+ All Categories
Home > Documents > Modeling, Analysis, and Simulation of Adsorption in ...

Modeling, Analysis, and Simulation of Adsorption in ...

Date post: 08-Nov-2021
Category:
Upload: others
View: 6 times
Download: 0 times
Share this document with a friend
239
Clemson University TigerPrints All Dissertations Dissertations 8-2016 Modeling, Analysis, and Simulation of Adsorption in Functionalized Memebranes Anastasia Bridner Wilson Clemson University Follow this and additional works at: hps://tigerprints.clemson.edu/all_dissertations is Dissertation is brought to you for free and open access by the Dissertations at TigerPrints. It has been accepted for inclusion in All Dissertations by an authorized administrator of TigerPrints. For more information, please contact [email protected]. Recommended Citation Wilson, Anastasia Bridner, "Modeling, Analysis, and Simulation of Adsorption in Functionalized Memebranes" (2016). All Dissertations. 1687. hps://tigerprints.clemson.edu/all_dissertations/1687
Transcript
Page 1: Modeling, Analysis, and Simulation of Adsorption in ...

Clemson UniversityTigerPrints

All Dissertations Dissertations

8-2016

Modeling, Analysis, and Simulation of Adsorptionin Functionalized MemebranesAnastasia Bridner WilsonClemson University

Follow this and additional works at: https://tigerprints.clemson.edu/all_dissertations

This Dissertation is brought to you for free and open access by the Dissertations at TigerPrints. It has been accepted for inclusion in All Dissertations byan authorized administrator of TigerPrints. For more information, please contact [email protected].

Recommended CitationWilson, Anastasia Bridner, "Modeling, Analysis, and Simulation of Adsorption in Functionalized Memebranes" (2016). AllDissertations. 1687.https://tigerprints.clemson.edu/all_dissertations/1687

Page 2: Modeling, Analysis, and Simulation of Adsorption in ...

Modeling, Analysis, and Simulation of Adsorption inFunctionalized Membranes

A Dissertation

Presented to

the Graduate School of

Clemson University

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

Mathematical Sciences

by

Anastasia Bridner Wilson

August 2016

Accepted by:

Dr. Lea Jenkins, Committee Chair

Dr. Chris Cox

Dr. Vince Ervin

Dr. Scott Husson

Page 3: Modeling, Analysis, and Simulation of Adsorption in ...

Abstract

The emergence of biopharmaceuticals, and particularly therapeutic proteins, as a

leading way to manage chronic diseases in humans has created a need for technologies that

deliver purified products efficiently and quickly. Towards this end, there has been a signif-

icant amount of research on development of porous membranes used in chromatographic

bioseparations. In this work, we focus on high-capacity multimodal membranes developed

by Husson and colleagues in the Department of Chemical and Biomolecular Engineering at

Clemson University.

Chromatographic performance of such membranes, particularly the adsorption ca-

pabilities of the membranes, depends of a large number of variables making it unrealistic to

scan the available options and determine the conditions resulting in the best performance

experimentally. Consequently, the goal of this work is to develop a modeling framework ca-

pable of describing the process under continuous flow conditions and software tools capable

of simulating the protein chromatography process under the effect of complex adsorption

relationships.

In this work, we consider the reactive transport, or advection-diffusion-reaction,

problem to model the chromatography process. We focus on the case of highly advective

flows as one of the advantages of using membranes in chromatography is the capacity to

maintain high protein binding capacity at high flow rates. Toward this end, we utilize a

streamline upwind Petrov-Galerkin (SUPG) finite element method to numerically solve the

advection-dominated advection-diffusion-reaction equation for porous media.

The complicating feature of the problem arises from modeling the adsorption reac-

ii

Page 4: Modeling, Analysis, and Simulation of Adsorption in ...

tion. The most accurate, thermodynamically consistent model, or isotherm, for multimodal

adsorption, recently developed by Nfor and colleagues, is highly nonlinear and implicitly

defined. Even the next best model, Langmuir’s isotherm, while not implicitly defined is still

nonlinear. As such we develop and analyze discretization methods incorporating nonlinear,

potentially implicit, adsorption isotherm models.

To gain insight into the advection-diffusion-reaction problem, we begin by analyzing

the SUPG formulation for the steady state case of the advection-diffusion equation. We

also analyze the time-dependent linear cases incorporating constant and linear adsorption

models. Although the constant and linear adsorption models do not represent realistic

adsorption relationships, the linear analysis serves as a template for the nonlinear analysis.

