forked from valdal/latex-template-dissertation-ucm
-
Notifications
You must be signed in to change notification settings - Fork 1
/
Copy path1_introduction.tex
214 lines (202 loc) · 17.3 KB
/
1_introduction.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
% INTRODUCTION
Marine aggregates are randomly formed particles composed of organic and inorganic matter, such as phytoplankton, detritus, sediment, and fecal pellets~\cite{jackson_simulation_1989}. Since these components stick together arbitrarily, a marine aggregate is often shown as a fractal structure.
These marine aggregates take in carbon dioxide (CO$_2$) during photosynthesis near the surface ocean and carry the dissolved carbon to the deep ocean.
% The ocean absorbs $40 \%$ of the anthropogenically produced carbon dioxide (CO$_2$) from the atmosphere~\cite{omand_sinking_2020}.
This process removes CO$_2$ from the atmospheric carbon cycle~\cite{honjo_understanding_2014} and plays a role in regulating atmospheric CO$_2$ and climate changes, which is one of the most significant environmental problems we face today.
\par
It has been observed that larger sinking aggregates tend to accumulate in thin layers where the ambient fluid is density stratified~\cite{alldredge_occurrence_2002}. The highly concentrated thin layers become biological hotspots for bacterial activity and animal feeding. It is ecologically important to understand the marine aggregates' dynamics as they delay in settling.
\par
We consider the flows around aggregates in the low Reynolds number regime, which is applicable for approximately the smallest 30\% of marine aggregates~\cite{alldredge_situ_1988}. Flow in this regime is governed by the Stokes equations, which allows solutions to be found via boundary integral methods~\cite{pozrikidis_boundary_1992}. Such methods have been implemented in several contexts, typically via a combination of analytical integration near the singularities that arise and quadrature methods away from the singularities. See~\cite{pozrikidis_boundary_1992} for a review. Boundary integral methods
are particularly well suited to computations of flow around complex solids~\cite{zinchenko_boundary-integral_2006} as velocity and forces may be expressed in terms of an integral over the boundary of the object.
They have also recently been combined with other methods, for example, to capture interactions between fluids and elastic solids~\cite{bao_immersed_2017} and stochastic fluctuations in suspensions~\cite{bao_fluctuating_2018}.
Here, we form aggregates using cubic particles, which results in a simple boundary over which the resulting integrals can be computed analytically, as described in Chapter 2.
This has the advantage of avoiding numerical singularities on the boundary, which otherwise require regularization~\cite{cortez_method_2001} or specialized numerical methods such as high-order product Nystr\"{o}m methods~\cite{atkinson_numerical_1997, delves_computational_1985}, adaptive sub-domain integration~\cite{chan_second-order_1992},
or nearest-neighbor discretization of the regularized Stokeslet boundary integral equation~\cite{smith_nearest-neighbour_2018}.
We introduce a novel implementation of boundary integral methods which is both simple and well-behaved numerically.
Details of this new methodology, along with its validation, are presented in Section~\ref{sec:numerical}. We compare two approaches, using a single-layer and double-layer integral representation of the velocity~\cite{pozrikidis_boundary_1992, power_second_1987, ingber_comparison_1999}, and determine which is the most suitable for fluid flow simulations around marine aggregates with this new method.
\begin{figure}[ht]
\begin{center}
\includegraphics[scale=0.3]{figures/fig_sample_cube10.pdf}
\caption{Example aggregate model with ten cubes. We denote $S$ as the aggregate boundary surface and $\hat{n}$ as its normal. The vectors $\vec{x}$ and $\vec{y}$ are points on and outside of $S$.}
\label{fig_cube10}
\end{center}
\end{figure}
\par
Our research is centered on computing the velocity field around settling marine aggregates and the forces acting on them. We do so using boundary integral equation (BIE) formulations. We construct marine aggregates using cubes to capture their fractal shape, as shown in Figure~\ref{fig_cube10}.
\section{Fluid momentum equations}
To describe the incompressible fluid motion that takes place around marine aggregates, we consider the Navier-Stokes equations,
\begin{align}
\nabla \cdot \vec{u} = 0
\label{eq_conserv_mass} \\
\rho
\left(
\frac{\partial \vec{u}}{\partial t} + \vec{u}\cdot \nabla \vec{u}
\right)
= \nabla \cdot \bar{\bar{\sigma}} + \rho \vec{g} ,
\label{eq_momentum_NS}
\end{align}
where $ \rho$ is fluid density (constant) and $\vec{u}, \ \vec{g}$ are fluid velocity and gravity vectors, respectively.
The first equation~(\ref{eq_conserv_mass}) shows the conservation of mass, and equation~(\ref{eq_momentum_NS}) describes the momentum conservation.
