S To Chas Tic Multi Scale
S To Chas Tic Multi Scale
S To Chas Tic Multi Scale
A Dissertation
by
Baskar Ganapathysubramanian
August 2008
c 2008 Baskar Ganapathysubramanian
The author was born in Madurai, India in October, 1981. After completing his
high school, the author was admitted into the Mechanical Engineering program
at IIT, Madras in 1999, from where he recieved his Bachelor’s degree in engineer-
ing in June, 2003. The author entered the doctoral program at the Sibley School
of Mechanical and Aerospace Engineering, Cornell University in August, 2003.
iii
The thesis is dedicated to my parents and brother for their constant support
and encouragement towards academic pursuits during my school years.
”We have not succeeded in answering all our problems. The answers we have found
only serve to raise a whole set of new questions. In some ways we feel we are as
confused as ever, but we believe we are confused on a higher level and about more
important things.” - Posted outside the mathematics reading room,
Tromsø University
iv
ACKNOWLEDGEMENTS
I would like to thank my thesis advisor, Professor Nicholas Zabaras, for his con-
stant support and guidance over the last 3 years. I would also like to thank
Professors Tomas Arias and Michael Thompson for serving on my special com-
mittee and for their encouragement and suggestions at various times during the
course of this work.
The financial support for this project was provided by NASA (grant NAG8-
1671) and the National Science Foundation (grant DMI-0113295). I would like
to thank the Sibley School of Mechanical and Aerospace Engineering for having
supported me through a teaching assistantship for part of my study at Cornell.
The computing for this project was supported by the Cornell Theory Center dur-
ing 2003-2008. The parallel simulators were developed based on open source
and Dmitry. You have made life beyond the lab fun and fullfilling!
v
TABLE OF CONTENTS
1 Introduction 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Some definitions and mathematical preliminaries . . . . . . . . . 12
vi
3 Data-driven approaches to constructing stochastic input models 71
3.1 Representing input randomness . . . . . . . . . . . . . . . . . . . . 71
3.1.1 Input parameters as random variables . . . . . . . . . . . . 71
3.1.2 Random field as a collection of random variables . . . . . 72
3.1.3 The Karhunen-Loève expansion . . . . . . . . . . . . . . . 73
3.1.4 Model reduction as a tool for constructing input models . 75
3.2 Problem definition . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.3 Nonlinear model reduction: Its necessity and some basic ideas . . 82
3.3.1 Linear model reduction: where does it fail? . . . . . . . . . 83
3.3.2 Non-linear model reduction: preliminary ideas . . . . . . 84
3.4 Mathematical formulation . . . . . . . . . . . . . . . . . . . . . . . 88
3.4.1 Some definitions and the compactness of the manifold M 89
3.4.2 Estimating the pair-wise geodesic distances: graph ap-
proximations . . . . . . . . . . . . . . . . . . . . . . . . . . 92
3.4.3 Estimating the optimal dimension, d, of the low-
dimensional representation . . . . . . . . . . . . . . . . . . 94
3.4.4 Constructing the low-dimensional parametrization, A . . 97
3.4.5 Construction of a non-parametric mapping F −1 : A → MS2 99
3.4.6 The low-dimensional stochastic input model F −1 : A → MS 2 107
3.5 Numerical implementation . . . . . . . . . . . . . . . . . . . . . . 107
3.5.1 Microstructure reconstruction: creating the samples xi . . 108
3.5.2 Constructing the low-dimensional region A . . . . . . . . 108
3.5.3 Utilizing this reduced-order representation: Stochastic
collocation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
3.6 Illustrative example . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
3.6.1 Two-phase microstructures . . . . . . . . . . . . . . . . . . 113
3.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
Bibliography 174
vii
LIST OF TABLES
viii
LIST OF FIGURES
ix
2.17 Realizations of the problem: Temperature, u velocity and v ve-
locity components. . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.18 Mean values of the variables: Top left: Temperature, Top right: u
velocity component, Bottom left: v velocity component, Bottom
right: pressure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
2.19 Mean velocity vectors and streamlines. . . . . . . . . . . . . . . . 56
2.20 Standard deviation of the variables: Top left: Temperature, Top
right: u velocity component, Bottom left: v velocity component,
Bottom right: pressure. . . . . . . . . . . . . . . . . . . . . . . . . 56
2.21 Probability distribution functions for the dependent variables
at point (0.25, 0). Top row: Temperature and pressure, Bottom
row: u and v velocity components. . . . . . . . . . . . . . . . . . . 58
2.22 Mode shift: Left: Temperature, Right: v velocity component. . . . 58
2.23 Mode shift: Left: Temperature, Right: v velocity component. . . . 59
2.24 Extreme realizations of temperature. . . . . . . . . . . . . . . . . 59
2.25 Mean and standard deviations of temperature. . . . . . . . . . . 60
2.26 Mode shift: Left: Temperature, Right: v velocity component. . . 60
2.27 Left: ACF of the surface roughness, Right: Eigen-spectrum of the
correlation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.28 Schematic of natural convection with wall surface roughness
computed from experiments. . . . . . . . . . . . . . . . . . . . . . 61
2.29 Realizations of temperature. . . . . . . . . . . . . . . . . . . . . . 62
2.30 Mean values of temperature and v velocity component. . . . . . 63
2.31 Mean velocity vectors and streamlines. . . . . . . . . . . . . . . . 64
2.32 Standard deviations of temperature and v velocity component. . 64
2.33 Schematic of convection in a medium with random porosity. . . 65
2.34 Fontainebleau sandstone: Correlation function and cross-
sectional image. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
2.35 Eigen spectrum of the correlation kernel. . . . . . . . . . . . . . . 66
2.36 Mean values of the variables: Top left: Temperature, Top right: u
velocity component, Bottom left: v velocity component, Bottom
right: pressure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
2.37 Mean velocity vectors and streamlines. . . . . . . . . . . . . . . . 67
2.38 Standard deviation of the variables: Top left: Temperature, Top
right: u velocity component, Bottom left: v velocity component,
Bottom right: pressure. . . . . . . . . . . . . . . . . . . . . . . . . 68
x
3.3 A set of points lying on a spiral in 3D space. The global coordi-
nates of any point on the spiral is represented as a 3-tuple. The
figure on the left (a) shows the reduced-order model resulting
from a linear PCA based reduction. The figure on the right (b)
depicts a non-linear strategy that results in an accurate repre-
sentation of the 3D spiral (data taken from [78]) that works by
“unravelling and smoothing” the 3D spiral into a 2D sheet. . . . 85
3.4 Figure (a) on the left shows images of the same object (from [78])
taken at various poses (left-right and up-down) while Figure (b)
on the right shows various two-phase microstructures that sat-
isfy a specific volume-fraction and two-point correlation. . . . . . 86
3.5 The figure above illustrates the simplest possible mapping be-
tween the low-dimensional region A and the high-dimensional
microstructural space MS2 . Given an arbitrary point ξ ∈ A, find
the point yk closest to ξ from the sampled points {yi }, i = 1, . . . , N.
Assign the image value of yk i.e. xk to the image of ξ. . . . . . . . 101
3.6 The figure above illustrates a local linear (k-neighbor) interpo-
lation mapping between the low-dimensional region A and the
high-dimensional microstructural space MS2 . . . . . . . . . . . . 102
3.7 Simple estimate for the interpolation error in Method 2: Let r de-
note the local radius of curvature of the function near the point
xexact (the filled square). We approximate the curve by a linear
patch, resulting in some error. This error is the distance between
the approximate linear image, x (the unfilled square) and the ac-
tual point, xexact (the filled square). This distance can be approx-
imated from simple geometry as a function of r and the local
geodesic distance between the points. . . . . . . . . . . . . . . . . 103
3.8 The local linear interpolation method can be made exact by sim-
ply projecting the image obtained after interpolation onto the
manifold. This is illustrated in the figure above, where the bot-
ted blue line represents the linear interpolation and the curved
line represents the projection operator that constructs the image
lying on the manifold. . . . . . . . . . . . . . . . . . . . . . . . . . 105
3.9 Sample illustration of the projection step: The figure on the left is
a microstructure after interpolation. Projecting it onto the man-
ifold yields the microstructure on the right. There is negligible
change in the three point correlation between the micorstructures. 105
3.10 Experimental image of a two-phase composite (from [66]). . . . . 114
3.11 The two-point correlation function. . . . . . . . . . . . . . . . . . 114
3.12 Comparison of the two-point correlation function from experi-
ments and from the GRF. . . . . . . . . . . . . . . . . . . . . . . . 116
3.13 One instance (realization) of the two-phase microstructure. . . . 116
3.14 Plot of the length functional of the MST of the graph G for vari-
ous sample sizes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
xi
Pd
i λ
3.15 (Left) The cumulative eigenspectrum of the data, Pi=1 N
λ
. (Right)
i=1 i
The residual variance for different dimensionalities of the region
A, computed from MDS. . . . . . . . . . . . . . . . . . . . . . . . 118
3.16 Reduction in the interpolation error with increasing number of
collocation points. . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
3.17 Steady-state mean temperature: (a) temperature contour, (b-d)
temperature iso-surfaces, and (e-g) temperature slices. . . . . . . 121
3.18 Standard deviation of temperature: (a) standard deviation con-
tours, (b) standard deviation iso-surfaces, (c) temperature PDF
at a point, and (d-f) standard deviation slices. . . . . . . . . . . . 122
xii
4.16 Convergence of the stochastic solution (mean and standard de-
viation) with increasing number of collocation points. (left)
Stochastic coarse-scale pressure, (right) Stochastic coarse-scale x-
direction flux. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
4.17 Mean contours of the stochastic coarse-scale solution. (left)
Coarse-scale pressure, (right) Coarse-scale x-direction flux. . . . . 152
4.18 Streamlines drawn on the mean stochastic coarse-scale velocity. . 153
4.19 Standard deviation of the stochastic coarse-scale solution. (left)
Coarse-scale pressure, (right) Coarse-scale x-direction flux. . . . . 153
4.20 Probability distribution functions of the coarse-scale flux (x-
direction) for two points in the domain. . . . . . . . . . . . . . . . 154
4.21 The refinement of the adaptive sparse grid. Each row of points
represents the refinement at a particular level of interpolation.
The interpolation starts at the bottom. . . . . . . . . . . . . . . . . 155
4.22 Probability distribution functions of the coarse-scale flux (x-
direction) for two points in the domain. . . . . . . . . . . . . . . . 156
4.23 (left and center) Two realizations of the coarse-scale flux (x-
direction) for a 1% perturbation in the x-location of the perme-
ability barrier, (right) Difference between the two solutions. . . . 156
4.24 Schematic of the domain. The permeability is uncertain in part
of the domain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
4.25 The fine-scale log permeability distribution in the domain. The
axes are scaled (1 unit = 50 ft). . . . . . . . . . . . . . . . . . . . . 158
4.26 The eigen-spectrum of the data set. The first 6 eigen-modes rep-
resent most of the variation that the input data set exhibits. . . . 159
4.27 Convergence of the stochastic solution (mean and standard de-
viation) with increasing number of collocation points. (left)
Stochastic coarse-scale pressure, (right) Stochastic coarse-scale x-
direction flux. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
4.28 Mean contours of the stochastic coarse-scale solution. (left)
Coarse-scale pressure, (right) Coarse-scale x-direction flux. . . . . 160
4.29 Standard deviation of the stochastic coarse-scale solution. (left)
Coarse-scale pressure, (right) Coarse-scale x-direction flux. . . . . 160
4.30 Probability distribution functions of the coarse-scale pressure
(left) and flux (x-direction) (right) for a point upstream of the
uncertainty. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
4.31 Probability distribution functions of the coarse-scale pressure
(left) and flux (x-direction) (right) for a point in the uncertain
domain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
4.32 Plot of the length functional of the MST of the graph G for vari-
ous sample sizes. The intrinsic dimension is related to the slope
of the graph. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
4.33 Mean contours of the stochastic coarse-scale solution. (left)
Coarse-scale pressure, (right) Coarse-scale x-direction flux. . . . . 163
xiii
4.34 Standard deviation of the stochastic coarse-scale solution. (left)
Coarse-scale pressure, (right) Coarse-scale x-direction flux. . . . . 163
4.35 Probability distribution functions of the coarse-scale pressure
(left) and flux (x-direction) (right) for a point upstream of the
uncertainty. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
4.36 Probability distribution functions of the coarse-scale pressure
(left) and flux (x-direction) (right) for a point in the uncertain
domain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
xiv
CHAPTER 1
INTRODUCTION
1.1 Introduction
With the rapid advances in computational power and easier access to high-
are composed of) are affected to a significant extent by the inherent (usually un-
resolvable) uncertainties in material properties, fluctuations in operating condi-
tions, and system characteristics. The scale at which these phenomena occur can
range from the micro-scale (MEMS devices), macro-scale (devices/components
effects of these input uncertainties into the system and understand how they
propagate and alter the final solution.
1
Another class of problems that illustrate the necessity of incorporating the
effects of uncertainty (in the lower scales) is in the design and analysis of com-
ponents made of polycrystalline, functionally graded and other heterogeneous
materials. Most engineering devices involve the use of such multi-component
neous media. Nevertheless, there has been no dedicated effort in the engineer-
ing community for incorporating the effect of such microstructural uncertainty
in process or device design even though it is well known that most performance
related failures of critical components are directly related with rare events (tails
2
zle flaps, gears, etc.) that are composed of polycrystals and/or functionally
graded materials, hydrodynamic transport through porous media and chem-
ical point operations (e.g. filtration, packed columns. fluidization, sedimen-
tation and distillation operations). Other sources of uncertainty include er-
tal studies implies that these input parameters have some inherent uncertain-
ties. The uncertainties in the physical model may also be due to inaccuracies
or indeterminacies of the initial conditions. This may be due to variation in
experimental data or due to variability in the operating environment. Uncer-
tainties may also creep in due to issues in describing the physical system. This
may be caused by errors in geometry, roughness, and variable boundary condi-
tions. Uncertainties could also occur due to the mathematical representation of
the physical system. Examples include errors due to approximate mathematical
Ingredients for stochastic analysis: There are two key ingredients for con-
structing a framework to analyze the effects of various sources of uncertainty in
a system.
3
ations and thermal perturbations, or more generally as noise. We denote
these as the input uncertainties.
ingful conclusions from the resulting solutions. For instance, trying to ac-
count for the variability in room temperature (in a simple heat transfer in
a room problem) using an input model that allows the room temperature
to vary from -100 OC to +100 OC is a valid albeit probably useless model.
works that have been developed in recent years for incorporating the effects of
uncertainties in various systems. The presence of uncertainties can be modelled
in the system through reformulation of the governing equations as stochastic
partial differential equations (SPDEs). Solution techniques for SPDEs can be
other major attraction of these methods are that their convergence rates do not
depend on the number of independent random variables. Further, they are very
easy to implement. The only necessity of computing uncertainties using the
statistical approach is the availability of a working deterministic code. How-
ever, the statistical approach becomes quickly intractable for complex problems
in multiple random dimensions. This is because the number of realizations re-
4
quired to acquire good statistics is usually quite large. Furthermore, the number
of realizations changes with the variance of the input parameters and the trun-
cation errors in this inherently statistical method are hard to estimate. This has
in part been alleviated by improved sampling techniques like Latin hypercube
has been some recent work in reformulating the problem as a set of moment
differential equations (MDE). The closure equations for these MDE’s are then
derived from Taylor expansions of the input uncertainty as well as perturba-
tion and asymptotic expansions. The random domain decomposition method
[33, 34, 35, 36] coupled with the MDE method has been shown to be success-
ful in solving complex fluid flow problems in porous media having large input
variances.
with space and time and a field variable is expanded along the uncertain di-
mension using suitable expansions. The basic idea is to project the dependent
5
variables of the model onto a stochastic space spanned by a set of complete
orthogonal polynomials. These polynomials are functions of a set of random
variables ξ(θ) where θ is a realization of the random event space. Ghanem and
Spanos first used SSFEM in linear elastic problems in [7]. Ghanem and co-
tion problems [13]. In the original work by Wiener, Gaussian random variables
were used with Hermite polynomials. The scheme has been extended to in-
clude other random distributions [37, 38, 39, 40, 41, 42]. Error bounds, and
convergence studies [43, 44, 45, 46, 47] have shown that these methods exhibit
Like the statistical methods, the SSFEM approach also reduces the problem
to a set of deterministic equations. But unlike the statistical methods, the re-
sulting deterministic equations are often coupled. This coupled nature of the
the number of expansion terms. This problem is especially acute when dealing
with realistic problems in fluid mechanics. The number of random dimensions
needed to represent surface fluctuations (roughness) or permeability variations
is relatively high (N ≥ 8). To solve these problems in high-dimensional spaces
and to allow non-smooth variations of the solution in the random space, there
6
have been recent efforts to couple the fast convergence of the Galerkin methods
with the decoupled nature of Monte-Carlo sampling [48]. Babuska et al. [15]
proposed a methodology wherein the Galerkin approximation is used to model
the physical space while a collocation scheme is used to sample the random
space. A tensor product rule was used to interpolate the variables in stochas-
tic space using products of one-dimensional interpolation functions (double or-
thogonal polynomials) based on Gauss points. Though this scheme effectively
decoupled the deterministic system of equations, the number of realizations re-
different fields [49, 50, 52, 53, 54]. Using this method, interpolation schemes can
be constructed with orders of magnitude reduction in the number of sampled
points to give the same level of approximation (up to a logarithmic factor) as
interpolation on a uniform grid.
In the standard sparse grid collocation approach, all the dimensions are
treated equally. In many problems usually encountered, not all the dimensions
are important. That is, the solution varies much more smoothly in some par-
ticular dimension than in others. This brings up the possibility of reduction in
the computational effort required to solve the stochastic problem by weighting
function evaluations.