When incorporating nonlinear adsorption, we consider two cases: adsorption with

an explicit representation as in Langmuir’s isotherm and adsorption with an implicit equa-

tion as in Nfor’s isotherm. In the case of an explicit adsorption relationship, three different

formulations are analyzed: a time-integrated mixed methods formulation, a time-integrated

SUPG formulation, and a fully implicit SUPG formulation. For the implicit adsorption rela-

tionship, a simple formulation is proposed which not only deals with the implicit definition

of the isotherm but also deals with the nonlinearity: the right hand side of the isotherm

relationship is evaluated at the previous time step. As expected, the solvability and stability

for this relationship are all shown to have a requirement on the time step size.

We provide numerical validation for each of the a priori error estimates. We also

compare results of our algorithm with data obtained from laboratory experiments. To im-

prove the accuracy of the numerical simulations, we incorporate non-instantaneous adsorp-

tion, considering both constant and transient adsorption rates. Additionally, we numerically

investigate the effects of varying velocity profiles by comparing results from simulations in-

volving five different profiles.

iii

Page 5: Modeling, Analysis, and Simulation of Adsorption in ...

Dedication

To my husband, Dustin, without whose continuous support over the years this work

would never have been finished.

iv

Page 6: Modeling, Analysis, and Simulation of Adsorption in ...

Acknowledgments

I would like to thank Dr. Lea Jenkins, my Ph.D. dissertation advisor, for the

extensive amount of time she has spent with me over the years helping me learn about

modeling and fluids in porous media, making me a better programmer, teaching me how to

write papers and present my research, thoroughly editing everything I wrote with her, and

advising me when I was looking for employment after graduation. Additionally I would like

to thank Dr. Scott Husson and Juan Wang for providing numerous experimental data sets

for comparison over the years and for answering my seemingly unending questions about

separations processes and membrane chromatography. I would also like to thank Dr. Vince

Ervin for all the help he has provided in checking the analysis portions of my research and

Dr. Chris Cox for serving on my committee with Drs. Husson and Ervin and providing

ideas for future collaborative research projects.

v

Page 7: Modeling, Analysis, and Simulation of Adsorption in ...

Table of Contents

Title Page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

Dedication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Review of Mathematical Methods . . . . . . . . . . . . . . . . . . . . . . . . 71.2 Structure of the Dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2 The Transport Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272.1 The Physics Involved . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272.2 The Reactive Transport Problem . . . . . . . . . . . . . . . . . . . . . . . . 40

3 SUPG Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423.1 Variational Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423.2 SUPG Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453.3 Finite Element Approximations . . . . . . . . . . . . . . . . . . . . . . . . . 483.4 Useful Inequalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4 Steady State Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 554.1 Finite Element Approximation & Assumptions . . . . . . . . . . . . . . . . 554.2 Boundedness and Coercivity of Bilinear Form . . . . . . . . . . . . . . . . . 564.3 Error Estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

5 Time-Dependent Linear Analysis . . . . . . . . . . . . . . . . . . . . . . . 665.1 Case 1: Constant Isotherm . . . . . . . . . . . . . . . . . . . . . . . . . . . 675.2 Case 2: Linear Isotherm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

6 Analysis for Nonlinear, Explicit Adsorption . . . . . . . . . . . . . . . . . 1236.1 Time-Integrated Mixed Method Formulation . . . . . . . . . . . . . . . . . 123

vi

Page 8: Modeling, Analysis, and Simulation of Adsorption in ...

6.2 Time-Integrated, SUPG Formulation . . . . . . . . . . . . . . . . . . . . . . 1486.3 Fully Implicit SUPG Formulation . . . . . . . . . . . . . . . . . . . . . . . . 154

7 Nonlinear Analysis with Implicit Adsorption . . . . . . . . . . . . . . . . 1687.1 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1687.2 Solvability and Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171

8 Numerical Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1808.1 Steady-State Convergence Rates . . . . . . . . . . . . . . . . . . . . . . . . 1818.2 Linear Convergence Rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1828.3 Nonlinear Convergence Rates . . . . . . . . . . . . . . . . . . . . . . . . . . 1848.4 Numerical Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

9 Summary and Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . 2109.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2109.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215

vii

Page 9: Modeling, Analysis, and Simulation of Adsorption in ...