We also introduce the stress tensor, $\bar{\bar{\sigma}}$ as sum of the fluid pressure $P$ and deviatoric stress tensor $\bm{\tau}$,
\begin{equation}
\bar{\bar{\sigma}} =-P \bar{\bar{I}} + {\bm \tau} = -P \bar{\bar{I}} + 2{\tilde{\mu}} {\bm D},
\label{eq_stress_tensor}
\end{equation}
where ${\tilde{\mu}}$ is constant fluid viscosity and ${\bm D}$ is the symmetric strain rate tensor, that is,
\begin{equation}
\boldsymbol{D} = \frac{1}{2} \left( \nabla \vec{u} + {(\nabla \vec{u} )}^T \right).
\label{eq_strain_rate}
\end{equation}
Since seawater is a Newtonian fluid, following Newton's law of viscosity, the strain rate~(\ref{eq_strain_rate}) becomes $\nabla \vec{u}$. This implies that we can rewrite the momentum equation~(\ref{eq_momentum_NS}) as
\begin{equation}
\rho \left(
\frac{\partial \vec{u}}{\partial t} + \vec{u}\cdot \nabla \vec{u}
\right)
= -\nabla P + {\tilde{\mu}} \nabla^2 \vec{u}+ \rho \vec{g}
\label{eq_stokes_momentum}
\end{equation}
For typical seawater, it is reasonable to say $\rho \approx 1025 \text{kg/m}^3$ and ${\tilde{\mu}} = 1.2 \times 10^{-3}\text{kg}/\text{ms}$.
Also, the gravity vector is $\vec{g} = - g\hat{k} \approx -9.8$m/$s^2 \times (0,0,1)$ where $\hat{k}$ points in the vertical direction (upward).
\par
The momentum equation~(\ref{eq_momentum_NS}) can be linearized for flows where inertial effects are small.
To estimate the size of the main forces at play, we consider a radius of marine aggregate, $R_a \approx 5 \times 10^{-5} (\text{m})$ and the reference Stokes settling speed of an aggregate,
\begin{equation}
U_s = \frac{gR_a^2}{{\tilde{\mu}}} (\rho_a-\rho) \approx 3.8 \times 10^{-4} ({\text{m/s}}),
\label{eq_U_s}
\end{equation}
where $\rho_a \approx 1400\text{kg/m}^3$ is the aggregate mass density.
When we non-dimensionalize the momentum equation using the length scale, $R_a$, and velocity, $U_s$, we obtain the following equation:
\begin{equation}
\left(\frac{\rho U_s R_a}{{\tilde{\mu}}} \right)
\left(
\frac{\partial \vec{u}'}{\partial t'} + \vec{u}'\cdot \nabla' \vec{u}'
\right)
= {\nabla'}^2 \vec{u}' - \nabla' P' + \vec{g}',
\label{eq_NS_moment_noD}
\end{equation}
where we find and compute the Reynolds number (Re),
\begin{equation}
\text{Re} = \frac{\rho U_s R_a}{{\tilde{\mu}}} \approx 10^{-2}
\ll 1.
\label{eq_Re}
\end{equation}
Note that we use the prime symbol to represent a dimensionless value.
Since we have a fairly small Reynolds number, we may neglect the inertial effects, limiting the left-hand side of equation~(\ref{eq_NS_moment_noD}) to zero,
\begin{equation}
{\nabla'}^2 \vec{u}' - \nabla' P' + \vec{g}' = 0.
\end{equation}
In the settling marine aggregate project part of this thesis, we, therefore, consider the following Stokes equations, written with dimensions, to describe the fluid flow around the settling aggregates,
\begin{align}
\nabla \cdot \vec{u} = 0
% \label{eq_conti2}
\nonumber \\
{\tilde{\mu}} \nabla^2 \vec{u} - \nabla P\ + \rho \vec{g} = 0.
\label{eq_stokes2}
\end{align}
The solutions of the system of equations are the fluid velocity, $\vec{u}$, and pressure, $P$. In general, pressure does not induce motion. We see the pressure at rest, denoted as $P_s$, contains the pressure of gravity, $P_s(\vec{y}) = \rho \vec{g} \cdot \vec{y}$, where $\vec{y} \in \mathbb{R}^3$ is a point in the fluid. Using this static pressure term, we introduce the dynamic pressure $P_d$, defined as
\begin{equation}
P_d(\vec{y}) = P(\vec{y}) - P_s(\vec{y}) = P(\vec{y}) - \rho \vec{g} \cdot \vec{y}.
\label{eq_def_Pd}
\end{equation}
By substituting the expression~(\ref{eq_def_Pd}), we obtain
\begin{align}
\nabla \cdot \vec{u} = 0
% \label{eq_conti2}
\nonumber \\
{\tilde{\mu}} \nabla^2 \vec{u} - \nabla P_d = 0.