7
In Chapter. 2, an adaptive extension to the sparse grid collocation method
is developed that can automatically detect the important dimensions and bias
the sparse sampling towards those dimensions. This results in substantial sav-
ings in the computational effort required to solve large dimensional problems.
the problem assuming that the property and topological variations are random
fields satisfying the experimental correlations. To perform any such analysis,
one must first construct models of these variations to be used as inputs in the
subsequent uncertainty analysis. The analysis of the effect of such uncertainties
8
dimensional, continuous mapping) that encodes and quantifies the variation of
material topology and properties in a mathematically rigorous way.
There have been very few previous investigations into developing stochas-
tic input representations. The recent work in [19] looks at developing proba-
bilistic models of random coefficients in SPDEs using a maximum likelihood
all of these techniques for constructing input models are based on the concept
of transforming experimental data and statistics into probability distributions
of the property. These techniques are usually highly application specific, re-
quire expert knowledge in assigning probabilities and invariably require some
9
absence of the requirement of any expert knowledge. In addition, the general-
ity of the mathematical developments results in a framework that can construct
input models for any property variability, seamlessly meshing with any recon-
struction algorithm and software that produce plausible data sets.
bed data variations into a low-dimensional manifold that can serve as the in-
put model for subsequent analysis. The significant features of this framework
are as follows: A completely general methodology to generate viable, realistic,
reduced-order stochastic input models based on experimental data is proposed.
To the best knowledge of the author, this is the first time that such a broad frame-
work for generating models for any property variation has been developed. We
provide a rigorous mathematical basis for the proposed framework. Further-
more, error estimates and strict convergence estimates of the reduced-order
model are provided. The basic model reduction ideas developed are not just
limited to generation of viable stochastic input models of property variations.
This framework has direct applicability to problems where working in high-
dimensional spaces is computationally intractable, for instance, in visualization
10
Complete response evaluation involving full (fine-scale) spatial and temporal
resolution simulations of multiscale systems is extremely expensive. Com-
putational techniques have been developed that solve for a coarse-scale so-
lution by defining an appropriate coarse-scale problem that captures the ef-
fect of the fine-scales [94]. The more popular techniques developed for such
upscaling fall under the category of multiscale methods viz. the variational
multiscale (VMS) method (operator upscaling) [95, 96, 97, 98, 99, 100], the
heterogenous multiscale method [101, 102] and the multiscale finite element
method [103, 104, 105, 106, 107]. Further related techniques are presented
in [108, 109]. These computationally scalable techniques have resulted in the
development of black box simulators that have been used with significant suc-
cess in the solution of large scale transport problems in complex geometries.
The multiscale analysis of such systems inherently assumes that the com-
The basic idea to solve the stochastic multiscale set of equations is to ex-
tend deterministic multiscale methods to their stochastic analogues. Spectral
11
strategies to pose and solve stochastic multiscale problems have been investi-
gated by Xu [112] and Asokan and Zabaras [42]. Collocation based strategies
have recently been developed in [20]. The key is to define appropriate ways to
link the fine-scale (subgrid) stochastic variation with the coarse-scale stochastic
In the final part of the thesis, we utilize the computational strategies de-
veloped to analyzing flow through random heterogeneous media given lim-
ited statistical information about the multiscale permeability variation. We
link stochastic analysis and multiscale methods to investigate this problem.
For the sake of clarity, let us denote the variable’s of interest as θ, u, p. Denote
by D, a region in Rnsd , the domain of interest. nsd = 1, 2, 3 is the number of
space dimensions over which D is defined. Denote by T ≡ [0, tmax ] the time do-
12
main of interest. Denote the boundary of D as ∂D. Denote the set of governing
(deterministic) differential equations defining the evolution of θ as
constitutive relations that the physical system may require. For instance in the
context of thermal problems, {q, α} could represent the heat flux and the thermal
conductivity of the system while in the context of elastic deformation problems,
{q, α} could represent the displacement boundary conditions and the elastic ten-
Suppose, without loss of generality, that the input quantity α is in fact un-
certain. Assume also, for the sake of clarity, that α is a scalar. Denote by Ωα the
space of possible values that α can take. Now, Ωα can be a set of discrete points
{α1 , . . . , αN } or it can be a continuous range of values α ∈ [a, b]. To rigorously
define this notion of variability, the notion of a σ-algebra is introduced. A σ-
to restrict our attention to a smaller, and usually more useful, collection of sub-
sets of a given set. The main use of σ-algebras is in the definition of measures on
13
Ωα . This is one of the cornerstones of probability theory (the interested reader
is referred to [17] for a gentle introduction to σ-algebras in probability theory).
The measure that we are usually interested in is the probability measure, P, i.e.
the probability of a particular subset in the σ-algebra occurring. That is the
of α(ωα ). Note that α(ωα ) is still an abstract quantity, that lies in an abstract
probability space (Ωα , F , P). In Chapter. 3, we will look at different ways of
representing this random process for straightforward utility in a computational
framework. This probabilistic definition of α(ωα ) necessitates a redefinition of
the governing equation. This can be done naturally as follows: define the evo-
lution of (u, p, θ) as
Now, the dependent variable θ depends not only on (x, t) but also on the space
Ωα . The solution of the above set of equations results in functions u : D × T ×
Ωα → Rnsd , p : D × T × Ωα → R, θ : D × T × Ωα → R
14
CHAPTER 2
SOLVING STOCHASTIC PARTIAL DIFFERENTIAL EQUATIONS: A
SPARSE GRID COLLOCATION APPROACH
defined over an abstract probability space (Ωα , F , P). The solution methodol-
ogy to solve of the above set of coupled differential equations is to first re-
duce the complexity of the problem by reducing the probability space into a
finite dimensional space. In some cases the random field α(x, ω) can be repre-
sented/described by a finite length random vector [ξ 1 , . . . , ξ N ] : Ω → RN . In
other cases the random field can have a spatial correlation or variation. There is
rich literature on techniques to extract/fit correlations for these random fields
15
Thus we can consider the following:
It is usually assumed that ξ i (ω) are independent random variables with prob-
that the set of Eqs.2.1) can now be written as a set of differential equations in
(N + nsd) dimensions, where N is the dimensionality of the truncated random
space Γ and nsd is the dimensionality of the physical space D:
B(u, p, θ : x, t, ξ) = 0. (2.3)
In this approach, the dependent variables (u, p, θ) are regarded as random pro-
cesses. These random fields are then expanded into a truncated series. One
could use the Karhunen-Loève expansion to represent these fields, but unfortu-
nately one does not a priori know the correlation functions for these variables.
16
This is written as
P
X
κ(x, t, ω) = κi (x, t)Φi (ξ i (ω)), (2.4)
i=0
where κ is any of the dependent variables. Here κi are the deterministic coef-
the input random field. For example, if Gaussian random variables are cho-
sen then the Askey based orthogonal polynomials {Φi } are chosen to be Hermite
polynomials, if ξ i are chosen to be uniform random variables, then {Φi } must
be Legendre polynomials (for a complete description of the GPCE scheme, the
(N + p)!
P+1 = . (2.5)
N!p!
The GPCE is used to expand all the dependent variables in terms of the orthog-
onal polynomials. Substituting these expansions into the governing equation,
where ha, bi is the inner product of the functions a and b over the ensemble
R
(ha, bi = Ω abdx). By using the orthogonality of the polynomial basis, we can
17
obtain a set of (P + 1) coupled equations for each random mode ui , pi , θi . By
utilizing the polynomial chaos expansion followed by the Galerkin projection,
the randomness has been transferred from the dependent variables to the basis
polynomials. The resulting governing equations for the expansion coefficients
are deterministic.
This methodology has been very successful in solving SPDEs [7, 11, 12, 13,
37, 39, 41, 42, 46]. We will restrict our discussion to the disadvantages of this
method so as to motivate the next few sections. The coupled nature of the
expansion coefficients makes the implementation of the GPCE method non-
data processing methodologies are necessary for the efficient solution of higher-
dimensional problems. The major constraint for the utility of this methodology
is the coupled nature of the resulting equations. This resulted in the search for a
method that combined the fast convergence and error estimates of the spectral
Galerkin method with the decoupled nature of the more statistical methods like
the Monte-Carlo methods.
In the spectral Galerkin method described in the previous section, the spatial
domain is approximated using a finite element discretization and the stochastic
18
resentation of the stochastic space causes the coupled nature of the resulting
equations. Another approach is to have a finite element approximation for the
spatial domain and approximate the multidimensional stochastic space using
interpolating functions. The interpolating functions are mutually orthogonal
and the resulting equations are decoupled. This approach is called the colloca-
tion approach, where one computes the deterministic solution at various points
in the stochastic space and then builds an interpolated function that best ap-
proximates the required solution. We now detail the issues involved in con-
structing the stochastic solution using the collocation method. We follow the
spirit of the discussion in Xiu et.al [16] to give an intuitive development of the
method.
for the dependent variables using their values at particular points in the stochas-
tic space. The Galerkin projection is then applied to find that interpolation func-
tion which minimizes the projected error of the approximated system. Denote
by ξ any point in the random space Γ ⊂ RN , by ΠN , the space of all N-variate
19
terpolation polynomials as follows:
M
X
I f (ξ) = f (ξk )Lk (ξ), (2.8)
k=1
using the nodes {ΘN }, the value of the function at any point ξ ∈ Γ is approxi-
mately I f (ξ). The Lagrange interpolated values of (u, p, θ), denoted by (û, p̂, θ̂)
are as follows:
M
X
(û(ξ), p̂(ξ), θ̂(ξ)) = (Iu(ξ), Ip(ξ), Iθ(ξ)) = (u(ξ k ), p(ξk ), θ(ξk ))Lk (ξ). (2.9)
k=1
Remark 1. Relation to the spectral Galerkin method: Babuska and coworkers [44,
45] realized that by choosing appropriate basis functions, the so-called double
orthogonal functions, the spectral stochastic Galerkin formulation can also be
made decoupled. This is specific to the case when the random input is a linear
functions of the input random variables. They also showed that in case the input
uncertainties are multi-linear combinations of the random variables {ξ i (ω)} (as
20
would happen if one uses the K-L expansion to truncate the input random fields
to a finite set of random variables), the solution obtained from the collocation
procedure coincides with the stochastic Galerkin solution [15]. They argue that
the collocation method can be seen as a pseudo-spectral Galerkin method with
21
However, uniform convergence (k f − In ( f )k∞ → 0 as n → ∞) is not guaranteed
for any arbitrary distribution of the nodes. Utilizing a Taylor series expansion,
the interpolation error for a function f ∈ Cn+1 [a, b] at a point x is given by [54]
n
f n+1 (ξ) Y
f (x) − In ( f )(x) = (x − x j ). (2.15)
(n + 1)! j=0
The magnitude of the (n + 1)th derivative could outweigh the product of the
nodes. But there are unique node distributions that exist where uniform con-
vergence with the number of nodes can be proved if f ∈ C1 [a, b].
where Λn (X) = max x∈[a,b] ( ni=0 |Li (x)|) where X is any distribution of the nodes.
P
It is apparent that Λn (X) does not depend on the function f but in fact only
depends on the distribution of the points X. Therefore it is possible to construct
a priori, sets of nodes X with a small Lebesgue constant. One such type of node
distribution is the interpolation based at the Chebyshev extrema [54].
Once the type of support nodes {X} has been chosen and the Lebesgue con-
stant for the node distribution computed, it is straightforward to arrive at er-
22
ror bounds for the polynomial interpolation function IX . For instance, the er-
ror bound for the Chebyshev node based interpolation function of the function
f ∈ Ck is given by [54],[55]
tion
where Iik are the interpolation functions in the ik direction and ξ ijkm is the jm th
point in the kth coordinate. Clearly, the above formula needs m1 ×. . .×mN function
evaluations, at points sampled on a regular grid. In the simplest case of using
only two points in each dimension, the total number of points required for a full-
tensor product interpolation is M = 2N . This number grows very quickly as the
number of dimensions is increased. Thus one has to look at intelligent ways of
sampling points in the regular grid described by the full-tensor product formula
23
2.4 Sparse grid collocation methods
For the univariate case, Gauss points and Chebechev points have least interpola-
tion error (for polynomial approximations). In the case of multivariate interpo-
lation, one feasible methodology that has been used is to construct interpolants
24
Let us denote the set of points used to interpolate the one-dimensional function
by Θ(k) . For instance, Θ(3) could represent the Chebyshev points that interpolate
a third-order polynomial. The Smolyak algorithm constructs the sparse inter-
polant Aq,N (where N is the number of stochastic dimensions and q − N is the or-
of one dimensional functions of order ik with the constraint that the sum total
(|i| = i1 +. . .+iN ) across all dimensions is between q − N +1 and q. The structure of
the algorithm becomes clearer when one considers the incremental interpolant,
∆i given by [49, 52, 53, 54]
U 0 = 0, ∆i = U i − U i−1 . (2.22)
To compute the interpolant Aq,N ( f ) from scratch, one needs to compute the func-
tion at the nodes covered by the sparse grid Hq,N
[
Hq,N = (Θ1(i1 ) × . . . × Θ(i1 N ) ). (2.24)
q−N+1≤|i|≤q
The construction of the algorithm allows one to utilize all the previous re-
sults generated to improve the interpolation (this is immediately obvious from
25
Eq. (2.23)). By choosing appropriate points (like the Chebyshev and Gauss-
Lobatto points) for interpolating the one-dimensional function, one can ensure
that the sets of points Θ(i) are nested (Θ(i) ⊂ Θ(i+1) ). To extend the interpolation
from level i to i + 1, one only has to evaluate the function at the grid points that
are unique to Θ(i+1) , that is, at Θi∆ = Θi \Θi−1 . Thus, to go from an order q − 1 inter-
polation to an order q interpolation in N dimensions, one only needs to evaluate
the function at the differential nodes ∆Hq,N given by
[
∆Hq,N = (Θi∆1 ⊗ . . . ⊗ Θi∆N ). (2.25)
|i|=q
where k is called the level of the Smolyak construction. The interpolation error
using the Smolyak algorithm to construct multidimensional functions (using
the piecewise multilinear basis) is [53, 54]
26
2.4.3 Solution strategy
(uhk (x), phk (x), θkh (x)) = (uh (x, ξk ), ph (x, ξk ), θh (x, ξ k )), (2.28)
such that
B(uh (ξi ), ph (ξ i ), θh (ξi ) : ξ i , x, t) = 0, i = 1, .., M. (2.29)
stochastic space, εI . This error can be written as ε ≤ (C1 ε2D + C2 ε2I )1/2 , where C1
and C2 are constants independent of the discretizations in physical space and
stochastic space.
27
Figure 2.1: Collocation nodes for a two-dimensional problem in a level
4 full (left) and sparse grid (right). Both the grids are con-
structed from the same level of one-dimensional interpolation
functions. The sparse grid offers marginally worse interpo-
lation properties with huge improvements in the number of
points sampled.
In this subsection, we illustrate the sparse grid interpolation method with a few
examples which will also serve as a motivation for the development of the adap-
tive sparse grid methodology.
Grid sizes
Fig. 2.1 shows the two-dimensional interpolation nodes for the sparse grid
along with the two-dimensional tensor product grid based on the same one-
28
Figure 2.2: Comparison of the exact (left) and interpolant (right) functions.
number of discrete points Θ. These function evaluations are usually finite el-
ement simulations of the deterministic problem at each stochastic point Θi . To
illustrate the methodology, we will instead assume that certain analytical func-
2 −̺y2
tions are the inputs at specific nodal points. Consider the function f = e−x ,
lute error max | f − Aq,2 ( f )| is 3 × 10−5 when considering 3329 points. This corre-
sponds to building the sparse interpolant using one-dimensional interpolating
functions up to order 9. To achieve a similar interpolation error using the full-
tensor product grid would require 263169 points.
grid is plotted in Fig. 2.3 for ̺ varying from 1 to 10000. Notice that as the
anisotropy increases, the variation of the function along one direction becomes
29
0
10
-1
10
-2
10
Error
ρ= 1
10
-3
ρ= 10
ρ= 100
ρ= 1000
-4 ρ= 10000
10
-5
10
0 1 2 3 4
10 10 10 10 10
Number of nodes
0.9
0.8
↑ρ
0.7
0.6
ρ =1
0.5
0.4
0.3
ρ =10
0.2
0.1 ρ =100
0
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
Figure 2.4: Slice of the function in the x plane. The arrow shows the direc-
tion of increasing ̺.
progressively steeper (see Fig. 2.4). Thus, more support nodes are required to
accurately interpolate the function. To reach an accuracy of 2 × 10−2 the number
of points required are 65, 145, 321, 705, 1537 and 3329, for ρ = 1, 10, 100, 1000
The critical point to note here is that even though only one dimension of
the function is very steep (as ̺ increases), the sparse grid increases the node
allocation in all directions. This suggests that one can further reduce the number
of function evaluations if we can detect the directions in which the function
30
varies more smoothly than the others.
analyzing mode jumps and equilibrium shifts that occur due to small variations
in the random variables. One extreme case is when there is a sharp (finite) dis-
continuity in the solution across a particular random variable. We are interested
in the performance of the sparse grid collocation method in capturing this dis-
continuity.
