k [S] ∂k+
− (k + k ) ∂k+
+ [E][S] 
= (2.35)

dt  ∂[S]  
 
∂[S] ∂[E] − ∂[ES]
−k + [E] ∂k − k +
[S] + k + [ES]

 ∂k−   − ∂k− ∂k− 
   
 ∂[E]   + ∂[S] + ∂[E] − c ∂[ES]
 ∂k−   −k [E] ∂k− − k [S] ∂k− + (k + k ) ∂k− + [ES] 

   
 ∂[P ]   
c ∂[ES]
 − 
 ∂k   k ∂k −


   
 ∂[ES]   + ∂[S] ∂[E] ∂[ES]
 ∂k−   k [E] ∂k− + k + [S] ∂k− − (k − + k c ) ∂k− − [ES] 

   
 ∂[S]   + ∂[S] + ∂[E] − ∂[ES] 
 ∂kc  
   −k [E] ∂kc − k [S] ∂kc + k ∂kc 

 ∂[E]   
 c   −k + [E] ∂[S]c − k + [S] ∂[E]c + (k − + k c ) ∂[ES] c + [ES] 
 ∂k   ∂k ∂k ∂k 
   
 ∂[P ]  
k c ∂[ES] + [ES]

 ∂kc   ∂k c 
   
∂[ES] + ∂[S] + ∂[E] − c ∂[ES]
∂kc
k [E] ∂kc + k [S] ∂kc − (k + k ) ∂kc − [ES]

We compare the solutions for the partial derivative of f with respect to the parameters

numerically integrated from (2.35) with a second-order finite difference method for decreasing

h as well as over time for a fixed h = 10−3 (Figure 2.6). We compare the wall clock time for

each method (Figure 2.7).


38

Figure 2.6: (Left) the average absolute error over time between Gronwall Gradients and the

second-order finite difference approximation for the gradient of f with respect to parameters.

(Right) the absolute error over time between Gronwall Gradients and the second-order finite

difference approximation for the gradient of f with respect to parameters for h = 10−3 .

Figure 2.7: The wallclock time in computing M second-order finite difference approxima-

tions to the gradient versus integrating the augmented system in (2.35), averaged over 15

runs.

Given that, in this case, Gronwall Gradients are less expensive and comparably accurate

to a second order finite difference method, we integrate the augmented system (2.35) to

obtain f (t, p) and ∇f (t, p).

We compute the first eigenvector of C from t0 = 0.0 to tF = 35.0 at 8000 points in


39

time by independently computing the active subspace at each time step. The time-series of

this eigenvector (with signs normalized according to (2.28)) will be used as the “truth” to

which we will compare all future approximations. For the following experiments we begin our

approximations from t0 = 0.001 because the initial condition of ∇f for Gronwall Gradients

is ∇f = 0.

2.3.3.1 Numerical Experiments

We again test the accuracy of DMD and SINDy in recovering the dynamic active

subspace as a function of (i) the proportion of time the snapshots represent, and (ii) the

number of samples in the data matrices.

In the following experiments we construct the data matrices XDM D and XSIN Dy with

snapshots from proportions of the data from 0.1 to 1.0. For each proportion we construct

the data matrices from 10, 20, 30, 40, 50, 75, 100, 200, 500, 1000, 2000, and 5000 snapshots.

To determine the influence of α in the accuracy of the SINDy approximation of the dynamic

active subspace, we test the absolute relative error and wall clock time for an approximation

with 30 snapshots from 60% of time for varying the α parameter in scikit-learn’s lasso

function.

Figure 2.8: The influence of varying α on the (left) cost and (right) absolute error of the

approximation of the dynamic active subspace with SINDy.


40

We determine that for these experiments α = 10−7 provides the most accurate approx-

imation with not significantly more cost (see Figure 2.8). Thus in the implementation of

SINDy we use scikit-learn’s lasso function with α = 10−7 .