List of Tables

8.1 Approximation errors and experimental convergence rates for the approxi-mation to the steady-state problem. As ∆t and h are cut in half, the H1

error is reduced the same amount which is consistent with the theoreticalconvergence rates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

8.2 Approximation errors and experimental convergence rates for the fully-explicitapproximation with a linear adsorption model. As ∆t and h are cut in half,the H1 error is reduced the same amount which is consistent with the theo-retical convergence rates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

8.3 Approximation errors and experimental convergence rates for the fully-implicitapproximation with a linear adsorption model. As ∆t and h are cut in half,the H2 error is reduced the same amount which is consistent with the theo-retical convergence rates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185

8.4 Time-integrated approximation errors and convergence rates for nonlinearadsorption with Langmuir adsorption model. As ∆t and h are cut in half,the time-integrated error is reduced the same amount which is consistentwith the theoretical convergence rates. . . . . . . . . . . . . . . . . . . . . . 186

8.5 Approximation errors and convergence rates for nonlinear adsorption withLangmuir adsorption model. As ∆t and h are cut in half, the H1 error isreduced the same amount which is consistent with the theoretical convergencerates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

8.6 Approximation errors and convergence rates for nonlinear adsorption withNfor adsorption model. As ∆t and h are cut in half, the H1 error for C andthe L2 error for q is reduced the same amount which is consistent with thetheoretical convergence rates. . . . . . . . . . . . . . . . . . . . . . . . . . . 189

viii

Page 10: Modeling, Analysis, and Simulation of Adsorption in ...

List of Figures

1.1 The protein chromatography process: the column materials separate the pro-teins as the solution is pushed through the column. . . . . . . . . . . . . . . 2

1.2 An electron microscopy image of a membrane used in membrane chromatog-raphy. Note the porous structure of the membrane that provides many ben-efits for chromatography. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

8.1 Comparison of numerical and experimental breakthrough curves using in-stantaneous adsorption. Note the typical S-shape characteristic and the closematch of breakthrough for both curves. . . . . . . . . . . . . . . . . . . . . 194

8.2 Effect of varying the mass transfer coefficient km on breakthrough curve.Notice as km increases, the breakthrough curves asymptotically approachthe results with instantaneous adsorption as shown in Figure 8.1. . . . . . . 197

8.3 Comparison of transient adsorption rate with constant adsorption rates. . . 1988.4 Comparison of numerical and experimental breakthrough curves using instan-

taneous adsorption and non-instantaneous adsorption. Note the numericalresult obtained using a transient adsorption rate are significantly closer tothe experimental results than those obtained using instantaneous adsorption. 199

8.5 The 2D evolution of the concentration in the membrane assuming a spatially-constant velocity profile. Notice that the membrane is essentially saturatedwith protein by t = 20 as shown in the last image. . . . . . . . . . . . . . . 204

8.6 The 2D evolution of the concentration in the membrane assuming a singleparabola velocity profile. Saturation of the membrane does not occur in thiscase until after the final image shown; specifically, it occurs at approximatelyt = 70. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205

8.7 The 2D evolution of the concentration in the membrane assuming a doubleparabola velocity profile. Saturation of the membrane does not occur in thiscase until after the final image shown; specifically, it occurs at approximatelyt = 65. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206

8.8 The 3D evolution of the concentration in the membrane assuming a singleparabola velocity profile. Saturation of the membrane does not occur in thiscase until after the final image shown; specifically, it occurs at approximatelyt = 72.5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207

8.9 The 3D evolution of the concentration in the membrane assuming a velocityprofile involving five parabolas. Saturation of the membrane does not oc-cur in this case until after the final image shown; specifically, it occurs atapproximately t = 80. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208

ix

Page 11: Modeling, Analysis, and Simulation of Adsorption in ...

8.10 Comparison of breakthrough curves for five different velocity profiles. Al-though some of the velocity profiles resulted in remarkably similar break-through curves (e.g. single and double parabola cases in 2D), there is enoughvariation in the curves to indicate further investigation into the velocity ifwarranted. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209

x

Page 12: Modeling, Analysis, and Simulation of Adsorption in ...

Chapter 1

Introduction

Biopharmaceuticals, and particularly therapeutic proteins and cells, are in high

demand for treating many severe and chronic illnesses such as diabetes, cancer, hemophilia,

anemia, and infectious diseases [101]. Currently, the U.S. Food and Drug Administration

has approved about 130 therapeutic proteins for use, and over 500 products are being

developed, 200 of which are cell therapeutics for transplant medicine, cancer treatment,

and aids therapy [1]. The global market for biophameceuticals is projected to exceed $142

billion in 2017 while the protein therapeutics market specifically is expected to grow 15%

annually through the next decade [72]. By the end of 2016, 50% of the top 100 drugs will

most likely be biologics, that is, drugs derived from biotechnology [44].