1
10
0
10
-1
10
Error
-2
10
-3
10
-4
10
0 1 2 3 4 5
10 10 10 10 10 10
Number of nodes
2 +2sign(y)
Consider a function f = e−x in the domain [−1, 1] × [−1, 1]. Here sign
is the signum function: 1, when y is positive and −1, otherwise. The function f
has a sharp discontinuity along the y = 0 plane. On the other hand, the function
is very smooth in the x direction. That is, the function f (x, c), for a constant c
is an exponentially decaying function. The sparse grid method is used to build
the interpolation for this function. Fig. 2.5 plots the error as a function of the
number of support nodes used by the algorithm. Notice that the error reduction
is sub-linear during the first few interpolation levels. This is because of the error
31
y x y x y x y x
1 1 1 1
0 0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Fig. 2.6 plots the ‘evolution’ of the interpolation function as the number of
support nodes increases. The first row shows a plot of the interpolated function
while the second row shows the corresponding support nodes required to build
The two examples discussed above illustrate the possibility of further re-
ducing the number of needed function evaluations for cases in which (i) the
32
The choice of support nodes now depends implicitly on the function that one
is interpolating. Therefore, one must decide which nodes to sample on-the-fly
for minimizing the number of function evaluations. This naturally leads to the
idea of adaptive sparse grid collocation, which is discussed in the next section.
As stated earlier, in the standard sparse grid approach, all the stochastic dimen-
sions are treated equally. By treated equally, we mean that the number of grid
points in each direction is equal. In most physical problems that one deals with,
dard sparse grid method is apparent by revisiting Eq. (2.23). The multi-index
|i| = i1 + . . . + iN that determines the number of sampling points in each dimen-
sion ik is sampled from the simplex |i| ≤ q. This ensures that if (n1 , n2 , . . . , nN ) is a
valid index, then (n2 , n1 , . . . , nN ) is also a valid index and so on. One simple ap-
33
dimensions is to consider a weighted simplex a · i ≤ q (this simple extension
is in fact not feasible [51]). Another approach is to allow any general index set
satisfying certain conditions. The problem essentially boils down to choosing
index sets from the set of conventional sparse grid points. One can heuristically
where e j is the j-th unit vector. This means that an admissible set contains for
every index k all indices which have smaller entries than k. The generalized
sparse grid includes both the conventional sparse grids as well as the classical
product formulae within it. This criterion still ensures the telescopic property
of the differential interpolant ∆i = U i − U i−1 . This means that one can construct
a better interpolant starting from a coarse interpolant by just sampling on the
unique nodes of the finer interpolant, see Eq. (2.23). Denote the space of all
34
admissible indices by F . The generalized sparse grid construction is then [51]
X
AS ( f ) = (∆i1 ⊗ . . . ⊗ ∆i1 )( f ), (2.32)
i∈S
for an admissible set S ∈ F . By carefully constructing the index sets, it is pos-
sible to construct polynomials at least as good as the case of the conventional
sparse grids.
satisfied: the resulting index sets all remain admissible and a large reduction
in the interpolation error is achieved. To achieve the second criterion, an error
estimator has to be defined that quantifies the error reduction. Once a suitable
error indicator is determined, the corresponding error indicators are computed
One starts the interpolation construction with the smallest index 1 and adds
the indices (2, . . . , 1), . . . , (1, . . . , 2). One then obtains an initial set of error indica-
tors. The index set is partitioned into two sets, the active indices A and the old
indices O [51]. Newly added indices are denoted as active indices. The error in-
dicators of the forward neighbors of active indices have not yet been computed.
All other indices are denoted as old indices.
At each step, the forward neighborhood of the index set with the largest error
indicator is considered, and the corresponding error indicators computed. The
35
forward neighborhood of an index i are the N indices {i + e j | j = 1, . . . , N}. These
new sets are added to the active index sets. Then the index set with the largest
error indicator is considered and so on. The error indicator is chosen to be the
deviation of the computed value of a particular point from its expected point
(from the interpolation polynomial built at this level). This is just the differential
interpolation operator ∆i−1 ( f ) (see Eq. (2.23)) i.e. it is the difference between
P
the computed value of the function at an active index and the interpolated value
of the function using all the old indices.
In this way, the algorithm adaptively samples the regions of the conven-
tional sparse grid support region where the error is maximized. The following
subsection illustrates the effectiveness of this methodology with some sample
examples.
1 1
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Figure 2.7: The adaptive grid used for ρ = 1 (left) and ρ = 1000 (right). No-
tice the increased node points in the y direction for the second
case.
36
2.5.3 Numerical illustrations
We revisit the first problem in the sparse grid collocation section. Briefly, con-
2 −̺y2
sider the function f = e−x , where ̺ is a positive number. This function is
analytically smooth. For ̺ = 1, the function is isotropically smooth. The com-
putational domain is [−1, 1] × [−1, 1]. We look at the problem for increasing ̺.
As the anisotropy increases the adaptive procedure must sample more points
in that direction where the error is maximum (the y direction in this particular
case). Fig. 2.7 plots the adaptive grids used for the two particular cases of ̺.
Fig. 2.8 plots the error reduction with the number of support nodes. Compare
this with the case of using the conventional sparse grids (Fig. 2.6). The number
of nodes required to achieve the same interpolation error of 5 × 10−2 for the case
of ρ = 1000 is 577 using the adaptive sparse grid method as compared to 1537
using the conventional sparse grid method. As the anisotropy of the problem
increases, the adaptive sparse grid method selectively improves the function in
the y direction, until the same level of accuracy is reached in both directions.
We revisit the second problem discussed in the previous section. Briefly, con-
2
sider a function f = e−x +2sign(y) in the domain [−1, 1] × [−1, 1], where sign is
the signum function. The function f has a sharp discontinuity along the y = 0
plane. Fig. 2.9 shows the adaptive sparse grid used to compute the interpola-
tion function for this discontinuous function. Notice the large bias towards the y
direction, which is to be expected from an adaptive algorithm. Fig. 2.9 also plots
37
0
10
-1
10
ρ =1
Error
-2
10
ρ = 10
ρ = 100
-3 ρ = 1000
10
-4
10
0 1 2 3
10 10 10 10
Number of nodes
Figure 2.8: Error versus the number of nodes using the adaptive sparse
grid method for interpolating anisotropic functions. Compare
with the error plot using the conventional sparse grid method
shown in Fig. 2.3.
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
y x
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Figure 2.9: Left: Adaptive mesh for the discontinuous problem, Right: In-
terpolated discontinuous functions.
the interpolated function. The error after using just 553 nodal points is 1 × 10−1 .
To achieve the same level of error using the conventional sparse grid method
(see Fig. 2.5) 3300 points have to be sampled.
38
lems where every stochastic dimension is equally important. In most physical
problems encountered, the solutions exhibit some level of dissimilarity along
various directions or have some structure that can be easily exploited using the
adaptive sparse grid collocation method. Furthermore, there is very little over-
head in computing the error indicator function (since the error indicators are
just the incremental interpolation function at the new indices (Eqn. 2.22), [54])
mentation
The solution procedure can be broadly divided into three distinct operations:
The concept of the collocation method lies in effectively decoupling the physical
domain computation from the computations in the stochastic dimensions. The
deterministic codes used in the present work took in as input, the sparse grid
39
coordinates. For example, consider the case of two stochastic dimensions (ξ1 , ξ2),
that determine the input random process ω. The range of (ξ1 , ξ2) is assumed to
be [0, 1] × [0, 1] without loss of generality. The deterministic executable must
be able to take in the different sampled two-tuples (ξ1i , ξ2i ) and output logically
In the present work, we use a simple MATLAB wrapper program that first ini-
tializes the stochastic dimensions and constructs the sparse grid coordinates.
In the case of the conventional sparse grid collocation method, all the sampled
nodes are known a priori. Hence in this case, the deterministic code is run and
the appropriately named result files are available. The wrapper program then
reads the input data and constructs the interpolation function. For the rapid
construction of the interpolation functions, we utilize the sparse grid interpo-
lation toolbox developed by Klimle [57, 58]. In the case of the adaptive sparse
grid collocation method, the wrapper program builds the input files for the de-
terministic code, calls the executable of the deterministic program and reads in
the corresponding result at each step of the process.
2.6.3 Post-processing
40
collocation, the integration is very straightforward because the weights corre-
sponding to the known nodal positions are computed a priori. A data set con-
taining the weights for each nodal position, for different dimensions and differ-
ent interpolation depths was first created. Then the cubature over the stochastic
a particular error tolerance has been reached), the interpolation functions are
numerically integrated to get the corresponding weights associated with each
nodal point. This results in significantly larger post processing time for higher-
dimensional integrations.
Clenshaw-Curtis type grid has been shown to outperform other types of grids
in numerical tests [53, 54]. In all the numerical examples, the global error com-
puted is the maximum of the absolute variation of the interpolated value from
the actual value at the newly computed sparse grid points (supi (∆i )). That is,
at each level, the function is evaluated at some points. The error is defined as
the difference between the interpolated value (using the interpolation function
at the previous level) and the actual function value at these points. The global
error reported is the maximum of these errors.
41
The main focus of this work is to investigate the effects of input uncertain-
ties in natural convection problems. Consider a d-dimensional bounded domain
D ⊂ Rd with a boundary ∂Dd ∂Dn . Dirichlet boundary conditions are applied
S
on ∂Dd , while Neumann boundary conditions are applied on ∂Dn . Natural con-
vection sets in due to the thermal differential maintained across ∂Dd . Consider
also a complete probability space (Ω, F , P), where Ω is the event space, F ⊂ 2Ω
the σ-algebra, and P : F → [0, 1] is the probability measure. In the problems
that we consider, the input uncertainties occur due to the following random
Under these conditions the problem is to find stochastic functions that describe
∇ · u = 0, (2.33)
∂u
+ u · ∇u = −∇p + Pr∇2 u + F(u, θ), (2.34)
∂t
∂θ
+ u · ∇θ = ∇2 θ, (2.35)
∂t
42
usually be the Bousinnessq approximated buoyant force term −RaPrθeg , where
Ra is the thermal Rayleigh number and eg is the gravity vector.
ferent points on the right hand side boundary is correlated. This is physically
feasible through, say, a resistance heater, where the mean temperature remains
constant, but material variations cause local fluctuations from this mean temper-
ature. We assume this correlation to be a simple exponential correlation func-
tion, C(y1 , y2 ) = exp(−c · |y1 − y2 |). Here, c is the inverse of the correlation length
that is taken as 1. For ease of analysis, we consider a non-dimensionalized form
of the governing equations. The Prandtl number is set to 1.0 and the thermal
Rayleigh number is set to 5000. The upper and lower boundaries are thermally
where ξi are normally distributed (N(0, 1)). The eigenvalues of the correlation
function decay very rapidly. The two largest eigenvalues contribute to 95% of
43
T (y) = f (y,ω)
T (y) = 0.5
the energy of the spectrum. To showcase the sparse grid collocation method,
three cases, where the Karhunen-Loève expansion is truncated at 2, 4 and 8
Fig. 2.11 plots the variation in the error with increasing number of support
points for both the conventional isotropic sparse grid and the adaptive sparse
grid methods. In the case of the isotropic sparse grid method, the level of in-
terpolation is increased linearly for the three problems of dimensions 2, 4 and
8, respectively. For the two-dimensional case, the adaptive sparse grid colloca-
tion (ASGC) is marginally better than the conventional sparse grid collocation
(CSGC) technique. The same level of accuracy is achieved in 137 function eval-
44
Error
Error
Error
Number of nodes
Figure 2.11: Error reduction with increased number of support nodes for
2, 4 (top figures) and 8 (bottom figure) dimensions.
uations as compared to the 321 required in the CSGC method. As the number of
stochastic dimensions is increased, the ASGC method provides orders of mag-
nitude reduction in the number of functional evaluations. For the same mag-
nitude of error (∼ 4 × 10−2 ), in 4 dimensions the ASGC required 751 function
evaluations compared to the 7537 that the CSGC required (see Fig. 2.11). This
is illustrated clearly in the case of 8 stochastic dimensions where a 65 times re-
duction in the number of function evaluation is seen going from the CSGC to
ASCG methods (15713 to 241). What has to be noted is that this dramatic re-
duction in the number of functional evaluations is mainly due to the structure
of the solution. While the adaptive strategy neglects the last four dimensions
and quickly converges to an accurate solution, the isotropic sparse grid method
45
0.45 8.00
0.36 6.40
0.27 4.80
0.18 3.20
0.09 1.60
0.00 0.00
-0.09 -1.60
-0.18 -3.20
-0.27 -4.80
-0.36 -6.40
-0.45 -8.00
0.45 8.00
0.36 6.40
0.27 4.80
0.18 3.20
0.09 1.60
0.00 0.00
-0.09 -1.60
-0.18 -3.20
-0.27 -4.80
-0.36 -6.40
-0.45 -8.00
8.00
0.45
6.40
0.36
4.80
0.27
3.20
0.18
1.60
0.09
0.00
0.00
-1.60
-0.09
-3.20
-0.18
-4.80
-0.27
-6.40
-0.36
-8.00
-0.45
Figure 2.12: Mean Temperature and u velocity contours from different so-
lution strategies. Top row: A level 6 collocation method, Mid-
dle row: Monte-Carlo sampling over 65000 points, Bottom
row: Second-order GPCE expansion.
46
takes many more functional evaluations to do the same because it treats all
dimensions equally. As the number of dimensions (i.e more terms in the KL
expansion are taken into account) increases, the isotropic method treats these
dimensions equally, even though the higher dimension have a negligible effect
on the stochastic solution. On the other hand, the adaptive method takes this
structure into account and preferentially samples the dimensions with the larger
eigen values.
In the case of the conventional sparse grid collocation method for the prob-
lem truncated to two-dimensions, a level 6 interpolation scheme was used in
the sparse grid interpolation method. This resulted in solving the determin-
istic problem at 321 different realizations. In the case of the adaptive sparse
grid collocation method, 137 realizations were required. The sparse grid points
correspond to the Clenshaw-Curtis cubature points in two-dimensions. The
solution procedure is embarrassingly parallel. The complete simulation was
space were generated. A second order polynomial chaos expansion of the above
problem was also performed. The mean and standard deviation values of the
three methodologies match very closely as seen in Fig. 2.12 and Fig. 2.13, respec-
tively. The total computational time for the GPCE problem was 235 minutes.
Table. 1 shows representative times for the natural convection problem solved
47
0.60 6.78
0.53 6.10
0.47 5.43
0.40 4.75
0.33 4.07
0.27 3.39
0.20 2.71
0.13 2.03
0.07 1.36
0.00 0.68
0.60 6.78
0.53 6.10
0.47 5.43
0.40 4.75
0.33 4.07
0.27 3.39
0.20 2.71
0.13 2.03
0.07 1.36
0.00 0.68
0.60 6.78
0.53 6.10
0.47 5.43
0.40 4.75
0.33 4.07
0.27 3.39
0.20 2.71
0.13 2.03
0.07 1.36
0.00 0.68
48
Probability
Probability
Temperature Pressure
Probability
Probability
U velocity V velocity
using GPCE and collocation based methods. The dimension 2 and 4 problems
were solved using a second-order polynomial chaos representation, while the
problem in 8 dimensions was solved using a first-order polynomial chaos rep-
49
Higher-order moments and PDFs
randomly 500000 times from the 2 dimensional stochastic space and the corre-
sponding PDF from its histogram distribution was computed. The probability
distribution functions for the temperature, pressure and velocities at this spatial
point are plotted in Fig. 2.14. The applied random temperature boundary con-
ditions results in some realizations having temperatures (along the right wall)
both higher and lower than the left wall. This leads to instances where there is
a complete reversal in the direction of circulation. Furthermore, as seen from
Fig. 2.12, most of the flow in the horizontal direction is concentrated in two
50
2.7.2 Natural convection with random boundary topology
waviness and roughness. Waviness refers to the large scale fluctuations in the
surface. Roughness is the small scale perturbations of the surface. Roughness is
described by two components: the probability of the point being z above/below
the datum (PDF), and the correlation between two points on the surface (ACF).
The type of surface treatment that resulted in the surface can be directly con-
cluded from the auto-correlation function. For example, if the surface has been
shot-peened, its ACF would have a larger correlation distance that if the surface
had been milled. Usually the PDF of the position of a particular point is well
and Xiu [61] deals with this problem in a spectral stochastic framework.
length is set to 0.10. The mean roughness (absolute average deviation from a
1 th
smooth surface) is taken to be 100
of the characteristic length (L = 1.0). The first
eight eigenvalues are considered to completely represent the surface random-
51
T (y) = -0.5
ness. The problem is solved in eight dimensional stochastic space. The bottom
wall is maintained at a higher temperature of 0.5. The top wall is maintained at
a lower temperature of −0.5. The side walls are thermally insulated. The Prandtl
number is set to 6.4 (corresponding to water) and the thermal Rayleigh number
is set to 5000. No slip boundary conditions are enforced on all four walls.
fi (x) are the eigenvalues and eigenvectors of the ACF. A reference domain
([−1.0, 1.0] × [−0.5, 0.5]) and grid (100 × 50 uniform quadrilateral elements) are
considered and the nodes on the bottom boundary are transformed according
to the above formula to get each realization of the random boundary. The Jaco-
bian of the transformation from the reference grid to the new grid can be easily
computed (It is just the ratio of the elemental areas before and after the transfor-
52
Error
Number of nodes
Figure 2.16: Error reduction using the adaptive sparse grid methodology.
tive sparse grid collocation method (ASGC) method required 820 realizations
to construct the interpolation function. 8 nodes on the Velocity-3 cluster at the
Cornell Theory Center took 500 minutes for simulating the problem at all the
collocation points. Fig. 2.16 plots the reduction in the error with the number of
Fig. 2.17 shows some sample realizations of the temperature and velocity
contours. The first row of figures is the case when the roughness is minimal. The
temperature and velocity contours closely resemble those of Rayleigh-Bénard
53
0.438 6.363 8.369
0.313 4.547 6.449
0.188 2.731 4.529
0.063 0.915 2.610
-0.063 -0.902 0.690
-0.188 -2.718 -1.230
-0.313 -4.534 -3.150
-0.438 -6.350 -5.070
convection with smooth boundaries. The next 4 rows of figures show extremal
values of the surface roughness. The thermal plume changes position according
to the local surface geometry. Also note that the flow field has been substantially
54
0.437 6.530
0.312 4.670
0.187 2.810
0.062 0.950
-0.062 -0.911
-0.187 -2.771
-0.312 -4.631
-0.437 -6.492
4.192 0.355
2.528 0.253
0.864 0.152
-0.800 0.050
-2.465 -0.051
-4.129 -0.153
-5.793 -0.254
-7.457 -0.356
Figure 2.18: Mean values of the variables: Top left: Temperature, Top
right: u velocity component, Bottom left: v velocity compo-
nent, Bottom right: pressure.