In order to construct a data matrix XSIN Dy with N snapshots at times t0 , t1 , · · · , tN ,

we compute an additional N eigenvectors at t0 + h, t1 + h, · · · , tN + h for h = 10−6 with

which to approximate Ẋ. In Figure 2.9, the error for each experiment is computed as the

maximum absolute error between the approximated dynamic active subspace, and the “true”

active subspace computed at all 8000 time steps. The library matrix Θ is constructed with

monomials of up to degree 5, trigonometric combinations of up to degree 3, and exponential

combinations of up to degree 3.

Figure 2.9: The maximum absolute error of the dynamic active subspace approximated

by (left) DMD and (right) SINDy. The data matrices for each algorithm have been created

with N samples from p proportion of the total time.

To understand the error in these approximations, we approximate the dynamic active

subspace with 10 samples from the first 50% of time (Figure 2.10). Recall that although there

are 10 samples in the XSIN Dy data matrix, there were 20 active subspace computations.
41

Figure 2.10: The (left) approximated dynamic active subspace and its (right) absolute

error using (top) DMD and (bottom) SINDy. Solutions have been approximated with 10

samples from the first 50% of time.

2.4 DISCUSSION

Our methodology has shown to be effective in recovering the dynamic active subspace

for (i) the linear case for which we also have the analytical form, and (ii) the nonlinear

enzyme kinetics system.

In the linear case, Figure 2.4 shows that the dynamic active subspace can be recovered

with DMD to within 10−8 from only ten samples from the first 10% of time. Conversely

we see that SINDy is only ever able to recover the dynamic active subspace to within 10−2 .

Given that it is necessary to compute the active subspace at twice as many points in time for

the SINDy approximation, it is clear that, for this case, DMD recovers the dynamic active

subspace for less cost and with more accuracy than SINDy.
42

In the nonlinear enzyme kinetics system Figure 2.9 shows that the dynamic active

subspace can be recovered with SINDy to within 10−3 with 40 samples (80 including those

used to approximate the derivative) from 60% of the total time. Conversely, we see that

the approximation given by DMD is bounded below by 10−1 . This example given in Figure

2.10 allows us to visualize what the error in Figure 2.9 means in the context of approxi-

mating the dynamic active subspace. We see that, qualitatively, both the DMD and SINDy

approximations recover the dynamic active subspace fairly well. However, we see that the

approximation found with SINDy overall captures the dynamic active subspace better than

DMD.

DMD so accurately recovered the dynamic active subspace for the linear case due

to the fact that the true dynamic active subspace is indeed a linear dynamical system.

Furthermore, DMD accurately recovered the dynamic active subspace for the snapshots that

were in some sense post-processed when multiplied by the square root of the time-dependent

first eigenvalue. Without this scalar multiplier, the snapshots in DMD would not be from

a linear dynamical system. We only knew that this was, in some sense, the correct scalar

multiplier because of the analysis in Section 2.2. In all other cases, this insight into what

the dynamic active subspace should be or look like does not exist.

SINDy, however, has accurately recovered the dynamic active subspace for the non-

linear case. In this case, it was not necessary to have any prior knowledge of the dynamic

active subspace in order for the methodology to recover the dynamics both qualitatively and

quantitatively well from very little data. In future applications of this work to analyze and

understand other systems, it is recommended that one use the SINDy approach to estimate

the dynamic active subspace. In these applications it has shown to provide approximations

to the one-dimensional dynamic active subspace from little computation and to within 10−3 .

To fully assess the usefulness of this methodology on other systems, however, we must either

find a mathematical guarantee of its ability to approximate the dynamics, or test it on bigger

and more complex systems than those tested here.


43

2.5 CONCLUSIONS AND FUTURE WORK

In this thesis we have presented a novel approach for uncertainty quantification and

sensitivity analysis in time-dependent systems. In the above sections we have discussed (i)

the ideas and algorithms in finding the active subspace of a function, (ii) the use of Gronwall

Gradients in approximating the gradient of a time-dependent function, and (iii) the methods

of dynamic mode decomposition and sparse identification for nonlinear dynamical systems

in recovering and approximating dynamical systems from data.

We have proven that, in the case where the QoI is a function of a linear dynamical

system dependent on its initial conditions, or of an inhomogeneous linear dynamical sys-

tem dependent on the inhomogeneous part of the system, there exists a dynamical system

describing the dynamic active subspace.