\label{eq_stokes3}
\end{align}
While an aggregate model settles, we focus on the velocity field and hydrodynamic forces around the marine aggregate model.
%===SECTION 2.2=========================================
\section{Boundary integral equation (BIE) formulations}
For our simulations, we consider a large fluid domain compared to the size of an aggregate, having zero fluid velocity at infinity. We also treat our aggregate surface as a solid. Although marine aggregates are porous, the solid boundary condition is reasonable to apply due to their low permeability.
This condition prevents flow through the aggregate, acting like a solid particle. For this reason, any flow inside of the aggregate is neglected in the remainder of this thesis.
\par
For a general surface S, the velocity $\vec{u}$ at a point $\vec{y}\in \mathbb{R}^3$ exterior to the surface may generally be expressed using the representation formula~\cite{pozrikidis_boundary_1992} when solving for the flow using the Stokes equations,
% around a solid object is expressed, using the stress vector, $\vec{f}$,
\begin{equation}
\vec{u}(\vec{y}) =
- \frac{1}{8 \pi {\tilde{\mu}}} \int_S \vec{f}(\vec{x}) \cdot \bar{\bar{G}}(\vec{x},\vec{y}) \ \text{d}S(\vec{x})
+ \frac{1}{8 \pi} \int_S
\vec{u}(\vec{x}) \cdot \bar{\bar{K}}(\vec{x},\vec{y})
\cdot \hat{n} ( \vec{x})
\ \text{d}S(\vec{x}),
\label{eq_BIE}
\end{equation}
% where $\vec{f}(\vec{x})$ is the stress vector that describes a point force at $\vec{x} \in S$.
where the integral is taken over points $\vec{x}$ on the surface $S$.
The kernel $\bar{\bar{G}}(\vec{x},\vec{y})$ is the Green's function of the Stokes equations at any point $\vec{y}$, that is the {\textit{Stokeslet}}, in the domain for a point-source located at $\vec{x}$,
\begin{align}
\bar{\bar{G}}(\vec{x},\vec{y}) =
\frac{\bar{\bar{I}}}{||\vec{x}-\vec{y}||} + \frac{(\vec{x}-\vec{y})(\vec{x}-\vec{y})}{||\vec{x}-\vec{y}||^3}.
\label{eq_stokeslet}
\end{align}
The kernel $\bar{\bar{K}}(\vec{x},\vec{y})$ is the stress tensor associated with this fundamental solution, named the {\textit{Stresslet}},
%
\begin{align}
\bar{\bar{K}}(\vec{x},\vec{y}) =
-6\frac{(\vec{x}-\vec{y})(\vec{x}-\vec{y}) (\vec{x}-\vec{y})}{||\vec{x}-\vec{y}||^5},
\label{eq_stresslet}
\end{align}
where $\| \cdot \|$ is the vector 2-norm.
The first integral distribution on the right-hand side of equation~(\ref{eq_BIE}) is called the \textit{single-layer potential}, and the second one is called the \textit{double-layer potential}.
To compute the velocity at a point on the surface $S$, i.e., $ \vec{x}_s \in S$, we use
\begin{equation}
\vec{u}(\vec{x}_s) = - \frac{1}{4 \pi {\tilde{\mu}}} \int_S \vec{f}(\vec{x}) \cdot \bar{\bar{G}}(\vec{x},\vec{x}_s) \ \text{d}S(\vec{x})
+ \frac{1}{4 \pi}
\int_S
\vec{u}(\vec{x}) \cdot \bar{\bar{K}}(\vec{x},\vec{x}_s)
\cdot \hat{n} ( \vec{x})
\ \text{d}S(\vec{x}).
\label{eq_BIE_onS}
\end{equation}
In general, $\vec{f}(\vec{x})$ is the unknown stress vector, or traction, on the surface, and $\vec{u}(\vec{x})$ is the, typically unknown, velocity on the surface $S$.
When $S$ is the boundary of a solid object, in our case, an aggregate, the general representation formula may be simplified in two different ways, depending on the fluid domain characteristics and the object's surface. We will introduce those two formulae in Chapter 2.
\section{Complex fluids}
\label{sec:intro_complex_fluid}
In the last part of this thesis, we will turn our attention to non-Newtonian fluids. This topic is added to my thesis as an extension of a summer internship at the Lawrence Berkeley National Lab in 2022 under the guidance of Dr. Ishan Srivastava.
\par
\vphantom{text}
\par
A broad class of complex fluids encompasses various materials of natural, industrial, and medical importance, such as granular materials, suspensions of micro-scale materials, and biological molecules.
Complex fluids show many interesting behaviors quite different than Newtonian fluids.