Fig. 2.18 plots the mean values of the temperature, velocity and pressure. The
rough lower surface causes the mean temperature to be much more homoge-
nized. The mean velocity is larger than the case with smooth lower surface.
In the case of smooth lower surfaces, the u velocity shows symmetry about the
horizontal centerline. Notice that this is lost when the lower surface is made
rough. The rougher surface enhances the circulation patterns in the lower half
of the system. Fig. 2.19 plots the mean velocity vectors and some streamline
contours. The top corners of the system consist of recirculating rolls which re-
sult in reduced heat transfer at the edges as can be seen in the mean temperature
contour plot.
Fig. 2.20 plots the second moment of the temperature, velocity and pressure
variables. The standard deviation of the velocity is of the order of the mean
velocity. This suggests that there are multiple modes at which the system exists
55
-0.742352
2
29
-0.245137
49
0. 7
0.2
-1.73678
5
20
77
-1 . 2
51
395
46
1. 2
7
-0.2
0.
25
451
20
37
77
Figure 2.19: Mean velocity vectors and streamlines.
0.301 15.875
0.260 13.758
0.220 11.642
0.180 9.525
0.140 7.408
0.100 5.292
0.060 3.175
0.020 1.058
18.599 0.801
16.119 0.694
13.639 0.587
11.159 0.480
8.679 0.374
6.200 0.267
3.720 0.160
1.240 0.053
and the variation of the random variables across the stochastic space causes the
system to move from one stable state to another. If the state changes abruptly
as the random variable moves across a threshold value, a mode shift is said to
have occurred. The existence of multiple patterns in the region of parameters
56
Mode shifts and equilibrium jumps
One way to identify multiplicity of stable states is through visualizing the prob-
ability distribution functions of the dependent variables at a physically relevant
point. Small changes in the roughness causes the central thermal plume in the
case of smooth surface Rayleigh-Bénard instability to change positions. This can
cause reversal of the direction of the fluid flow in parts of the domain (compare
rows 2 and 3 in Fig. 2.17). We choose one such point, (0, 0.25) which has a large
deviation in the temperature to plot the PDFs. The PDFs were computed from
500000 Monte-Carlo realizations of the dependent variables. Fig. 2.21 plots the
corresponding distribution functions. Notice that there exist two peaks in the
distribution functions for the temperature pressure and the v velocity. This con-
firms the existence of two stable states. We assert that these shifts also occur
due to the non-linearities introduced by the surface roughness. To verify if this
is indeed the case, Fig. 2.22 shows the temperature and v velocity variations at a
specific point ([0.0, 0.25], chosen because of its large temperature and v velocity
second moment) plotted over the first two random variables. Notice the abrupt
jump in the temperature and v velocity as the random variables cross 0. Fig. 2.23
shows the projection of the variation onto a single dimension. The mode shift is
apparent in the abrupt change of the dependent variables as the random vari-
able crosses over zero.
57
Probability
Probability
Temperature Pressure
Probability
Probability
U velocity V velocity
58
0.2
15
0.1
10
0
5
Temperature
Y velocity
-0.1 0
-5
-0.2
-10
-0.3
-15
-0.4
-5 0 5 -5 0 5
Random variable Random variable
0.438
0.438
0.313
0.313
0.188
0.188
0.063
0.063
-0.063
-0.063
-0.188
-0.188
-0.313
-0.313
-0.438
-0.438
Rayleigh number of 1700 is chosen. This is just below the critical Rayleigh num-
ber of 1708. The other system and domain specifications are the same as the
previous problem. We study the problem in four stochastic dimensions.
The critical Rayleigh number of 1708 is computed for a infinitely long, flat
layer of fluid. The variations in the lower surface coupled with the fact that we
are in fact only considering a finite sized domain causes the onset of convection
in some of the stochastic space. Fig. 2.24 shows two such extreme realizations
of temperature at two different collocation points. Notice that the two patterns
almost mirror images of each other. But the behavior of the mean temperature is
along expected lines. Fig. 2.25 plots the mean and the standard deviation of the
temperature contours. The temperature evolution is predominantly conductive.
59
0.437 0.182
0.312 0.158
0.187 0.134
0.062 0.109
-0.062 0.085
-0.187 0.061
-0.312 0.037
-0.437 0.012
-0.01 0.5
-0.02
0
-0.03
-0.04 -0.5
Temperature
Y velocity
-0.05
-1
-0.06
-0.07 -1.5
-0.08
-2
-0.09
-5 0 5 -2.5
Random variable -5 0 5
Random variable
To further verify that no mode shift occurs as the stochastic phase space is
sampled, Fig. 2.26 plots the projection of the temperature and y velocity varia-
tions onto one dimension at a specific point ([0.0, 0.0]). The v velocity variation
is smooth with small gradients (compare with Fig. 2.23) while the temperature
variation is negligible.
Large dimensions
60
1
16
0.8
12
0.6
Eigenvalue
V2
8
0.4
0.2
4
0
0
0 0.5 1 1.5 2 5 10 15 20
V1 Index
T (y) = -0.5
type simulation. To represent the surface roughness, the ACF was considered to
be an exponential function. In this problem, we consider realistic ACF from ex-
perimental data. The surface roughness correlation function is given in Fig. 2.27
(taken from [72]). The corresponding eigen-spectrum is given in Fig. 2.27. The
first 20 eigenvalues correspond > 99% of the spectrum.
61
0.44
0.31
0.19
0.06
-0.06
-0.19
-0.31
-0.44
0.44
0.31
0.19
0.06
-0.06
-0.19
-0.31
-0.44
0.44
0.31
0.19
0.06
-0.06
-0.19
-0.31
-0.44
The upper wall is smooth while the lower wall is rough. The mean roughness
1 th
(absolute average deviation from a smooth surface) is taken to be 200
of the
characteristic length (L = 1.0). The first twenty eigenvalues are considered to
62
0.44
0.31
0.19
0.06
-0.06
-0.19
-0.31
-0.44
15.93
11.40
6.87
2.33
-2.20
-6.73
-11.26
-15.79
took about 1400 minutes on 48 nodes of the Velocity-3 cluster at the Cornell
Theory Center. Fig. 2.29 shows temperature contours at some of the collocation
points in 20 dimensional stochastic space. Fig. 2.30 shows the mean values of
the temperature and the v velocity. Fig. 2.31 plots the mean velocity vectors
and some streamlines. The standard deviations of the temperature and the v
velocity are shown in Fig. 2.32. Notice that the surface roughness causes thermal
plumes to form, which are seen in the variations in the temperature and the
v velocity. These plumes caused by the non-uniformities in the lower surface
result in improved heat transfer across the domain. This concept can be further
extended to the design of the ACF of the surface roughness to achieve enhanced
heat transfer characteristics. This will showcased in one of our forthcoming
publications.
63
35
-1.721
46
7 9 3. 4 11
84
06
0.
356
-4.2
79
0.844
-1.721
882
- 4. 28
821
1
Figure 2.31: Mean velocity vectors and streamlines.
0.17
0.14
0.12
0.10
0.08
0.06
0.03
0.01
7.63
6.62
5.60
4.58
3.56
2.54
1.53
0.51
medium with deterministic variable porosity has been well studied [63]. But
in most cases of heterogeneous porous media, the porosity either varies across
multiple length scales or cannot be represented deterministically. In the follow-
ing example, we focus on the case when the porosity is described as a random
64
u=v=0
T=1 T=0
u=v=0 Free fluid u=v=0
Porous medium
u=v=0
field. The usual descriptor of the porosity is its correlation function and the
mean porosity.
be free fluid. The rest of the domain is filled with a porous material. The porous
material is assumed to be Fontainebleau sandstone, whose experimental corre-
lation function is taken from [64]. The correlation function and a cross-sectional
image of the sandstone is shown in Fig. 2.34.
The governing equation for the fluid flow in the variable porosity medium
is given by (see Nithiarasu et. al[63] and Samanta and Zabaras [65]):
∂v v Pr (1 − ǫ)2 1.75 k v k (1 − ǫ)
+ v · ∇( ) = − 2
v − v + Pr∇2 v −
∂t ǫ Da ǫ (150Da)1/2 ǫ 2
ǫ∇p − ǫPrRaTeg (2.37)
where ǫ is the volume fraction of the fluid and Da is the Darcy number. Nat-
65
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 10 20 30 40 50 60
15
10
Eigenvalue
0
5 10 15 20
Index
porosity is ǫ = 0.8 and the porosity of the free fluid in the interior of the domain
is 1.0.
66
0.94 7.0
0.81 4.4
0.69 1.8
0.56 -0.8
0.44 -3.4
0.31 -6.0
0.19 -8.6
0.06 -11.2
9.5 0.6
7.1 0.5
4.6 0.4
2.2 0.3
-0.3 0.3
-2.7 0.2
-5.1 0.1
-7.6 0.0
Figure 2.36: Mean values of the variables: Top left: Temperature, Top
right: u velocity component, Bottom left: v velocity compo-
nent, Bottom right: pressure.
-0.020765
8 -1.06042
95
-0.020765
52
- 0.
-0.52958
-0.020765
6
2
91
.5
-1
221
4
29
65
26
-2.1
-2 .
591
-2.122
-1.
1
- 1. 0
604
2 -1.06042
-0.52958
-0.020765
67
0.097 5.056
0.084 4.382
0.071 3.708
0.058 3.034
0.045 2.359
0.032 1.685
0.019 1.011
0.006 0.337
6.936 0.054
6.011 0.047
5.086 0.040
4.161 0.032
3.237 0.025
2.312 0.018
1.387 0.011
0.462 0.004
porosity is 0.8. A level 4 interpolatory sparse grid with 3937 collocation points
was utilized for the stochastic simulation. A fractional time step method based
on the work of Nithiarasu et al. [63] is used in the simulation of the deterministic
problem. Fig. 2.36 plots the mean temperature, velocity and pressure contours.
Fig. 2.37 shows the mean velocity vectors and some streamlines. The standard
deviation of the dependent variables is shown in Fig. 2.38.
68
2.8 Conclusions
The generalized polynomial chaos approach and its derivatives have been ex-
tensively utilized to solve thermal and fluid flow problems involving random
inputs and boundary conditions. The GPCE is restrictive in the sense that it
drawbacks of the spectral stochastic methodology. We show that this leads nat-
urally to the tensor product collocation method [15] and the sparse grid colloca-
tion method [16]. An extension of the sparse grid collocation method to adap-
ward to convert any deterministic code into a code that solves the correspond-
ing stochastic problem. We provide a detailed road-map, description of the nec-
essary changes in the deterministic code and the required sub-routines for pre-
and post-processing to easily accomplish this change.
69
We have attempted to showcase the utility of the sparse grid collocation
method to solve large stochastic dimensional heat transfer problems. We have
solved realistic natural convection problems in rough surfaces with stochastic
dimensions up to d = 20 as well as natural convection problems through hetero-
geneous porous media. We further looked at the shifts in equilibrium that occur
due to slight variations in the boundary conditions and successfully captured
these equilibrium jumps.
70
CHAPTER 3
DATA-DRIVEN APPROACHES TO CONSTRUCTING STOCHASTIC
INPUT MODELS
The first step towards solving Eq. (2.1) is some form of numerical representa-
tion of the input random processes. A finite dimensional representation of the
abstract probability space is a necessary prerequisite for developing any viable
framework to solve Eq. (2.1). In this chapter, we discuss various techniques that
can be utilized to represent the input uncertainties, with an emphasis on utiliz-
ing experimental data to construct these representations. Note that the input
parameters to Eq. (2.1) can be either be scalars or functions.
Consider the case when the input parameter, α is a scalar. The problem of inter-
est is now to represent the uncertainty of α in a numerically feasible way that
utilizes experimental data. Experimental data, in this case, refers to a set of real-
71
where ξα is the random variable which represents the variability in the input pa-
rameter α. In case there are multiple (uncertain) scalar parameters {q, α} , each of
these input parameters should be represented as appropriate random variables
{ξq , ξα}. Assume that there are n = 2 scalar input parameters, the governing
In some cases, the input parameters are functions showing variability in space.
This could be, for instance, imposed boundary conditions or property variabil-
ity in heterogeneous media. Assume (for clarity), that q(x) is one such input pa-
rameter that is uncertain. One simple way of representing the uncertainty of this
spatially varying random process (also called a random field) is to consider q(x)
as a collection of scalar values at different spatial locations. That is, the function
q(x) is approximated as a collection of points q′ = {q(x1 ), q(x2 ), . . . , q(xk )}. Each
of these individual parameters q(xi ) show some uncertainty and hence can be
represented as appropriate random variables. Note that the representation of
72
nite value of k, the above procedure of approximating a random field as a set of
independent random variables results in a nsd + 1 + k dimensional stochastic dif-
ferential equation for the evolution of the dependent variable θ. The governing
stochastic differential equation is written as:
One problem with the above strategy is that the number of random variables
(the stochastic dimensionality) quickly becomes very large and can pose signif-
expansion.
73
random process q(x, ω) ∈ L2 (Ω, F , P) can be expressed as a summation
∞
X
q(x, ω) = qi (x)ζi (ω), (3.4)
i=1
field onto the basis. Eq. (3.4) involves an infinite summation and hence is
computationally intractable. The Karhunen-Loève expansion (KLE) considers
finite-dimensional subspaces of L2 (Ω, F , P) that best represent the uncertainty
in q(x, ω). The KLE for a stochastic process q(x, t, ω) is based on the spectral
decomposition of its covariance function Rhh (y1 , y2 ). Here, y1 and y2 denote the
spatio-temporal coordinates (x1 , t1) and (x2 , t2 ), respectively. By definition, the
covariance function is symmetric and positive definite and has real positive
eigenvalues. Further, all its eigenfunctions are orthogonal and span the space to
where E(q(x, t, ω)) denotes the mean of the process and {ξi (ω)}∞
i=0 forms a set of
Z
Rhh (y1 , y2 ) fi (y2 )dy2 = λ f1 (y1 ). (3.6)
D×T
The chief characteristic of the KLE is that the spatio-temporal randomness has
sion represents most of the process with arbitrary accuracy. The expansion in
Eq. (5.5) is typically truncated to a finite number of summation terms, k. Using
74
this technique, the representation becomes
In many cases, one can only obtain limited experimental information about the
variation in a particular property/conditions that is essential to analyzing a par-
ticular system. This is particularly apparent when one is dealing with systems
ture at a few locations or some higher order correlation functions, Fig. 3.1). This
leads to an analysis of the problem assuming that the properties are random
fields satisfying some experimentally determined statistics. In the context of a
computational framework to perform such analysis, one needs to construct a
75
Exact distribution of components
Polycrystals and unknown. Topology and properties
multicomponent Application
vary.
microstructures
mate space, we are essentially sampling over all possible, plausible microstruc-
tures (or any other property variations).
76
(or microstructures). These snapshots are either experimentally obtained or re-
constructed using the experimental statistical information available. A model
reduction scheme (based on Principle Component Analysis (PCA)) was for-
mulated to represent this set of snapshots/property maps (which is a large-
boundary distribution [20]. Each point on the linear manifold shown corre-
sponds to a microstructure with a unique grain distribution. By sampling over
this space, one essentially samples over the space of all allowable microstruc-
tures. The figure on the right shows a similar orientation distribution model.
Following this, in [21], a more flexible non-linear method to encode the mi-
crostructural variability was developed. This framework is based on concepts
borrowed from psychology [22, 23], cognitive sciences [24, 25], differential ge-
ometry [26, 27, 28] and graph theory [29, 30]. The key mathematical concept on
which the framework is constructed on is the so-called principle of ‘manifold
learning’: Given a finite set of unordered data points in high dimensional space,
construct a low dimensional parametrization of this dataset. The finite set of
77
plausible realizations of the property map. In [21], we show that this set of
unordered points (satisfying some experimental statistics) lies on a manifold
embedded in a high dimensional space. That is, all the possible variations of
the quantity of interest (given a limited set of information about its variability, in
78
The corresponding governing stochastic differential equation is
We detail the non-linear model reduction strategy in the rest of the chapter:
uniquely occupied by one of the two phases (0 or 1) and that each phase has
a given diffusivity). That is, for a microstructure specified as a distribution of
two phases in a domain, the thermal diffusivity distribution is given by simply
replacing the phase description (0 or 1) at each point on the domain by its cor-
responding thermal diffusivity (α0 or α1 ). From this simple scaling, for each
79
realization of the microstructure, we can compute the corresponding realization
of the diffusivity, α (α = k/ρC p ).
Let MS be the space of all microstructures that satisfy the statistical proper-
ties S . This is our event space. Every point in this space is equiprobable. Conse-
quently, we can define a σ-algebra G and a corresponding probability measure
∂u(x, t, ω)
= ∇ · [α(x, ω)∇u(x, t, ω)] + f (x, t), x ∈ D, t ∈ [0, T f ], ω ∈ MS , (3.11)
∂t
the random topology field into a finite set of uncorrelated random variables.