For other time-dependent QoI, we have developed a methodology for approximating

the dynamic active subspace given snapshots of the active subspace. We have approximated

the dynamic active subspace using this methodology for (i) a linear harmonic oscillator

and (ii) the enzyme kinetics system. We have quantitatively and qualitatively compared

the recovered dynamic active subspace obtained with DMD and SINDy. We quantitatively

assessed the tradeoff between the proportion of time represented in the data snapshots, and

the number of snapshots, used in DMD and SINDy, and showed that for the linear harmonic

oscillator, DMD was able to recover the dynamics with more accuracy and for less cost.

Conversely, in the enzyme kinetics example, SINDy was able to recover the dynamics with

more accuracy and for less cost.

The insight given from a time-dependent sensitivity metric allows for an in-depth un-

derstanding of the underlying system. Furthermore, the parameters in biological systems

are often approximated with uncertainty resulting from measurement or observation error or

parameter identifiability problems. Our methodology provides a means by which to identify

the linear combination of parameters that will change the function the most on average, and
44

how the linear combination changes in time.

There is great potential for future work in line with this research. By properly ensuring

orthogonality between states, the methodology presented in this work could be extended to

recover the k-dimensional dynamic active subspace given by the first k eigenvectors. It is pos-

sible that the analytical form for the dynamical system of the dynamic active subspace exists

for other time-dependent systems; further research in the overlap with differential geome-

try, classical dynamical systems analysis, and the study of manifolds could yield promising

results in the study of dynamic active subspaces.

The concept of using data-driven recovery techniques (such as DMD and SINDy) to

approximate a time-dependent sensitivity metric given snapshots of the sensitivity metric is,

to the best of our knowledge, novel. Furthermore, the methodology presented in this work

can be extended to develop time-dependent insight for other global sensitivity metrics. This

intellectual and technical contribution of the novel ideas and methodology developed in this

work lend to the advancement of uncertainty quantification and sensitivity analysis for time-

dependent functions, with applications ranging from epidemiological studies, to chemical

kinetics, to biological systems.


BIBLIOGRAPHY

[1] N. T. J. Bailey and J. Duppenthaler, Sensitivity analysis in the modelling of


infectious disease dynamics, Journal of Mathematical Biology, 10 (1980), pp. 113–131.

[2] S. M. Blower and H. Dowlatabadi, Sensitivity and uncertainty analysis of


complex models of disease transmission: An hiv model, as an example, International
Statistical Review / Revue Internationale de Statistique, 62 (1994), pp. 229–243.

[3] N. Borisov, E. Aksamitiene, A. Kiyatkin, S. Legewie, J. Berkhout,


T. Maiwald, N. Kaimachnikov, J. Timmer, J. Hoek, and B. Kholodenko,
Systems-level interactions between insulin–egf networks amplify mitogenic signaling,
Molecular Systems Biology, 5 (2009), pp. 1–15.

[4] S. L. Brunton, J. L. Proctor, and J. N. Kutz, Discovering governing equations


from data by sparse identification of nonlinear dynamical systems, Proceedings of the
National Academy of Science, 113 (2016), pp. 3932–3937.

[5] E. Candes and J. Romberg, l1-magic: A collection of MATLAB routines for solving
the convex optimization programs central to compressive sampling, www. acm. caltech.
edu/l1magic, (2006).

[6] E. J. Candes and T. Tao, Decoding by linear programming, IEEE Transactions on


Information Theory, 51 (2005), pp. 4203–4215.

[7] H. Caswell, Sensitivity analysis of transient population dynamics, Ecology Letters,


10 (2007), pp. 1–15.

[8] R. Chartrand, Numerical differentiation of noisy, nonsmooth data, International


Scholarly Research Network Applied Mathematics, 2011 (2011), pp. 1–11.

[9] S. S. Chen, D. L. Donoho, and M. A. Saunders, Atomic decomposition by basis


pursuit, SIAM Journal on Scientific Computing, (1999), pp. 33–61.