While the increase in product development is immensely beneficial to society, there

are significant challenges that must be addressed to ensure the supply meets the needs of

the population. In particular, manufacturers are striving to develop new production meth-

ods to increase production capacity while also minimizing the cost of production. Although

building new biomanufacturing facilities is an option for integrating new production meth-

ods, the significant financial risk of building and validating the new facilities often outweigh

the benefits of the new method. Consequently, technologies that could be integrated into

current biomanufacturing facilities would have the most economic benefit.

A typical biomanufacturing facility comprises upstream and downstream processes.

1

Page 13: Modeling, Analysis, and Simulation of Adsorption in ...

In the upstream process, biopharmaceuticals are produced by cell culture technologies us-

ing genetically engineered cells that excrete the product of interest into a solution. The

downstream process is designed to recover and purify the product from solution.

Downstream purification, normally including chromatography, often limits the pro-

duction capacity. In capture-step protein chromatography, the protein serum is passed

through a column containing materials that separate the target proteins (see Figure 1.1).

Once the materials in the column are saturated with protein, breakthrough occurs. Break-

through is evidenced by the outflow from the column containing a “significant” amount of

protein, for example 10% of the value of the feed concentration appears in the effluent of

the column for 10% breakthrough. After breakthrough occurs, an elution solution is pushed

through to remove the protein from the column.

Figure 1.1: The protein chromatography process: the column materials separate the proteinsas the solution is pushed through the column.

Current practice utilizes resin chromatography in several steps. In resin-based chro-

matography, small resin beads are used to isolate and purify the product or to remove

process impurities from the product solution usually through ion-exchange and diffusion-

based processes. Although reliable, it has low throughput, characterized by long processing

times that cause product degradation, with losses as high as 50% of the product manufac-

tured [139]. Its low efficiency is considered a “must be addressed” factor to avoid production

capacity constraints and decrease production cost in biopharmaceutical industries [99].

2

Page 14: Modeling, Analysis, and Simulation of Adsorption in ...

To address these challenges, membrane chromatography has been studied as an

alternative to resin chromatography [12, 13, 14, 33, 53, 107, 148, 149, 154, 158, 175, 176],

particularly now that strategies have been developed by Husson and colleagues and others

to increase the binding capacities of membrane adsorbers to values that meet or exceed

the corresponding resins [12, 13, 14, 33, 107, 148, 149]. Membrane chromatography uses

a porous, adsorptive membrane as the packing medium instead of small resin beads (see

Figure 1.2). Due to the near absence of diffusional mass-transfer limitations, adsorptive

membranes can maintain high protein binding capacity at high flow rates; on the contrary,

resin columns lose binding capacity at higher flow rates. Membranes therefore enable higher

productivity.

Figure 1.2: An electron microscopy image of a membrane used in membrane chromatog-raphy. Note the porous structure of the membrane that provides many benefits for chro-matography.

Protein binding capacity is a key performance metric in membrane chromatography

as it determines the volume of membrane needed for purification. Currently, commercial

membranes are limited to products in which protein capture is accomplished primarily by

one adsorption mechanism, e.g., ion exchange, hydrophobic interactions. Most ion-exchange

membranes lose protein binding capacity at relatively low conductivity and do a poor job of

removing unwanted protein aggregates. Thus, additional processing steps often are required

to reach high purity causing lower yield and higher costs [89, 170]. Recent research efforts

by Husson and colleagues have focused on multimodal membrane-based chromatography, in

which protein capture and aggregate removal are enhanced by coupling orthogonal mech-

anisms for adsorption [166]. Initial work produced multimodal membranes characterized

by high protein binding capacities at ionic strengths and used a thermodynamic model to

3

Page 15: Modeling, Analysis, and Simulation of Adsorption in ...

provide insights on the nature of protein-membrane interactions under different test condi-

tions.

Chromatographic performance by multimodal membranes depends on membrane

properties, protein properties, and operating conditions, e.g., protein concentration, pH,

ionic strength, salt type, flow rate. The number of variables makes it unrealistic to scan the

available options experimentally and discover the conditions that would result in the best

performance. Thus, to limit the number of experiments needed for process development,

it is imperative to develop a modeling framework capable of describing the process under

continuous flow conditions.