The major difference is that the viscosity (denoted $\tilde{\mu}$ in the Newtonian fluid case) is not constant anymore; rather, it is attributed to the shear rate $\dot{\gamma} = \left| {\boldsymbol{D}} \right|$~\cite{kundu_chapter_2016}.
Due to this varying viscosity, the constitutive behavior of non-Newtonian fluids is highly complex.
To understand their behaviors, we shall focus on the viscous stress tensor, $\bm \tau$, introduced in equation~(\ref{eq_stress_tensor}).
In particular, we explore four features of viscosity in complex fluids: 1) (shear) rate-dependent flow, 2) yield stress, 3) a second-order strain rate rheology, and 4) granular material flow.
Note that {\textit{rheology}} is the study of material deformation and flows as responses to applied stress, including those that exhibit co-existing two states.
\begin{figure}[h]
\begin{center}
\includegraphics[scale=0.2]{figures/fig_yield_stress_graph.pdf}
\end{center}
\caption{Classification of fluid types based on descriptions in~\cite{irgens_rheology_2014}. Solid lines show different types of non-Newtonian behavior with examples in the parenthesis. The red star describes yield stress. The dashed line represents Newton's law of viscosity.}
\label{fig_rheology}
\end{figure}
\par
Figure \ref{fig_rheology} demonstrates how the shear stress, $|\bm{\tau}|$, varies on the shear rate $\dot{\gamma}$, with an example.
Two types of rate-dependent flow, shear-thinning, and thickening, are plotted with the green and purple lines, respectively.
The viscosity of a shear-thinning fluid, such as shampoo or mayonnaise, decreases as the shear rate increases~\cite{singh_introduction_2013}. On the other hand, shear-thickening fluid viscosity gets higher under shear strain~\cite{barnes_introduction_1989}.
Meanwhile, the orange straight line represents the Bingham plastic fluid.
One defining characteristic of many complex fluids is the presence of yield stress, which is shown as the red star on the vertical axis.
Yield stress is an instantaneous stress value to make a material flow.
For an insufficiently stressed state, the material behaves like an elastic solid, allowing deformation plastically; once the yield stress is exceeded, it flows like a fluid~\cite{irgens_rheology_2014}.
Examples of yield stress fluids include toothpaste, cement, mud, and blood.
We will discuss more of the relationship between viscosity and shear rate in Chapters 4.2 and 4.3.
\par
While the viscous stress tensor of the rate-dependent fluids has a linear relationship with the strain rate, $\bm D$, other flow effects can be captured by a higher-order strain rate.
In this thesis, we explore up to the second-order case and call them secondary flows.
In the recent paper by~\cite{tanner_review_2018}, Tanner discussed several examples of non-colloidal suspensions and highlighted the importance of examining these secondary flows.
For instance, a curvature in free-surface flows~\cite{couturier_suspensions_2011}, anomalous stress profile in cylindrical Couette flows~\cite{krishnaraj_dilation-driven_2016}, or negative rod climbing effect (Weissenberg effect)~\cite{boyer_dense_2011} are found for a non-linear strain rate effect. We thus want to implement computational tools to study the second-order strain rate effect.
\par
Lastly, we present further work for granular materials.
They are highly ubiquitous and the second most common material on earth after water~\cite{richard_slow_2005}.
For example, we can find grains in the food industry, pharmaceutical powder to make medicine tablets, and geological materials, such as sand or soil.
One fascinating property is that they can behave like a liquid, solid, and gas, coexisting simultaneously. For instance, sand in a desert can stay still like a solid without any external forces; it can also flow or slide like a fluid when there is wind.
There are still many open questions we should study to understand the behaviors of granular materials due to the complicated dynamics.
% and their flow in a long cube-shaped channel.
\par
Our ultimate goal for this project is to have a constitutive rheology code that can capture complex continuum dynamics.
We use the exascale computing package \href{https://amrex-codes.github.io/index.html}{{\color{blue}AMReX-Codes}}, which is a block-Structured Adaptive Mesh Refinement (AMR) Software Framework and Applications. We implement the granular rheology in an individual module of AMReX called \href{https://amrex-codes.github.io/incflo/}{{\color{blue}\textit{incflo}}}. This code allows us to model the incompressible Navier-Stokes equations with various types of rheology in complex geometries without subcycling in time.
We will describe the details of the numerical methods in Chapter 4.4.
\par
This dissertation is ordered as follows: in Chapter 2, we introduces the simulations of settling marine aggregates in a homogeneous fluid. As an extended study, we explore a density-stratified fluid to understand more realistic marine aggregate behaviors in Chapter 3. Lastly, Chapter 4 addresses complex fluid rheology and granular materials using the AMReX-Codes.