[10] N. Chitnis, J. M. Cushing, and J. M. Hyman, Bifurcation analysis of a


mathematical model for malaria transmission, SIAM Journal on Applied Mathemat-
ics, 67 (2006), pp. 24–45.
46

[11] P. G. Constantine, Active Subspaces: Emerging Ideas for Dimension Reduction in


Parameter Studies (SIAM Spotlights), SIAM-Society for Industrial and Applied Math-
ematics, 2015.

[12] P. G. Constantine, Z. del Rosario, and G. Iaccarino, Many physical laws are
ridge functions, arXiv:1605.07974 [math.NA], (2016), pp. 1–20.

[13] P. G. Constantine and P. Diaz, Global sensitivity metrics from active subspaces,
Reliability Engineering & System Safety, 162 (2017), pp. 1–13.

[14] P. G. Constantine and A. Doostan, Time-dependent global sensitivity analysis


with active subspaces for a lithium ion battery model, Statistical Analysis and Data
Mining, 10 (2017), pp. 243–262.

[15] R. Cooke, Experts in Uncertainty: Opinion and Subjective Probability in Science,


Oxford University Press, 1991.

[16] J. Crutchfield and B. McNamara, Equations of motion from a data series, Com-
plex Systems, 1 (1987), pp. 417–452.

[17] R. Cukier, C. Fortuin, K. Shuler, A. G. Petschek, and J. Schailby, Study


of the sensitivity of the coupled reaction systems to uncertainties in rate coefficients: I.
theory, Journal of Chemical Physics, 59 (1973), pp. 3873–3878.

[18] G. Dieter, Engineering design: a materials and processing approach, McGraw-Hill,


2 ed., 1991.

[19] M. Elad and M. Aharon, Image denoising via sparse and redundant representations
over learned dictionaries, IEEE Transactions on Image Processing, (2006), pp. 3736–
3745.

[20] H. Enderling and M. Chaplain, Mathematical modeling of tumor growth and


treatment, Current Pharmaceutical Design, 20 (2014), pp. 1–7.

[21] T. Gronwall, Note on the derivatives with respect to a parameter of the solutions of
a system of differential equations, Annals of Mathematics, 20 (1919), pp. 292–296.

[22] T. J. Hastie, R. J. Tibshirani, and J. H. Friedman,


The elements of statistical learning, Springer Series in Statistics, New York, NY,
2 ed., 2009.

[23] B. Ioos and P. Lemaı̂tre, A review on global sensitivity analysis methods, 2015.

[24] I. Jolliffe, Principal Component Analysis, Springer Series in Statistics, New York,
1986.

[25] D. Jones, M. Plank, and B. Sleeman, Differential Equations and Mathematical


Biology, Chapman and Hall/CRC, 2 ed., 2009.
47

[26] J. Keener and J. Sneyd, Mathematical Physiology (Interdisciplinary Applied


Mathematics), Springer, 1998.

[27] J. Kim and H. Park, Fast active-set-type algorithms for l1- regularized linear
regression, Proceedings of the International Conference on Artificial Intelligence and
Statistics, 13 (2010), pp. 397–404.

[28] A. Kiparissides, S. S. Kucherenko, A. Mantalaris, and E. N. Pistikopou-


los, Global sensitivity analysis challenges in biological systems modeling, Industrial
and Engineering Chemistry Research, 48 (2009), pp. 7168–7180.

[29] P. Krishnaiah, Analysis of Variance, Elsevier: New York, 1981.

[30] J. N. Kutz, S. L. Brunton, B. W. Brunton, and J. L. Proctor, Dynamic


Mode Decomposition: Data-Driven Modeling of Complex Systems, SIAM-Society for
Industrial and Applied Mathematics, 2016.

[31] V. Kuznetsov, I. Makalkin, M. Taylor, and A. Perelson, Nonlinear dynamics


of immunogenic tumors: Parameter estimation and global bifurcation analysis, Bulletin
of Mathematical Biology, 56 (1994), pp. 295–321.

[32] R. LeVeque, Finite Difference Methods for Ordinary and Partial Differential
Equations: Steady-State and Time-Dependent Problems (Classics in Applied
Mathematics), SIAM, Society for Industrial and Applied Mathematics, 2007.