Upon decomposition and characterization of the random inputs into d random
80
variables, ξi (ω), i = 1, . . . , d, the solution to the SPDE Eq. (3.11) can be written as
as follows:
∂u(x, t, ξ)
= ∇ · [α(x, ξ)∇u(x, t, ξ)] + f (x, t), x ∈ D, t ∈ [0, T f ], ξ ∈ Γ. (3.13)
∂t
For the sake of brevity, we will denote the system of equations consisting of
Eq. (3.13) and appropriate initial and boundary conditions (here for simplicity
assumed to be deterministic) as B(u : x, t, ξ) = 0.
81
where ρ(ξ) : Γ → R is the joint probability distribution function for the set of ran-
dom variables {ξ1, . . . , ξd }. In the following sections, we describe the non-linear
model reduction framework for computing α(x, ξ). Details of the implementa-
tion of the Smolyak algorithm for Eq. (3.13) can be found in [20].
ideas
In our recent work [20], a linear model reduction strategy was developed to
convert experimental statistics into a plausible low-dimensional representation
of two-phase microstructures. The first step in that formulation was the con-
version of the statistical information into a set of plausible 3D realizations. This
feature of first converting the given experimental statistics into a data set of
plausible 3D microstructures is continued in the developments featured here.
This is motivated by the fact that there exists a variety of mature mathemati-
cal and numerical techniques that convert experimental data and statistics into
various techniques that have been developed [20, 74, 75, 76, 77], to convert ex-
perimental statistics into a plausible 3D reconstructions of two-phase composite
microstructures as well as polycrystalline materials.
82
3.3.1 Linear model reduction: where does it fail?
The PCA based linear model reduction technique that we had formulated [20]
has some drawbacks. The most critical of these is that any PCA based approach
can only identify the closest linear subspace to the actual space (which is possi-
bly non-linear) in which the data reside. This directly translates into the fact that
Number of samples
Figure 3.2: Plot of the number of samples versus the number of eigenvec-
tors required to represent 80% of the ‘energy’ spectrum con-
tained in the samples.
83
The issue brought out in the discussion above can be understood in a more
intuitive way by looking at the simple surface shown in Fig. 3.3(a). The set of
points shown in Fig. 3.3(a) all lie on a spiral in 3D space. The global coordinates
of any point on the spiral is represented as a 3-tuple. Any PCA based model
tries to fit a linear surface such that the reconstruction error is minimized. This
is shown by the green plane which is a 2D representation of the data. This
clearly results in a bad representation of the original data. On the other hand,
knowledge of the geometry of the 3D curve results in representing the global po-
tation naturally arises. Thus, while PCA based methods would require a 3D
representation to accurately describe the spiral shown in Fig. 3.3, a non-linear
model reduction method would identify the geometry of the curve, unravel it
and provide a simple 2D representation of the data.
The essential idea of nonlinear model reduction finds its roots in image com-
pression and related digital signal processing principles. Figure 3.4(a) shows
multiple images of the same object taken at different left-right and up-down
84
PCA
3D 40
20
data
15
10
15
10
-5
5
0
-10 -5
-10
(a) (b)
Figure 3.3: A set of points lying on a spiral in 3D space. The global coordi-
nates of any point on the spiral is represented as a 3-tuple. The
figure on the left (a) shows the reduced-order model resulting
from a linear PCA based reduction. The figure on the right (b)
depicts a non-linear strategy that results in an accurate repre-
sentation of the 3D spiral (data taken from [78]) that works by
“unravelling and smoothing” the 3D spiral into a 2D sheet.
curve as the manifold) representing all possible pictures of this object is embed-
Analogous to the problem defined above (using Fig. 3.4 (a)), we define a
problem based on the images in Fig. 3.4 (b). Fig. 3.4 (b) shows multiple mi-
crostructures that satisfy experimentally determined two-point correlation and
volume-fraction. Each microstructure is a 65 × 65 × 65 binary image. Each mi-
crostructure that satisfies the given experimental statistics is a point that lies
on a curve (manifold) embedded in 65 × 65 × 65 = 274625 dimensional space.
The problem of ‘manifold learning’ or parameter estimation as applied to this
85
(a) (b)
Figure 3.4: Figure (a) on the left shows images of the same object
(from [78]) taken at various poses (left-right and up-down)
while Figure (b) on the right shows various two-phase mi-
crostructures that satisfy a specific volume-fraction and two-
point correlation.
situation is as follows:
Classical methods in manifold learning have been methods like the Princi-
ple Component Analysis (PCA), Karhunun-Loève expansion (KLE) and multi-
dimensional scaling (MDS) [79]. These methods have been shown to extract
local geometry of the data. They aim to map nearby points on the manifold
86
to nearby points in the low-dimensional representation. Such methods, Locally
Linear Embedding (LLE) [22], Laplacian Eigen Maps, Hessian Eigen maps, es-
sentially construct a homeomorphic mapping between local sets in the manifold
to an open ball in a low-dimensional space. The complete mapping is a union
suitable metric) map to nearby points in the low-dimensional space and far-
away points map to faraway points in the low-dimensional space. Though both
approaches are viable, we focus our attention to global methods of non-linear
dimension reduction. The global approach has been shown to provide a faith-
The basic premise of the global methods (particularly the Isomap [23] algo-
rithm) is that ‘only geodesic distances reflect the true low-dimensional geome-
try of the manifold’. The geodesic distance (between two-points) on a manifold
can be intuitively understood to be the shortest distance between the two-points
along the manifold. Subsequent to the construction of the geodesic distance be-
tween the sample points (x) in the high-dimensional space, the global methods
construct the low-dimensional parametrization simply as a set of points (y) ly-
ing in a low-dimensional space that most accurately preserve the geodesic dis-
87
high-dimensional data into low-dimensional points. The global based methods
solve the following version of the problem statement:
This section is divided into five parts. We first introduce some mathemati-
cal preliminaries and define an appropriate distance function D between two
points in MS . For the straightforward construction of a transformation F :
88
Section 3.4.4 details the application of the Isomap algorithm (along with the
estimate of the optimal dimensionality from Section 3.4.3) to construct the low-
dimensional parametrization of the input data. Finally Section 3.4.5 details a
non-parametric mapping that serves as the reduced-order, data-driven model
the microstructure is n = q × q × q.
Definition 4.2: Let MS denote the set of microstructures {xi } satisfying a set of
material topology and property distribution than, say, the volume fraction.
89
correlation function.
For clarity of presentation, we restrict our analysis to the set of two-phase mi-
Definition 4.4: The function D : MS2 × MS2 → [0, ∞) is defined for every x1 , x2 ∈
90
MS2 as
Remark 4.2: Notion of distance and equivalence between two microstructures: The
function D(x1 , x2 ) provides a notion of distance between two microstructures
x1 , x2 ∈ MS2 . Since x1 , x2 belong to MS2 , both have identical volume fraction, S 1
and two-point correlation S 2 . We naturally denote them as equivalent if they
have the same upper-correlation function, S 3 . Since we are dealing with statisti-
cally reconstructed microstructures, this definition of equivalence ensures that
two microstructures are identical if their higher-order topological characteriza-
tion is identical.
Remark 4.3: Any other mapping that satisfies the three properties of a met-
ric [82] can be used as a measure of equivalence and distance. Since we are
91
Lemma 4.2: The metric space (MS2 , D) is totally bounded.
Proof: Follows from Lemma 4.2 and Lemma 4.4 (see Theorem 45.1 in [26]).
proximations
Definition 4.5: Denote the intrinsic geodesic distance between points in MS2
by D M . D M is defined as
where γ varies over the set of smooth arcs connecting x1 and x2 . We wish to
remind the reader that length of the arcs in the equation above are defined using
the distance metric D.
lying in MS2 . Hence, the Definition 4.5 of the geodesic distance is not partic-
ularly useful in numerically computing the distance between two-points. An
92
approximation of the geodesic distance is required to proceed further. Such an
approximation is provided via the concept of graph distance. We subsequently
show that this approximation asymptotically matches the actual geodesic dis-
tance (Eq. (3.18)) as the number of samples, N, increases (see Theorem 4.2 be-
low).
The unknown geodesic distances in MS2 between the data points are ap-
proximated in terms of a graph distance with respect to a graph G constructed
on the data points. This neighborhood graph G is very simple to construct [23].
Two points share an edge on the graph if they are neighbors. The neighbor-
For points close to each other, the geodesic distance is well approximated
by the distance metric D. This is because the curve can be locally approxi-
mated to be a linear patch, and the distance between two points on this patch
is the straight line distance between them. This straight line distance is given
by the distance metric D, which is just the edge length between the points on
the graph, G. On the other hand, for points positioned faraway from each other,
93
trinsic geodesic distance. The asymptotic convergence of the graph distance to
the geodesic distance is rigorously stated in Theorem 4.2. This theorem utilizes
some parameters for the quantification of the geometry of the manifold, partic-
ularly the minimal radius of curvature, ro and the minimal branch separation
so . For the sake of completeness, we state the theorem below. For the sake of
brevity, we leave out the definitions of these abstract parameters (the interested
reader is referred to [23, 26] for discussion of these terms).
dimensional representation
94
We draw from the recent work in [27, 28], where the intrinsic dimension
of an embedded manifold is estimated using a novel geometrical probability
approach. This work is based on a powerful result in geometric probability -
the Breadwood-Halton-Hammersley [83] theorem where d is linked to the rate
The sequel utilizes concepts from graph theory. We provide some of the
essential definitions below. For a detailed discussion of graphs, trees and related
constructs, the interested reader is referred to [29]. Consider a set of k points
(vertices) and a graph defined on this set. A graph consists of two types of
elements, namely vertices and edges. Every edge has two endpoints in the set
• A tree is a graph in which any two vertices are connected by exactly one
path.
form a tree.
Definition 4.7: The geodesic minimal spanning tree (GMST) is the minimal
spanning tree of the graph G. The length functional, L({x}) of the GMST is de-
95
P
fined as L({x}) = minT ∈T N e∈T |e|, where T N is the set of all spanning trees of the
graph G and e are the edge-weights of the graph. This is in fact, simply the total
weight of the tree.
theorem:
Theorem 4.3, particularly the asymptotic limit given in Eq. (3.19) for the
length functional of the GMST provides a means of estimating the intrinsic di-
mension of the manifold. It is clearly seen that the rate of convergence of L({x})
is strongly dependent on d. Following [30], we use this strong rate dependence
to compute d. Defining lN = log(L({x})), gives the following approximation for
lN = a log(N) + b + ǫN , (3.20)
96
where
d−1
a = , (3.21)
d
b = log(βmC), (3.22)
and ǫN is an error residual that goes to zero as N → ∞ [27, 28, 30]. The intrinsic
dimensionality, d can be estimated by finding the length functional for different
number of samples N and subsequently finding the best fit for (a, b) in Eq. (3.20).
Section 3.4.2 presented the construction of the geodesic distance between all
pairs of input points, while Section 3.4.3 provided an estimation of the dimen-
sionality, d, of the low-dimensional representation. Denote as M, the pair-wise
distance matrix (based on the geodesic distance), with elements di j = DG (xi , x j )
dimensional space such that the coordinates of these N points yield a Euclidean distance
matrix whose elements are identical to the elements of the given distance matrix M.
97
N × N centering matrix [79, 84], H = I − N1 11′ with elements hi j = δi j − 1/N. Define
the N × d matrix of points Y = (y1 , . . . , yN )T . Assume without loss of generality
PN
that the centroid of the set of points is the origin ( i=1 yi = 0). Define the N × N
matrix B whose components are the scalar products yTi · y j .
d
X
bi j = yik y jk = yTi · y j . (3.24)
k=1
B = HAH. (3.25)
B = ΓΛΓ. (3.26)
B = YYT . (3.27)
From Eq. (3.26) and Eq. (3.27), an estimate of Y in terms of the largest d eigen-
1
vectors of B follows: Y = Γd Λd 2 , where Λd is the diagonal (d × d) matrix of
98
PCA based techniques truncate the expansion based on representing some per-
centage of the eigen-spectrum i.e. choose the largest d eigenvalues that account
Pd
i λ
for, say, 95% of the eigen-spectrum. ( Pi=1
N
λ
> 0.95).
i=1 i
The stochastic collocation procedure for the solution of SPDEs involves com-
MS2
(Section 3.4.3) convert these points into a set of points {yi }, i = 1, . . . , N belonging
to a convex set A. This convex region A ⊂ Rd , defines the reduced representation of
the space of microstructures, MS2 , satisfying the given statistical correlations S2 . A
can be considered to be a surrogate space to MS2 . One can access the complete
99
egy for solving SPDEs, one constructs the statistics of the dependant variable by
sampling over a discrete set of microstructures. Since we propose to utilize A as
a reduced representation of MS2 , we sample over a discrete set of points ξ ∈ A
instead. But we have no knowledge of the image of a random point ξ ∈ A in
the microstructural space MS2 (we only know that the points {yi } for i = 1, . . . , N
map to the microstructures {xi } for i = 1, . . . , N). For a usable reduced-order model of
the microstructure space, an explicit mapping F −1 from A to MS2 has to be constructed.
can be trained using the sample points ({yi }, {xi }), i = 1, . . . , N to construct a non-
parametric mapping F −1 . There have been recent reports of variants of the
Isomap algorithm that along with constructing the reduced-order representa-
tion of the samples also construct an explicit mapping between the two sets [85].
But there are two significant issues that have to be considered when one utilizes
such mapping strategies: 1) Most of these explicit mapping strategies are es-
sentially some form of interpolation rule that utilize the sample set of values
({y}, {x}). One has to make sure that the interpolated result, x (x = F −1 (ξ)) for
some arbitrary point ξ ∈ A actually belongs to MS2 . 2) Care must be taken to for-
mulate the explicit mapping in a way that results in a computationally simple
methodology of finding the images of points. This is very significant consid-
ering the fact that we will potentially deal with very large pixel sized images
tionally simple mappings between the two spaces MS2 and A keeping in mind
100
the issues raised above.
This is the simplest map (illustrated schematically in Fig. 3.5) that sets the image
of an arbitrary point in the region A to the image of the nearest sampled point.
The nearest point is the point that is the smallest Euclidean distance from the
given arbitrary point. This method is particularly useful when the sampling
density (α as defined in Theorem 4.5) is large, i.e. this method results in a reliable
A⊂Rd MS ⊂ R n
Figure 3.5: The figure above illustrates the simplest possible mapping be-
tween the low-dimensional region A and the high-dimensional
microstructural space MS2 . Given an arbitrary point ξ ∈ A,
find the point yk closest to ξ from the sampled points {yi },
i = 1, . . . , N. Assign the image value of yk i.e. xk to the image of
ξ.
actual image and the mapped image is given by e = eA . This error can be made
arbitrarily small by increasing N.
101
Method 2: Local linear interpolation
Utilizing the fact that the isometric embedding that generated the points yi from
xi conserves distances, the image of ξ is then given by
Pk x̂i
i=1 l
x = Pk 1i . (3.30)
i=1 li
That is, the image of ξ is the weighted sum of the images of the k-nearest neigh-
bors of ξ (where the nearest neighbors are taken from the N-sampled points).
A⊂Rd MS ⊂ R n
Figure 3.6: The figure above illustrates a local linear (k-neighbor) interpo-
lation mapping between the low-dimensional region A and the
high-dimensional microstructural space MS2 .
The linear interpolation procedure is based on the principle that a small re-
gion in a highly-curved manifold can be well approximated as a linear patch.
102
This is in fact one of the central concepts that result in local strategies of non-
linear dimension reduction [22, 24] (see Section 3.3 for a discussion of global
versus local strategies of non-linear dimension reduction). This linear patch is
constructed using the k-nearest neighbors of a point ξ. As the sampling density
(the number of sample points, N) used to perform the nonlinear dimension re-
duction increases, the mean radius of the k-neighborhood of a point approaches
zero (limN→∞ maxi=1,...,k kξ − ŷi k2 → 0), ensuring that the linear patch represents
the actual curved manifold arbitrarily well.
■
□
l1 l2
Figure 3.7: Simple estimate for the interpolation error in Method 2: Let r
denote the local radius of curvature of the function near the
point xexact (the filled square). We approximate the curve by a
linear patch, resulting in some error. This error is the distance
between the approximate linear image, x (the unfilled square)
and the actual point, xexact (the filled square). This distance can
be approximated from simple geometry as a function of r and
the local geodesic distance between the points.
proximated from the geodesic distances between the points. Let r denote the
radius of curvature of the manifold at x. The error in the interpolation based
103
representation is caused by considering the space to be locally linear, when it
is curved (see Fig. 3.7). This error is approximated from simple geometry as
q
e = maxi=1,...,k (r − r2 − l2i ).
The local linear interpolation method (Method 2) can be made exact by sim-
ply projecting the image obtained after interpolation onto the manifold (shown
schematically in Fig. 3.8). This ensures that the image lies on the manifold MS2 .
provides a good approximation of the exact image, x̂ ≈ xexact , where xexact is that
microstructure on MS2 whose geodesic distances from each of the k microstruc-
tures xi , i = 1, . . . , k is li . The errors in this approximation are due to the fact that
the approximation x̂ does not usually lie on the manifold (as seen in Fig. 3.8). That
is, x̂ does not satisfy the statistical correlations S = {S 1 , . . . , S p } that all points on
the manifold satisfy.