To model this process and thereby create a simulation tool, we consider the reac-

tive transport problem (advection-diffusion-reaction problem) given by the coupled set of

equations [154, 50, 5, 156]

ω∂C

∂t+ ρs (1− ω)

∂q

∂t+∇ · (uC)−∇ · (D∇C) = f

q = g (q, C)

where C is the liquid phase concentration of the material of interest, q is the solid phase

concentration of the material of interest, ω is the porosity of the membrane (0 ≤ ω ≤ 1),

ρs is the density of the membrane, u is the Darcy velocity, and D represents the diffusivity

of the fluid through the membrane. A more detailed description of the modeling equations

and the involved physics is given in Chapter 2.

The solid phase concentration is the concentration of material adsorbed onto the

membrane. The equation q = g (q, C) describes the (usually explicit and nonlinear) func-

tional relationship between the adsorptive properties of the membrane and the liquid phase

concentration C. Notable example relationships include the Langmuir and Freundlich

isotherm functions. However, recent developments in the understanding of these adsorp-

tion processes has led to newly defined, implicit representations of q as a function of C

[114, 116, 120, 167]. (See section 2.1.4 for more details on the different adsorption models.)

4

Page 16: Modeling, Analysis, and Simulation of Adsorption in ...

Etzel and colleagues were some of the first to mathematically model membrane

chromatography [154, 175]. In [154], Suen and Etzel first analyzed protein separation in

membranes due to affinity using a Langmuir model for adsorption. Affinity separations rely

on a very specific chemical reaction that causes the protein to bind to the membranes, or

more specifically to the ligands in the coating on the membranes. Etzel and colleagues were

able to use analytical solutions to their model to determine the effects of flow velocity and

membrane thickness and porosity on the performance of the membrane.

As affinity membranes use just one type of reaction to separate proteins, Yang, Bitzer

and Etzel continued analyzing membranes in the context of ion-exchange chromatography in

[175]. The authors investigated the performance of ion-exchange membranes by comparing

numerical results from different isotherm models to experimental results from literature.

Fitting the numerical solution to experimental results provided a means of determining

the phenomena that most affected the mass transfer rate and predicting the operating and

membrane-design parameters needed for sharp breakthrough curves.

The two most common isotherm models used in the literature are Langmuir’s and

Freundlich’s isotherms which were both proposed in the early nineteenth century. However,

both models have their disadvantages especially when it comes to predicting multimodal

adsorption. Additionally, neither model is able to accurately predict adsorption relation-

ships for varying salt concentrations: if the salt concentration of the solution changes, then

the fitted isotherm parameter values also change.

Consequently, much research and experimentation has been conducted in the past

decade to develop isotherm models capable of accurately predicting multimodal adsorption

under varying salt concentrations. In [67], Ghose et al. explored protein separations based on

hydrophobic tendencies of proteins in a liquid. Based on experimental results, the authors

developed an isotherm model to better predict the relationship between the amount of

protein in the liquid and the amount of protein bound to the membrane; the isotherm

model that was developed is an exponentially modified version of Langmuir’s isotherm.

While this isotherm accurately simulated adsorption in a range of circumstances, it was

5

Page 17: Modeling, Analysis, and Simulation of Adsorption in ...

developed in the context of one mode of adsorption.

In [65], Gao et al. investigated mixed-mode protein adsorption focusing on the effect

of salt concentration and pH levels. Based on experimental data, they proposed an empir-

ical modification to Langmuir’s isotherm which was similar to the model proposed in [67].

Although accurate in its predictions, the proposed isotherm was not thermodynamically

consistent.

Understanding the importance of having an accurate, thermodynamically-consistent

mixed-mode adsorption model, Mollerup and colleagues developed an adsorption model

based on thermodynamic principles [112, 113, 114, 115, 116]. In [112], Mollerup laid the

foundation of thermodynamic principles in the context of biopharmaceutical development,

specifically the area of applied thermodynamics of chromatography. Mollerup later discussed

thermodynamic principles of ligand binding in order to develop models for multi-component

adsorption due to ion-exchange and hydrophobic chromatography [113]. In [115], Mollerup

et al. focused specifically on modeling adsorption due to ion-exchange, describing a strategy

for estimating the parameters in the ion-exchange adsorption model. Mollerup and asso-

ciates summarized the development of their multimodal adsorption model in [114, 116] in

order to support an in-depth understanding of their model, both the development and im-

plementation. The authors emphasized that the parameters in their model are not only easy

to determine from experimental data but also have useful, well-defined physical meaning.