[33] T. Loudon and S. Pankavich, Mathematical analysis and dynamic active subspaces
for a long term model of HIV, Mathematical Biosciences & Engineering, 14 (2016),
pp. 709–733.

[34] S. Marino, I. B. Hogue, C. J. Ray, and D. E. Kirschner, A methodology for


performing global uncertainty and sensitivity analysis in systems biology, Journal of
Theoretical Biology, 254 (2008), pp. 178–196.

[35] M. McKay, R. Beckman, and W. Conover, A comparison of three methods for


selecting values of input variables in the analysis of output from a computer code,
Technometrics, 21 (1979), pp. 239–245.

[36] M. Morris, Factorial sampling plans for preliminary computational experiments, Tech-
nometrics, 33 (1991), pp. 161–174.

[37] J. Newman and W. Tiedemann, Porous-electrode theory with battery applications,


American Institute of Chemical Engineers Journal, 21 (1975), pp. 25–41.

[38] S. Pankavich and D. Shutt, An in-host model of HIV incorporating latent infection
and viral mutation, arXiv:1508.07616 [q-bio.PE], (2015).

[39] K. Pearson, Notes on regression and inheritance in the case of two parents, Proceed-
ings of the Royal Society of London, 58 (1895), pp. 240–242.
48

[40] R. Sachs, L. Hlatky, and P. Hahnfeldt, Simple ode models of tumor growth
and anti-angiogenic or radiation treatment, Mathematical and Computer Modeling, 33
(2001), pp. 1297–1305.

[41] A. Saltelli, D. Gatelli, F. Campolongo, J. Cariboni, M. Ratto,


M. Saisana, and T. Andres, Global Sensitivity Analysis: The Primer, John Wi-
ley & Sons, 2008.

[42] H. Schaeffer, S. Osher, R. Caflisch, and C. Hauck, Sparse dynamics for


partial differential equations, Proceedings of the National Academy of Sciences, 110
(2013), pp. 6634–6639.

[43] M. Schilling, T. Maiwald, S. Hengl, D. Winter, C. Kreutz, W. Kolch,


W. Lehmann, J. Timmer, and U. Klingmüller, Theoretical and experimental
analysis links isoform-specific erk signalling to cell fate decisions, Molecular Systems
Biology, 5 (2009), pp. 1–18.

[44] P. Schmid, Dynamic mode decomposition of numerical and experimental data, Journal
of Fluid Mechanics, 656 (2010), pp. 5–28.

[45] M. Schmidt and H. Lipson, Distilling free-form natural laws from experimental data,
Science, 324 (2009), pp. 81–85.

[46] J. Schmitz, N. V. Riel, K. Nicolay, P. Hilbers, and J. Jeneson, Silencing


of glycolysis in muscle: experimental observation and numerical analysis, Experimental
Physiology, 95 (2010), pp. 380–397.

[47] S. Tavener, M. Mikucki, S. G. Field, and M. F. Antolin, Transient sensitivity


analysis for nonlinear population models, Methods in Ecology and Evolution, 2 (2011),
pp. 560–575.

[48] R. Tibshirani, Regression shrinkage and selection via the lasso, Journal of the Royal
Statistical Society, (1996), pp. 267–288.

[49] G. Tran and R. Ward, Exact recovery of chaotic systems from highly corrupted
data, arXiv:1607.01067 [math.DS], (2016).

[50] J. Tu, C. Rowley, D. Luchtenburg, S. Brunton, and J. Kutz, On dynamic


mode decomposition: Theory and applications, Journal of Computational Dynamics, 1
(2014), pp. 391–421.

[51] T. Turányi, Sensitivity analysis of complex kinetic systems. tools and applications,
Journal of Mathematical Chemistry, 5 (1990), pp. 203–248.

[52] H. Voss, P. Kolodner, M. Abel, and J. Kurths, Amplitude equations


from spatiotemporal binary-fluid convection data, Physical review letters, 83 (1999),
pp. 3422–3425.

You might also like