The projection operation essentially modifies the point x̂ to satisfy these cor-
relations. This can be achieved computationally by performing a stochastic op-
timization problem starting from x̂ [76, 80, 86]. Since x̂ is very close to xexact ,
these algorithms are guaranteed to reach the local minima defined by xexact . In
the context of the numerical examples presented in this work, using two-phase
microstructures (see Section 3.6), the stochastic optimization is done as follows:
Starting from the approximate microstructure x̂, compute the volume fraction
and two-point correlation of this image. Change the pixel values of t sites in this
104
A⊂Rd MS ⊂ R n
Figure 3.8: The local linear interpolation method can be made exact by
simply projecting the image obtained after interpolation onto
the manifold. This is illustrated in the figure above, where
the botted blue line represents the linear interpolation and the
curved line represents the projection operator that constructs
the image lying on the manifold.
as a version of simulated annealing, with the starting point, x̂ being close to the
optimal point, xexact .
Figure 3.9: Sample illustration of the projection step: The figure on the
left is a microstructure after interpolation. Projecting it onto
the manifold yields the microstructure on the right. There is
negligible change in the three point correlation between the mi-
corstructures.
105
One question that arises from this mapping strategy is the following: The
initial goal of the mapping was to construct a microstructure (lying on the man-
ifold) that was a distance (geodesically) li from xi . The interpolation step en-
sures this distance (but the microstructure does not lie on the manifold). How
much deviation from this distance does the projection operator cause? By swap-
ping the pixels in the microstructure during the stochastic optimization step, the
three-point correlation is being changed (which is a measure of distance). But in
all of our numerical experiments, this optimization converged within 1000 such
pixel flips, thus negligibly affecting the three-point correlation (since the three-
point correlation is computed by sampling over 65 × 65 × 65 ∼ 300, 000 points,
changing 0.3% of the pixels is negligibly small). This is illustrated pictorially
using a sample example shown in Fig. 3.9. The number of neighbors used for
In all three methods detailed above, we only utilize the given input data {xi }
to construct the mapping for an arbitrary point ξ ∈ A. The mapped microstruc-
ture, F −1 (ξ) is constructed solely based on the available input microstructures
and not based on direct reconstruction from moments. In case of the projection
operation used in method 3, the mapped microstructure is moved onto the man-
ifold MS 2 . This operation changes less that 0.3% of the microstructure having
negligible effect on the thermal behavior of the microstructure.
106
3.4.6 The low-dimensional stochastic input model F −1 : A →
MS 2
Each of the microstructures in MS2 (by definition) satisfies all required sta-
tistical properties, therefore they are equally probable to occur. That is, ev-
This section contains recipes for the numerical implementation of the theoreti-
cal developments detailed in the previous sections. We divide this section into
various subsections that sequentially discuss the data-driven strategy, starting
with the generation of the samples given some limited microstructural informa-
107
tion (Section 3.5.1), the algorithm for constructing the low-dimensional region
A (Section 3.5.2) and the Smolyak algorithm for the solution of the SPDEs (Sec-
tion 3.5.3).
correlations are enforced during the reconstruction of the GRF using the given
information [87, 69, 88]. With this method, a set of N 3D models of the property
variations can be generated.
108
Step 2: Construct the neighborhood graph, G, of this sample set. That is,
determine which points are neighbors on the manifold based on the distance
P(i, j). Find the nearest k-neighbors of each point. This is performed using a
sorting algorithm (O(NlogN) complexity). Connect these points on the graph G
and set the edge lengths equal to P(i, j). The total complexity of this operation
is O(N 2 logN).
for i = 1:N
G(1:k,i) = I(2:k+1)
Step 3: Estimate the geodesic distance M(i, j) between all pairs of points on
the manifold. This can be done by computing the shortest path distances in the
graph G. There are several algorithms to compute the shortest path on a graph.
In our implementation, we utilize Floyd’s algorithm to compute M(i, j). The
complexity of this step is O(N 3 ).
109
Initialize M(i, j) as
for k = 1:N
space A. Using the graph G [30], estimate the average geodesic MST length.
This is done as shown below.
L(p) = a log(p) + ǫ p
We utilize the code in [89] to compute the length functional of the MST. The
complexity of this step is O(NlogN).
110
Step 5: Construct the d-dimensional embedding using MDS.
1
Compute A(i, j) = − M2 (i, j)
2
Compute B = HAH
A = convex hull(y1 , . . . , yN )
111
3.5.3 Utilizing this reduced-order representation: Stochastic
collocation
tic model for the solution of SPDEs. We utilize a sparse grid collocation strategy
for constructing the stochastic solution [14]. The method essentially solves the
problem at various points ξ on the stochastic space and constructs an interpola-
tion based approximation to the stochastic solution. The sparse grid collocation
images are utilized as inputs (property maps) in the solution of the SPDEs.
Remark 5.1: In the stochastic collocation approach, the collocation points are
usually given in the unit hypercube, i.e. p ∈ [0, 1]d . As a first step, this point p
must be mapped to a corresponding point ξ ∈ A.
for i = 1 : M
compute xi = F −1 (ξ i )
112
3.6 Illustrative example
solution into multiple sections. 1) The first step is the extraction of topological
statistics from the experimental image provided. These statistics are then uti-
lized to reconstruct a large set of 3D microstructures {xi }, i = 1, . . . , N. 2) The
next step is to construct the low-dimensional representation of the class of mi-
113
Data extraction and sample set construction
formation from the experimental image. The image is cropped, deblurred and
discretized. The volume fraction of silver is p = 0.2. The experimental two-point
correlation is extracted from the image. The normalized two-point correlation
L2 (r)−p2
(g(r) = p−p2
), is shown in Fig. 3.11. The data extraction was performed in
Matlab.
1
0.8
0.6
g(r)
0.4
0.2
0
0 5 10 15 20
r (um)
114
The next step is to utilize these extracted statistical relations (volume frac-
tion and two-point correlation) to reconstruct a class of 3D microstructures. We
utilize a statistics based reconstruction procedure based on Gaussian Random
Fields (GRF). In this method, the 3D microstructure is obtained as the level cuts
d and a cutoff scale rc . For a specific choice of (β, d, rc), one can construct a
microstructure from the resulting GRF. The (theoretical) two-point correlations
corresponding to this reconstructed microstructures is computed. Optimal val-
ues of (β, d, rc) are obtained by minimizing the error between the theoretical
Using the optimal parameters of the GRF (to match with the experimental
data), realizations of 3D microstructure were computed. Each microstructure
consisted of 65 × 65 × 65 pixels. This corresponds to a size of 20µm × 20µm × 20µm.
115
1
Reconstructed
0.8
Experimental
0.6
g(r)
0.4
0.2
0
0 5 10 15 20
r (um)
The GRF based reconstruction detailed above was used to generate a set of
N = 1000 samples of two-phase microstructure. Each microstructure is repre-
sented as a 65 × 65 × 65 pixel image. The three-point correlations of all these
gle lie on the same phase. S 3 (a, b, c) is the ratio of the number of such successful
placements over the total number of tries. In our computations, we randomly
116
place 500, 000 triangles to compute S 3 for each value of (a, b, c). The total com-
putational time to reconstruct 1000 microstructures along with their S 3 was 30
minutes on 25 nodes of our in-house Linux cluster.
40000
30000
20000
Log(length of MST)
10000
Figure 3.14: Plot of the length functional of the MST of the graph G for
various sample sizes.
the slope of this line (see Eq. (3.20)). The slope of the curve is computed using a
simple least squares fit. The optimal dimensionality was estimated to be d = 9
(a ∼ 0.885). The total computational time to estimate the dimensionality was 8
minutes on a 3.8GHz PC.
117
Multi Dimensional Scaling is performed using the geodesic distance matrix
M. The 9 largest eigenvalues and their corresponding eigenvectors are used to
represent the input samples. The low-dimensional region A is constructed as
the convex hull of these N(= 1000) 9-dimensional points ξi . This region coupled
with the mappings developed in Section 3.4.5 define the reduced-order stochas-
tic input model F −1 : A → MS 2 .
0.0068
Residual variance
0.95
Eigenspectrum
0.0064
Optimal dimension
Optimal dimension
0.006
0.9
0.0056
0.85
5 10 15 20 5 10 15
Dimension Dimension
Pd
i λ
Figure 3.15: (Left) The cumulative eigenspectrum of the data, Pi=1 N
λ
.
i=1 i
(Right) The residual variance for different dimensionalities of
the region A, computed from MDS.
the region A based on simple variance errors (i.e. choose the value of d that ac-
counts for 90% of the variance in the data). Fig. 3.15(a) plots the eigenspectrum
Pd
i λ
of the computed eigenvalues of B ( Pi=1
N
λ
). All dimensionalities beyond d = 4
i=1 i
account for over 90% of the variance in the data. Hence, there arises some ambi-
guity in simply choosing the dimensionality of the reduced mode based on this
plot. Fig. 3.15(b) plots the residual variance of the low-dimensional representa-
tion for various dimensionalities. The residual variance measures the difference
between the intrinsic manifold distance matrix, M and the pairwise Euclidian
118
matrix, DA , recovered from MDS (see Eq. 3.23) for various dimensionalities d.
It is defined in terms of the element-wise correlation between the two matrices,
e = 1 − r2 (DA , M), where r here is the standard linear correlation coefficient [91].
Notice that all the dimensions above d = 8 have fairly small variance, with
ξ ∈ A serve as the stochastic input for the diffusion equation. A simple diffusion
problem is considered (Eq. 3.13). A computational domain of 65 × 65 × 65 is
considered (this corresponds to a physical domain of 20µm × 20µm × 20µm ).
The random heterogeneous microstructure is constructed as a 65 × 65 × 65 pixel
119
1
10
0
10
-1
10
-2
10
Error
-3
10
-4
10
-5
10
-6
10 0 1 2 3 4
10 10 10 10 10
# of collocation points
nodes (each with two 3.8G CPUs) of our 64-node Linux cluster were utilized to
solve these deterministic equations. These are dual core processors with hyper-
threading capabilities thus each node was used to perform the computation for
4 such problems. The total computational time was about 210 minutes. Each
maxi=1:nno (| f − I( f )|), where nno is the number of nodes). As the level of interpo-
lation increases, the number of sampling points used to construct the stochastic
solution increases [14]. Notice that there is a slight jump in the error going
from an interpolation of depth 2 to depth 3. This is probably due to the pres-
120
ence of some highly localized fluctuations of the stochastic solution that is cap-
tured only when the depth of interpolation reaches 3. Nevertheless, the error
reduction shown above follows the theoretical convergence estimates for using
Smolyak based interpolation [14].
e
b c d
a g
The mean temperature is shown in Fig. 3.17. The figure plots iso-surfaces of
temperatures −0.25 (Fig. 3.17(b)), 0.0 (Fig. 3.17(c)) and 0.25 (Fig. 3.17(d)). The
figure also shows temperature slices at three different locations of the xz plane:
y = 0 (Fig. 3.17(e)), y = 8µm (Fig. 3.17(f)) and y = 16µm (Fig. 3.17(g)).
121
6
1
d
0
-0.2 0 0.2
b c Temperature
a
f
the given finite input data set into a set of low-d points. In case the input data,
122
{xi } all belong to a localized region of MS 2 , the model reduction strategy frame-
work will construct a low-dimensional parametrization of only this localized
region. By increasing the amount of data utilized, one can make sure that the
complete space, MS 2 is sampled, ensuring that the data-driven model reflects
3.7 Conclusions
ability has been developed in the present work. The major advantages of
the proposed developments are: it seamlessly meshes with any reconstruction
method, directly converts samples into an equiprobable low-order model of
the property, and is applicable to any property variation (for instance, property
123
how uncertainty in the topology of the microstructure affects the evolution of a
dependant variable (temperature).
The basic model reduction ideas envisioned in this work are not limited to
generation of viable stochastic input models of property variations. This frame-
work has direct applicability to problems where working in high-dimensional
duction strategy formulated here. This is an area that is unexplored and could
potentially result in very efficient, real time, data-driven, stochastic reduced-
order model generation techniques. In addition to the importance of such mod-
els in process modeling of heterogeneous materials (polycrystals, composites,
124
CHAPTER 4
ANALYSIS OF FLOW THROUGH HETEROGENEOUS RANDOM MEDIA:
A STOCHASTIC MULTISCALE SYSTEM
4.1 Introduction
putational techniques have been developed that solve for a coarse-scale so-
lution by defining an appropriate coarse-scale problem that captures the ef-
fect of the fine-scales [94]. The more popular techniques developed for such
upscaling fall under the category of multiscale methods viz. the variational
multiscale (VMS) method (operator upscaling) [95, 96, 97, 98, 99, 100], the
heterogenous multiscale method [101, 102] and the multiscale finite element
method [103, 104, 105, 106, 107]. Further related techniques are presented
in [108, 109]. These computationally scalable techniques have resulted in the
development of black box simulators that have been used with significant suc-
125
cess in the solution of large scale transport problems in complex geometries.
The multiscale analysis of such systems inherently assumes that the com-
plete, fine-scale variation of the permeability is known. This assumption lim-
its the applicability of these frameworks since it is usually not possible to ex-
perimentally determine the complete structure of the media at the finest scale.
In the past decade, there has been tremendous progress in posing and
solving SPDEs. Several techniques like Generalized Polynomial Chaos expan-
sion [7, 8, 11, 12, 37, 39, 41], wavelet expansion and collocation based strate-
gies [14, 16, 110, 111] have been developed to solve SPDEs. These techniques
have been applied with significant success to solve single scale stochastic prob-
lems. In recent years, there has been significant interest in coupling determin-
istic multiscale methods with techniques in stochastic analysis to investigate
There are two key questions that have to be sequentially answered to con-
struct a stochastic multiscale framework that models transport phenomena only
given limited information: (1) Techniques to convert/encode limited statistical
information and/or sampled property values into a viable (stochastic) input
model for the (multiscale) permeability variation and (2) Given this input mul-
126
tiscale stochastic model, techniques to solve the stochastic multiscale equations.
The basic idea to solve the stochastic multiscale set of equations is to ex-
tend deterministic multiscale methods to their stochastic analogues. Spectral
strategies to pose and solve stochastic multiscale problems have been investi-
gated by Xu [112] and Asokan and Zabaras [42]. Collocation based strategies
have recently been developed in [20]. The key is to define appropriate ways to
link the fine-scale (subgrid) stochastic variation with the coarse-scale stochastic
variation of the dependent variables.
Fig. 4.1 shows a schematic of the problem of interest. Denote the domain as
D ⊂ Rnsd , where n sd is the number of spatial dimensions. The characteristic
tion, k(x) in the domain. However, the complete permeability distribution in un-
127
D
known. Only some limited statistics and/or snapshots of the permeability are
given. This limited information available to characterize the permeability neces-
sitates assuming that the permeability is a realization of a random field. This is
mathematically stated as follows: Let Ω be the space of all allowable permeabil-
ity variations. This is our event space. Every point k = {k(x, ω), ∀x ∈ D, ω ∈ Ω}
in this space is equiprobable. Consequently, we can define a σ-algebra F and
a corresponding probability measure P : F → [0, 1] to construct a complete
probability space (Ω, F , P) of allowable permeability. To make this abstract de-
128
where Y1 , . . . , YN are uncorrelated random variables.
The pressure and velocity are characterized by the following set of equations
is also assumed to not have a multiscale character. These assumptions are sim-
ply to make the subsequent developments clear. It is fairly straightforward to
account for uncertainties in f as well as to analyze the effects of a multiscale
f [113].
while taking into account the fine-scale variation in the stochastic permeability.
In the next section, we detail a stochastic extension to the variational multiscale
method to solve this problem. The stochastic multiscale formulation is based
on the multiscale formulation detailed in the papers by Juanes et al. [113] and
tions. For brevity of presentation, we will concentrate on these two aspects. The
interested reader in referred to [113, 114] for other details.
129
4.3 Stochastic variational multiscale formulation
For the problem to be physically relevant, we assume that the stochastic perme-
ability satisfies some conditions, particularly:
Assumption 1: k is positive and uniformly coercive:
The governing equations for the velocity and pressure given in the mixed
form are as follows:
k−1 u + ∇p = 0, (4.5)
∇ · u = f. (4.6)
The above constitutive and conservation equations are supplemented with the
following boundary conditions
p = po on ∂D p , (4.7)
u · n = uo on ∂Du . (4.8)
Without loss of generality, we assume that the boundary conditions are deter-
ministic and that the Neumann condition is homogeneous [113], uo = 0 on ∂Du .
The next step is to introduce the appropriate function spaces in which the
velocity and pressure lie. In contrast to their deterministic counterparts, the ve-
130
locity and pressure defined here are stochastic processes. Most dependent vari-
ables one encounters in stochastic analysis are usually random processes [45,
115, 116, 117] that are defined in product spaces. These variables (for instance
p(x, Y)) usually have one structure in the stochastic space p(., Y) and another
in the physical space p(x, .). The numerical analysis/approximation [45, 115]
of such functions can be performed by defining the appropriate tensor product
spaces. The interested reader is referred to [45, 115] for insightful discussions
on the definitions of these product spaces. Following [45, 115], we define ap-
propriate function spaces that encode variations of the function in the physical
domain D and in the stochastic space Γ.
Remark 3.1: We will consider stochastic functions that belong to S, the space
of all square integrable functions, with measure ρ(Y). Given a deterministic
function space A, its stochastic counterpart is the tensor product space A ≡ S⊗A.
The problem defined by Eqs. (4.5) and (4.6) along with the boundary conditions
Eqs. (4.7) and (4.8) can be written in mixed variational form: Find (u, p) ∈ V×W
such that
R R
where h f, gi is defined as Γ
dρ(Y) ∂D p
f gdx.
131
Variational multiscale approach: The exact solution u is assumed to be
made up of contributions [98] from two different (spatial) scales namely, the
coarse-scale solution uc (x, .) that can be resolved using a coarse (spatial) mesh
and a subgrid solution u f (x, .):
u = uc + u f , p = pc + p f . (4.14)
This additive sum decomposition induces a similar decomposition for the spa-
tial part of the fine-scale tensor-product function spaces into a direct sum of a
W = Wc ⊕ W f , H = Hc ⊕ H f , V = Vc ⊕ V f . (4.15)
The main idea is to develop models for characterizing the effect of the subgrid
solution u f (x, .) on the coarse-scale solution and to subsequently derive a mod-
ified coarse-scale formulation that only involves uc (x, .). The additive decom-
position provides a way of splitting the fine-scale problem given by Eqs. (4.12)-
(4.13) into a coarse-scale problem and a subgrid problem. Testing against the
coarse-scale test functions results in the coarse-scale variational problem: Find
Similarly testing against the subgrid test functions results in the subgrid varia-
(w f , ∇ · (uc + u f )) = (w f , f ), ∀w f ∈ W f . (4.19)
Note that both the coarse-scale variational equations Eqs. (4.16)-(4.17) as well
as the subgrid variational equations Eqs. (4.18)-(4.19) contain coarse and sub-
grid variations in the dependent variable ( uc , pc ; u f , p f ). The key is to solve
132
Eqs. (4.18)-(4.19) for u f and construct a functional representation of the subgrid
variation, u f and p f in terms of the coarse-scale variation, uc :
scale are first defined. As stated before (Remark 3.1), the stochastic function
spaces are simply tensor product spaces of S with the corresponding determin-
istic spaces.
Fine-scale approximation spaces: Let Vh and Wh be finite-dimensional subspaces
W h = S ⊗ Wh , V h = S ⊗ V h . (4.21)
Note that Wh and Vh should satisfy the discrete inf-sup condition [118]. Consider
SNh
a partition, Th of the domain D into non-overlapping elements ei , Th = i=1 ei ,
where Nh is the number of elements of the grid. Following [113], define also the
S Mh
skeleton of the partition, SPh = a=1 γa , where Mh is the number of element faces
denoted by γa . The partition Th is denoted as the fine-grid, on which the fine-
133
113] space, RT 0 (Th ), and the deterministic finite element approximation space
for the fine scale pressure, Wh is taken to be the space of piece-wise constants on
the fine mesh, P0 (Th ).
ministic developments, the two choices for such spaces were the lowest order
Raviart-Thomas space (used in [113]) and the Brezzi-Douglas-Marini space of
order 1 (used in [96]). We choose to utilize the stochastic analog to the low-
est order Raviart-Thomas space in the current work. We can now associate the
Mc
X
Vch =S⊗ Vch, Vch = {uc : uc = N ca uca , uca = 0 ∀Λa ∈ ∂Du }, (4.22)
a=1
134
mesh, P0 (Tc ):
Nc
X
Wch =S⊗ Wch , Wch = {wc : wc = φci wci }, (4.23)
i=1
which is equal to one in element E i and zero in all other elements. wci is the cor-
responding pressure degree of freedom (the average pressure in coarse-element
E i ).
coarse-element [113]. Denote by E i,h = Th (E i ) the fine grid defined over the coarse
element E i . The subgrid-scale velocity field defined on each coarse element sat-
isfies the condition
Similarly, the subgrid pressure space is restricted to belong to the space of piece-
wise constant functions on each coarse element
135
where Whf,i is the finite-dimensional approximation to the subgrid-scale con-
tinuum space W f , constrained to a coarse element E i . The elements of these
spaces are extended to zero functions to the entire spatial domain [113], D, and
the subgrid pressure space is defined as
Nc
M
Whf = Whf,i . (4.29)
i=1
also assume that this condition is satisfied not in a distributed sense, but for each
stochastic realization (i.e. point-wise in stochastic space). This leads to the following
condition
where (., .)Ei is the corresponding inner product defined over the coarse-element
E i . This condition (in its strong, point-wise form) is equivalent to [113]
∇ · uc = Πc f, (4.31)
where Πc f is the projection of the source/sink term onto the space Wc of piece-
wise constants on the coarse grid. As stated in the beginning of this section, we
assume that the sourse/sink function does not display a multiscale nature, i.e.
136
it is equal to its projection on the space of coarse-scale pressures ( f = Πc f ). Sub-
stituting Eq. (4.30) into the coarse-scale conservation equation Eq. (4.17) results
in
Since wc is constant in each coarse element, using the divergence theorem and
the assumption that this condition is valid point-wise in stochastic space, results
Eq. (4.33) is the essential condition [113] that guarantees mass conservation at
both scales and allows for the localization of the subgrid problem. This local-
ization assumption converts the global subgrid problem into a set of local Neu-
mann problems with boundary conditions defined by Eq. (4.33). Furthermore,
the subgrid test function v f must satisfy v f · n = 0 on ∂E i which results in the
following constraint
(∇ · v f , pc )Ei = 0, ∀E i ∈ Tc . (4.34)
Finally, we define the space of subgrid pressure Whf as the orthogonal comple-
ment of Wch in Wh (i.e. (w f , wc ) = 0). Now, since WCh = div Vch , this results in an
additional constraint
(∇ · vc , p f )Ei = 0, ∀E i ∈ Tc . (4.35)
The subgrid problem: Using the localization assumption Eqs. (4.33), (4.34), (4.35)
the localized subgrid problem can be defined as follows: For each coarse ele-
ment, E i = 1, . . . , Nc , find (u f , p f ) ∈ Vhf,i × Whf,i , such that
137
(v f , k−1 u f )Ei − (∇ · v f , p f )Ei = −(v f , k−1 uc )Ei , ∀v f ∈ Vhf,i , (4.36)
element.
The above equations transform the problem of finding the function Φ(uc ) and
Ψ(uc ) (defined in Eq. (4.20)) into finding the subgrid scale basis functions (N af , φaf )
associated with each coarse-scale degree of freedom. For computational effi-
ciency, the multiscale basis function representing the multiscale velocity is de-
138
fined as the sum of the coarse-scale and subgrid-scale components as
Mc
X Mc
X
c f c
u = (N a + N a )ua = N ms c
a ua . (4.39)
a=1 a=1
The multiscale basis functions are associated with each coarse-scale inter-
face Λa . The multiscale basis functions are constructed based on defining the
Greens function to the subgrid Eqs. (4.36)-(4.37) [98, 99]. In the context of de-
terministic mixed multiscale methods, various authors have constructed these
multiscale basis functions [96, 106, 107]. We utilize the deterministic definition
of the multiscale basis functions used in [120, 113] to define the stochastic mul-
tiscale basis functions. The multiscale basis function is the solution to a flow
problem restricted to a pair of adjacent coarse elements that share a common
coarse interface. The course terms are specified in such a way that the flow
through the interface is identically one (based on the Green’s function idea). It
is straightforward to extend this definition of the multiscale basis function to
the case when the permeability is stochastic.
is shared by the coarse elements E i and E j ) are the solution to the following
k−1 Nms ms
a + ∇φa = 0, in E i ∪ E j , (4.40)
N ms
a · n = 0, on ∂(E i ∪ E j ), (4.41)
R
if x ∈ E i ,
+k(x, .)/ Ei
k(x, .)dx,
∇ · Nms = (4.42)
a R
−k(x, .)/ k(x, .)dx, if x ∈ E j .
Ej
139
Vch × Wch , such that
1
R
where φ̃ms ms
a = φa − |Ei | Ei
φms
a dx.
Once the coarse-scale variables are solved, the fine-scale velocity and pres-
sure can be reconstructed (if necessary) by resorting to the additive decompo-
sition Eq. (4.14). The subgrid part is obtained by a linear combination of the
coarse-scale fluxes and the subgrid scale basis functions.
140
Set coarse and fine discretization
• Find the stochastic multiscale basis functions for each coarse element in-
terface by solving Eqs. (4.40)-(4.42).
• Use these multiscale basis functions to solve for the stochastic coarse-scale
velocities and pressure given by Eqs. (4.47)-(4.48). Adaptive techniques to
solve the above set of SPDEs are discussed in Section 3.5.3.
Fig. 4.3. The data-driven strategies (discussed in Chapter. 3) convert the limited
information available into a plausible, realistic stochastic representation of the
fine-scale permeability variation. The adaptive sparse grid collocation strategy
141
is used to construct the stochastic multiscale basis functions over all the coarse
elements. Following this, the stochastic coarse-scale equations are solved for the
coarse-scale stochastic pressure and velocity.
In this section, we apply the complete stochastic analysis framework- from data-
driven model generation, the stochastic variational mixed multiscale formula-
tion to the adaptive solution of the resulting SPDEs. In the first example, we
domain.
142
4.5.1 Effect of uncertainties at different scales
The schematic of the domain of interest is shown in Fig. 4.4. The region is part
of an outcrop (where extensive measurements of permeability have been per-
formed [124]) in Lawyer Canyon in Texas. The domain of interest is a square
of length 200 ft. Flow is driven by an injection well at the left bottom corner
of the domain and a production well at the top right corner. There is a low-
permeability fault running across the domain. Limited information about the
spatial variation in the fine-scale permeability is available in the form of semi-
variograms and mean permeability (Fig. 4.5). Furthermore, the exact location
and characteristics of the fault are unknown. Hence both the fine-scale perme-
ability variation as well as the large-scale fault features have to be considered to
be stochastic.
Fault
200 ft
Figure 4.4: Schematic of the domain. A fault runs across the middle of the
domain.
143
Effect of stochastic fine-scale permeability: We first investigate the effect
of the stochastic fine-scale permeability keeping the location of the fault fixed.
The fault is assumed to originate at (40, 100) and have a length of 100 feet and a
width of 20 feet. A reduced model for the fine-scale permeability is constructed
from the limited statistics available. Since in this case, the correlation structure
of the permeability variation is available, we utilize the KL expansion to con-
struct a viable stochastic model.
Note the logarithmic scale used. The first 25 eigenvalues represent about 95% of
the total information. The KL expansion is truncated to 25 terms and this gen-
erates a 25-dimensional input stochastic model for the fine-scale permeability.
Three eigenvectors utilized in the KL expansion are shown in Fig. 4.7. Note
144
3
0.45 10
0.44
0.43
0.42
Semivariogram, V(r)
Eigenvalue
0.41
102
0.4
0.39
0.38
0.37
1
10
0.36
0.35
0 100 200 300 400 0 10 20 30 40
Distance, r Number of eigenvectors
50 50 50
0 0 0
0 50 100 150 200 0 50 100 150 200 0 50 100 150 200
Fig. 4.8 shows two extreme realizations of the log-permeability field con-
structed utilizing the KL model. Note the close to 7 orders of magnitude varia-
145
Figure 4.8: Two extreme realizations of the log-permeability generated
from the KL expansion.
coarse variables.
ing number of collocation points, while Fig. 4.9 (right) plots the convergence of
the stochastic x-direction flux with increasing number of collocation points. For
the same depth of interpolation about 2.9 × 109 collocation points would have
been required using conventional sparse grid collocation.
Fig. 4.10 plots the mean values of the coarse-scale pressure and x-direction
flux. Note that the pressure is stratified around the region where the permeabil-
ity fault is located.
Fig. 4.11 plots the standard deviation of the stochastic coarse-scale pres-
sure and x-direction flux. The low-permeability fault leads in preferential flow
around the fault. This leads to stratification of pressure around the fault as well
as negligible flow. This is clearly seen as the very small standard deviation in
146
1
10
Mean p
Mean vel
-1
Var pp
Std
10 Var
Std vel
0
10
Error
Error
-2
10
-1
10
-3
10
-2
10
2
10
3
10
4 10 2 3 4
10 10 10
Number of collocation points Number of collocation points
Figure 4.9: Convergence of the stochastic solution (mean and standard de-
viation) with increasing number of collocation points. (left)
Stochastic coarse-scale pressure, (right) Stochastic coarse-scale
x-direction flux.
2.6
1.5 25.3
0.3 20.8
0.0 16.4
-0.6 11.9
-1.8 7.5
-2.9 3.0
-4.1 -1.5
-5.2 -5.9
-6.4 -10.4
-7.6 -14.9
pressure in the middle of the domain where the fault is located. Furthermore,
this results in a large variation in flux around the permeability fault as can be
seen in Fig. 4.11 (right).
147
0.064 3.299
0.052 2.480
0.039 1.661
0.027 0.842
0.015 0.547
0.003 0.117
gies. 106 samples were utilized to compute the statistics using the Monte Carlo
method. This took about 900 hours using 30 nodes of our in-house Linux clus-
ter. Fig. 4.12 plots the difference between the mean values of the coarse-scale
2.80E-03 2.52E-02
2.35E-03 1.84E-02
1.90E-03 1.16E-02
1.46E-03 4.85E-03
1.01E-03 -1.95E-03
5.64E-04 -8.74E-03
1.18E-04 -1.55E-02
-3.29E-04 -2.23E-02
Figure 4.12: Difference in the mean value of the stochastic solutions using
the adaptive sparse grid strategy and Monte Carlo sampling.
(left) Mean pressure, (right) Mean x-direction flux.
Notice the difference in the mean pressure is very close to zero close to the
injection well as well as on the permeability fault. On the other hand, since
the flow will preferentially try to avoid the fault (resulting in the largest devi-
148
ation around this region), there will be high variability in the x-direction flux
immediately above and below the permeability fault. This leads to the largest
error occurring here as seen clearly in Fig. 4.12. This is also clearly illustrated
in Fig. 4.13 which plots the difference between the standard deviation of the
coarse-scale pressure and x-direction flux computed from the two methods.
5.53E-04 1.12E-01
1.77E-04 7.51E-02
-1.99E-04 3.81E-02
-5.75E-04 1.09E-03
-9.51E-04 -3.59E-02
-1.33E-03 -7.30E-02
-1.70E-03 -1.10E-01
-2.08E-03 -1.47E-01
Fig. 4.14 plots streamlines based on the mean coarse-scale velocity distribu-
tion in the domain. The streamlines preferentially circumvent the permeability
barrier.
(180, 160). The PDFs are plotted in Fig. 4.15. PDFs of coarse-scale pressure (not
shown) also exhibited similar trends. The fine-scale stochastic permeability es-
sentially caused a ‘smearing’ of the coarse-scale pressure and velocity fluxes.
149
Figure 4.14: Streamlines drawn on the mean stochastic coarse-scale veloc-
ity.
0.6 0.7
Probability distribution function
0.6
0.5
0.5
0.4
0.4
0.3
0.3
0.2
0.2
0.1
0.1
0
3 4 5 6 7 0
1 2 3 4 5
Flux (x-direction) Flux (x-direction)
coarse-scale stochastic pressure and velocity flux. We set the fine-scale perme-
ability variation to its mean value and utilize the same 100-fold coarse-grained
discretization as in the previous problem. We assume that there is 10% uncer-
tainty in the location, length and width of the permeability barrier. This is quan-
tified by 4 independent uniform random variables, (Y1 , Y2 , Y3 , Y4 ), on the domain
150
[−1, 1] as
where (xl , yl ) is the start position of the barrier, and Lbarrier , Wbarrier are the
length and width of the barrier, respectively. The variational stochastic mixed
multiscale framework is utilized to solve this problem. 40 nodes of our local
Linux cluster (corresponding to 160 processors) is utilized to solve the problem.
Mean p 10
1 Mean vel
10-1 Var
Std p Var
Std vel
Error
Error
-2
10 10
0
-3
10
-1
10
500 2000 500 2000
Number of collocation points Number of collocation points
Fig. 4.16 (left) plots the convergence of the stochastic pressure with increas-
ing number of collocation points, while Fig. 4.16 (right) plots the convergence
of the stochastic x-direction flux with increasing number of collocation points.
151
For the same depth of interpolation about 8000 collocation points would have
been required using conventional sparse grid collocation. The adaptive sparse
grid strategy sampled the 4D stochastic space anisotropically with the random
dimension corresponding to the x-location of the permeability barrier being the
Fig. 4.17 plots the mean contours of the stochastic coarse-scale pressure and
flux in the x-direction. The effects of the fine-scale variation in the structure of
permeability is visible in the uneven pressure contours near the injection and
production wells. Furthermore, the pressure is essentially uniform close to the
middle of the domain where the permeability barrier lies. This is also seen in
Fig. 4.17 (right), where most of the flow circumvents the location of the fault.
This is even more clearly seen in the streamlines drawn on the mean coarse-
scale velocity field as shown in Fig. 4.18.
Fig. 4.19 plots the standard deviation contours of the coarse-scale stochastic
pressure and x-direction flux. As expected, most of the deviation in the pressure
occurs close to the region of the permeability fault. Interestingly, most of the
152
Figure 4.18: Streamlines drawn on the mean stochastic coarse-scale veloc-
ity.
right.
153
0.8
2
0.4
1
0.2 0.5
0 0
9 10 11 12 13 -2 -1.5 -1 -0.5 0
Flux (x-direction) Flux (x-direction)
ond point (100, 60) has a smaller standard deviation. The PDFs are plotted in
Fig. 4.20. Notice that the PDFs at both points have several peaks (in contrast
with the previous case) pointing to the possibility of mode shifts existing. In
particular, Fig. 4.20 (right) is the PDF of a point that lies very close to the perme-
ability barrier and it clearly shows up to 10 distinct peaks. It appears that small
variations in the location of the peak results in significant changes in the flow
patterns leading to the multi-modal structure of the PDF.
variations and discontinuities. Most of the refinement proceeded along the di-
mension that characterized the uncertainty in the x-location of the permeability
154
22
20
18
16
Depth of interpolation
14
12
10
8
6
4
2
0
0 0.25 0.5 0.75 1
Random variable
Figure 4.21: The refinement of the adaptive sparse grid. Each row of points
represents the refinement at a particular level of interpolation.
The interpolation starts at the bottom.
barrier (i.e. Y1 ). We solved a problem with this as the only input stochastic vari-
able. Fig. 4.21 plots the evolution of the adaptive grid. Notice that after the 8-th
level of interpolation, most of the refinement is concentrated at 5 points. This
implies the definite existence of steep gradients or discontinuities around these
points in stochastic space. This is clearly seen in Fig. 4.22, where the PDF of the
coarse-scale flux in the x-direction is plotted at two spatial points. The PDF in
both cases consist of 5 sharp peaks.
In Fig. 4.23, we illustrate this mode shift behavior by plotting two real-
izations of the coarse-scale stochastic flux (x-direction). These two realiza-
tions correspond to two points in stochastic space that are separated by 1%
(Y1 = 0.295, 0.305). This corresponds to a mode-shift located at Y1 = 0.30 as
seen in Fig. 4.21. The flow structure close to the permeability barrier is signifi-
cantly different for the two cases.
155
10.00 450
9.00 400
0.00 0
5.00 5.50 6.00 6.50 7.00 0.08 0.09 0.1 0.11 0.12 0.13
Flux (x-direction) Flux (x-direction)
Figure 4.23: (left and center) Two realizations of the coarse-scale flux (x-
direction) for a 1% perturbation in the x-location of the perme-
ability barrier, (right) Difference between the two solutions.
156
domain is shown in Fig. 4.24. The region consists of a 250×150 ft2 domain where
the fine-scale permeability is given in 1 ft2 blocks. An injection well is located
at the bottom left corner of the domain while a production well is placed at the
top right corner of the domain. No flow boundary conditions are enforced on
all the four boundaries. The fine-scale discretization of the domain corresponds
to a 250 × 150 element triangulation.
Uncertain permeability
block
150 ft
250 ft
The fine-scale permeability in the domain is shown in Fig. 4.25. The deter-
ministic permeability is taken from the 10-th SPE comparative project [125]. The
permeability shows variation of 5 orders of magnitude. Furthermore, there is
a localized region of uncertain fine-scale permeability in the central portion of
the domain, corresponding to a 50 × 50 ft2 domain. Aggressive coarse-scaling is
157
Figure 4.25: The fine-scale log permeability distribution in the domain.
The axes are scaled (1 unit = 50 ft).
the fine-scale permeability in the domain, and (b) if a large number of snapshots
are available, a non-linear model reduction strategy is utilized.
meability distributions and the eigenvalue problem solved. Fig. 4.26 plots the
eigen-spectrum of the covariance matrix. The first six eigenmodes accurately
represent the variability in the given data-set. These eigen-modes are utilized to
construct the corresponding low-dimensional representation of the stochastic-
fine-scale permeability distribution in the central block of the domain. The di-
mensionality of the representation is 6.
158
2.5
1.5
Eigenvalue
1
0.5
0
1 2 3 4 5 6 7 8 9 10
Number of eigenvectors
Figure 4.26: The eigen-spectrum of the data set. The first 6 eigen-modes
represent most of the variation that the input data set exhibits.
vergence of the stochastic pressure (mean and standard deviation) with increas-
ing number of collocation points, while Fig. 4.27 (right) plots the convergence
of the stochastic x-direction flux with increasing number of collocation points.
-2 -2
10 10
Mean vel
Mean p Std vel
Var vel
Var
Std p
-3
10
Error
Error
-3
10
-4
10
-5 -4
10 1 2 3 10 1 2 3
10 10 10 10 10 10
Number of collocation points Number of collocation points
Fig. 4.28 plots the mean contours of the stochastic coarse-scale pressure and
159
flux in the x-direction. Note the highly dendrite-like structure of the x-direction
flux. Due to the large variation in the permeability, the flow tries to find the path
with the largest permeability resulting in this dendrite-like structure.
-0.7 90.
-1.8 69.
-2.8 49.
-3.5 29.
-3.8 8.
-4.2 4.
-5.2 -6.
-6.3 -26.
-7.3
Fig. 4.29 plots the standard deviation contours of the coarse-scale stochastic
pressure and x-direction flux. Upstream of the uncertain permeability block, the
standard deviation of the coarse-scale pressure and velocity are both negligible.
Most of the deviation in the pressure occurs in the uncertain domain with its ef-
fects felt downstream of the uncertain domain. The velocity flux has maximum
deviation in the region of uncertainty and the deviation rapidly decays in the
rest of the domain.
0.88
3.5E-04
0.75
3.0E-04
0.63
2.4E-04
0.50
1.9E-04
0.38
1.3E-04
0.29
7.4E-05
0.21
1.9E-05
0.08
0.00
160
first point (50, 50) is upstream of the localized uncertain region, while the second
point (125, 75) is in the region of uncertain fine-scale permeability. The PDFs are
plotted in Figs. 4.30 and 4.31. For the point upstream, the probability distribu-
tions are peaked and have a very small range of variation. In comparison, the
2500
60
Probability distribution function
40
1500
30
1000
20
500
10
0 0
-2.459 -2.4585 -2.458 -2.4575 12.25 12.275 12.3 12.325
Pressure Flux (x-direction)
40 4
Probability distribution function
3.5
30 3
2.5
20 2
1.5
10 1
0.5
0
-3.925 -3.9 -3.875 -3.85 -3.825 2 2.5 3
Pressure Flux (x-direction)
161
Large number of snapshots
In this case, a data-set containing 1500 plausible fine-scale permeability distri-
butions in the uncertain part of the domain is available. Based on the discus-
sion in Chapter. 3, the neighborhood graph G is constructed using these 1500
the dataset (see [21]). Fig. 4.32 plots the variation in the length functional of the
MST. The optimal dimensionality was estimated to be d = 5. The corresponding
low-dimensional representation of the stochastic-fine-scale permeability distri-
bution in the central block of the domain is constructed.
4
10
Length of the MST
3
10
2
10 1 2 3
10 10 10
Number of samples
Figure 4.32: Plot of the length functional of the MST of the graph G for
various sample sizes. The intrinsic dimension is related to the
slope of the graph.
162
(stochastic + spatial domain) dimensional problem.
Fig. 4.33 plots the mean contours of the stochastic coarse-scale pressure and
flux in the x-direction. The mean results are very similar to the results obtained
by generating a data-driven model utilizing a much lower amount of initial
information.
-0.7 90.
-1.8 69.
-2.8 49.
-3.5 29.
-3.8 8.
-4.2 4.
-5.2 -6.
-6.3 -26.
-7.3
On the other hand, Fig. 4.34 plots the standard deviation contours of the
coarse-scale stochastic pressure and x-direction flux. The utilization in more
information (larger number of snapshots to build the input model) results in an
overall reduction in the standard deviation in the domain, though the general
2.4E-04
0.43
2.1E-04
0.38
1.7E-04
0.32
1.3E-04
0.26
9.0E-05
0.21
5.1E-05
0.15
1.3E-05
0.09
0.04
0.00
163
1500 40
30
1000
20
500
10
0
-2.459 -2.4585 -2.458 -2.4575 12.25 12.275 12.3
Pressure Flux (x-direction)
2
Probability distribution function
20
1.5
1
10
0.5
0
-3.925 -3.9 -3.875 -3.85 2 2.5 3
Pressure Flux (x-direction)
164
coarse-scale pressure and x-direction flux is plotted at the same two spatial lo-
cations. The PDFs are plotted in Figs. 4.35 and 4.36. For the point upstream,
the probability distributions are peaked and have a very small range of varia-
tion. In comparison, the point in the region of uncertainty has a significantly
larger range of variability. Comparing Fig. 4.36 with Fig. 4.31 illustrates the
effect of the incorporation of more data into the construction of the stochastic
input model.
165
CHAPTER 5
CONCLUSIONS AND SUGGESTIONS FOR THE FUTURE RESEARCH
put models of the permeability. The resulting multiscale SPDEs are solved uti-
lizing an adaptive sparse grid collocation strategy. We showcase the complete
formulation through realistic large scale applications of flow through hetero-
geneous random media. This is (to the best knowledge of the authors) the first
Although this thesis has addressed the fundamental issues with (a) solving
stochastic partial differential equations in a computationally tractable form (b)
generating reliable stochastic input models from limited experimental data, and
166
5.1 Stochastic inverse problems
167
5.2 Incorporating correlation statistics and investigating regu-
larization
In classical inverse problems [126, 127, 128, 129, 130, 131], the ill-posednes re-
quires some form of regularization or smoothing assumptions for computing
the solution. But in the context of stochastic inverse problems, we are search-
ing for a solution in a distribution sense in a much larger space (tensor product
space of spatio-temporal and stochastic variations). This expanded search space
essentially guarantees a stochastic solution [137]. Nevertheless, regularization
arguments play a very significant role in reducing the computational effort by
168
(m x1 + 1) × (m x2 + 1) control points Pi, j as
m x1 X
X m x2
q(x1, x2, Y) = Pi, j (Y)Bmi x1 (x1)Bmj x2 (x2), where, (5.1)
i=0 j=0
mx !
Bmi x (x) = xi (1 − x)mx −i . (5.2)
i!(m x − i)!
The design variables are now the stochastic control values Pi, j (Y). These design
variables can be subsequently represented in terms of nq collocated values as
nq
X
Pi, j (Y) = Pri, j (Y r )Lr (Y). (5.3)
r=1
The sensitivity equations are now defined in terms of these nq deterministic pa-
rameters. Since the choice of the parametrization controls the spatial smooth-
ness of the solution, it is necessary to construct convergence estimates as the
spatial parametrization is varied. Some preliminary work on convergence of
relation kernel C(x) determines the stochastic field and imposes some smooth-
ness on its spatial variability. This will convert the problem of designing the
stochastic field q to the problem of designing its correlation structure C(x). The
unknown correlation function is represented in terms of its spectral expansion
K
X
C(r) = ai sin(ix/L), (5.4)
i=1
where L is the characteristic length of the system and {ai } is the set of K
Fourier modes of the correlation function (and we have assumed that the field
is isotropic, without loss of generality). By representing the correlation this way,
169
the stochastic optimization problem is converted to computing the modal coef-
ficients {ai }. The design parameters are now the set of K scalars {ai }. Given any
correlation kernel, the Karhunen-Loève expansion can be utilized to explicitly
represent the stochastic field as
N p
X
q(x, ω) = E(q(x, ω)) + λi fi (x)Yi (ω), (5.5)
i=0
where (λi , fi (x)) are the eigen-pairs of the correlation function. The stochastic
sensitivity equations are formulated in terms of these K deterministic scalars.
The choice of the number of modes K used in the representation of the unknown
correlation provides a measure of the regularization. Including higher frequen-
cost functional is important along with developing a strategy for estimating the
number of modes required based on the statistics of the measurements.
170
5.3 Stochastic low-dimensional modeling
sentation can either be created off-line or adaptively modified on-the-fly [133, 134, 135].
dent variable into stochastic spatial modes. Consider a random field u(x, t; Y) in
a space-time domain and we look for biorthogonal representations as
K
X
u(x, t; Y) = b(h) (h)
i (t)ai (x; Y). (5.6)
i=1
stochastic field (either from off-line calculations, available numerical data or ex-
suggested that this stochastic counterpart of the POD framework will produce
low-order ODEs (in terms of the temporal coefficients bi(h) (t)) that are indepen-
171
dent of the stochastic dimensionality of the system. It is important that care
be taken in choosing the appropriate norm over which to construct the spatial
stochastic basis. As a preliminary step, one could utilize inner products based
on the mean, variance and standard deviation [139]:
Z Z Z
{b, c}0 := hbi · hcidx, {b, c}1 := hb · cidx {b, c}2 := (hb · ci − hbi · hci) dx.(5.7)
D D D
The unknowns are the deterministic coefficients {bi }. Inserting this representa-
tion, Eq. (5.9) into Eqs. (2.3) and using the orthogonality properties of the modes
results in a set of ODEs for the modal coefficients, {bi }.
K
X
(B( b(h) (h) (h)
i (t)ai : {q, α}), b j ) = 0, j = 1, ..., K, (5.10)
i=1
∂bhj
= F j (bhj ), j = 1, ..., K. (5.11)
∂t
The cost functional, its gradient and the stochastic optimization problem can
now be posed in terms of this set of ODEs instead of the set of SPDEs.
172
There has been recent progress on developing reduced order models to control
systems defined by PDEs [133, 134, 135]. In such approaches, an adaptive pro-
cedure is used that successively updates the ensemble of data snapshots and
the reduced order model during the optimization procedure. It would be help-
173
BIBLIOGRAPHY
174
[12] D. Xiu, G. E. Karniadakis, Modeling uncertainty in flow simulations via
generalized polynomial chaos, J. Comp. Physics 187 (2003) 137–167.
[16] D. Xiu, J. S. Hesthaven, High order collocation methods for the differential
equation with random inputs, SIAM J. Sci. Comput. 27 (2005) 1118–1139.
175
[24] V. deSilva, J. B. Tenenbaum, Global versus local methods in nonlinear di-
mensionality reduction, Advances in Neural Information Processing Sys-
tems 15 (2003) 721–728.
[28] J. A. Costa, A. O. Hero, Geodesic Entropic Graphs for Dimension and En-
tropy Estimation in Manifold Learning, IEEE Trans. on Signal Processing
52 (2004) 2210–2221.
176
[36] C. L. Winter, D. M. Tartakovsky, A. Guadagnini, Moment equations for
flow in highly heterogeneous porous media. Surv. Geophys., 24 (2003) 81-
106.
177
[46] D. Xiu,, D. Lucor, C.-H. Su, G. E. Karniadakis, Performance Evaluation
of Generalized Polynomial Chaos, International Conference on Computa-
tional Science, Lecture Notes in Computer Science, Vol. 2660, pp. 346-354,
Springer, 2003.
[49] E. Novak, K. Ritter, The curse of dimension and a universal method for
numerical integration, Multivariate approximation and splines, G. Nurn-
berger, J. W. Schmidt, G. Walz (eds.), 1997, 177–188.
178
[58] A. Klimke, Sparse Grid Interpolation Toolbox – User’s Guide, IANS re-
port 2006/001, University of Stuttgart, 2006.
[59] http://mpdc.mae.cornell.edu/Software.
[60] Y. Shen, P. Tong, K. -Q. Xia, Turbulent convection over rough surfaces,
Phys. Rev. Let. 76 (1996) 908–911.
[69] L. Arleth, S. Marc̆elja. T. Zemb, Gaussian random fields with two level-
cuts– Model for asymmetric microemulsions with nonzero spontaneous
curvature, The Journal of Chemical Physics 115 (8) (2001) 3923–3936.
[70] http://www.mathworks.com/products/matlab/.
179
[71] F. F. Abraham, J. Q. Broughton, N. Bernstein, E. Kaxiras, Spanning the
Length Scales in Dynamic Simulation, Computers In Physics 12 (1998)
538–546.
180
[82] C. Goffman, G. Pedrick, First course in functional analysis, Prentice-Hall,
2002.
[83] J. Breadwood, J.H. Halton, J.M. Hamersley, The shortest path through
many points, Proc. Cambridge Philos. Soc. 55 (1959) 299–327.
[85] C-G. Li, J. Guo, G. Chen, X-F. Nie, Z. Yamg, A version of Isomap with ex-
plicit mapping, Proceedings of the Fifth International conference on ma-
chine learning and cybernetics, Dalian, Aug 2006.
181
[95] T. Arbogast, S. E. Minkoff, P. T. Keenan, An operator-based approach
to upscaling the pressure equation, Computational Methods in Water
Resources XII, Computational Methods in Contamination and Remedi-
ation of Water Resources Vol. 1, Computational Mechanics Publications,
Southampton, UK (1998).
[103] T. Y. Hou, X. H. Wu, A multiscale finite element method for elliptic prob-
lems in composite materials and porous media, J. Comput. Phys. 134
(1997) 169-189.
182
[106] Y. Efendiev, V. Ginting, T. Y. Hou, R. Ewing, Accurate multiscale finite
element methods for two-phase flow simulations, J. Comput. Phys., sub-
mitted.
[107] J.E. Aarnes, V. Kippe, K.A. Lie, Mixed multiscale finite elements and
streamline methods for reservior simulation of large geomodels, Adv. wa-
ter. Resources. 28 (2005) 257–271.
[115] Ch. Schwab, R. A. Todur, Sparse finite elements for stochastic elliptic
problems–higher moments, Computing 71 (1) (2003) 43–63.
[116] Ch. Schwab, R. A. Todur, Sparse finite elements for elliptic problems with
stochastic loading, Numerische Mathematik 95 (2001) 707–734.
[117] T. Petersdorff, Ch. Schwab, Sparse finite element methods for operator
equations with stochastic data, Applications of Mathematics 51 (2006)
145–180.
183
[118] F. Brezzi, M. Fortin, Mixed and hybrid finite element methods, Springer-
Verlag 1991
[119] P.A.Raviart, J.M. Thomas, A mixed finite element method for second
order elliptic problems, in: Mathematical aspects of the finite element
method, Lecture Notes in Mathematics 606, Springer-Verlag, Berlin 1977
[124] J.W. Jennings, S.l. Bryant, Outcrop permeability statistics and implication
for fluid flow and reactive transport modeling, Bureau of Economic Geol-
ogy, U. Texas Austin, 2000 report.
[125] M.A. Christie, M.J. Blunt, Tenth SPE comparative solution project: A com-
parison of upscaling techniques, SPE Reserv. Eval. Eng. 4 (2001) 308–317.
184
[131] V. A. Morozov, Methods for Solving Incorrectly Posed Problems.
Springer-Verlag, New York, 1984.
[135] K. Ito, S.S. Ravindran, Reduced Basis Method for Optimal Control of Un-
steady Viscous Flows, International Journal of Computational Fluid Dy-
namics 15 (2001) 97–113.
185