In [120], Nfor et al. applied Mollerup’s thermodynamic framework for adsorption to

the case hydrophobic interaction and ion-exchange chromatography. The resulting adsorp-

tion model gave extremely good fits with various sets of data found in literature and thus

can be used with any mixed mode adsorbent having hydrophobic and ion-exchange inter-

actions. Additionally, the adsorption model developed by Nfor et al. is applicable under

varying salt concentrations. As one of the known parameter values is the salt concentra-

tion of the solution, then changing the salt concentration of the solution changes only that

known value while the fitted parameters remain unchanged.

While the isotherm model developed by Nfor was very accurate for a wide variety of

6

Page 18: Modeling, Analysis, and Simulation of Adsorption in ...

data sets, it is by definition highly nonlinear and implicit in its definition making it difficult

to incorporate into numerical simulation. As a result, it has yet to be incorporated into a

full-model simulations of protein chromatography.

The goal of this work, therefore, is to develop software tools capable of simulating

the protein chromatography process under the effect of complex, implicit adsorption rela-

tionships in the presence of highly advective flows. Toward this end, we choose to utilize

a streamline upwind Petrov-Galerkin (SUPG) finite element method to numerically solve

the advection-dominated advection-diffusion-reaction equation for porous media. In this

work, we develop and analyze SUPG discretizations schemes for the cases of linear, non-

linear explicit, and nonlinear implicit adsorption relationships. We first review previous

mathematical analysis and numerical methods applied to the advection-diffusion-reaction

equation.

1.1 Review of Mathematical Methods

Extensive research has been conducted over the past decades mathematically study-

ing the general advection-diffusion equation. Despite this extensive research, designing, an-

alyzing, and implementing a scheme to solve the problem is still challenging especially in

the cases of advection dominated flow and nonlinear adsorption. Many times mixed meth-

ods are used in conjunction with the advection-diffusion equation because of the inherent

mass conservation properties of the mixed method. However, in the case of flow that is

dominated by advection, it is well known that numerical instabilities occur when standard

numerical methods are used (even mixed methods) because of sharp layers in the solutions

that are not resolved correctly (unless the spatial discretization parameter is taken to be

extremely small) [8, 17, 42, 70, 82]. While many numerical solutions have been proposed

to resolve these instabilities, two major families of methods have emerged: the streamline

upwinding Petrov Galerkin finite element method (SUPG) and the discontinuous Galerkin

finite element method (DG) [74]. These techniques often complicate the analysis of the

7

Page 19: Modeling, Analysis, and Simulation of Adsorption in ...

scheme so much of the analysis of the advection-diffusion-reaction problem is completed

without any stabilizing element involved or in a different context such as a mixed finite

element method.

The presence of a reaction terms, especially a nonlinear one, generally complicates

the analysis and numerical solution. The analysis is complicated even further by the consid-

eration of equilibrium adsorption instead of non-equilibrium (see [134] for a specific compar-

ison). The one case of reaction which actually simplifies the analysis is when the adsorption

is accounted for by a function of the unknown being added to the diffusive and reactive

terms as in

∂C

∂t−∇ · (D∇C) +∇ · (uC) +RC = f.

Since the reaction term RC does not have a temporal derivative associated with it, the

analysis is somewhat simpler. Also many times the reaction is taken to be linear, i.e. R is

not a function of C, to simplify the analysis and implementation even further.

Below we present a review of previous literature dealing with the advection-diffusion-

reaction equation and/or transport in porous media, specifically that which studies advection-

dominated flow. Since DG is one of the most common techniques for dealing with advection-

dominated flow, a brief summary of literature using DG is included; however since we have

chosen to employ the SUPG method, most of the review pertains to the SUPG method or

techniques inspired by and/or combined with upwinding.

1.1.1 Finite Element Analysis of Advection-Diffusion Equation

Extensive analysis is available for DG, SUPG, and mixed methods when applied

to the advection-diffusion-reaction equation. However, much of the analysis is conducted

under simplifying assumptions such as diffusion dominance and linear or no adsorption.

8

Page 20: Modeling, Analysis, and Simulation of Adsorption in ...

1.1.1.1 Discontinuous Galerkin Method

Just recently, Miller et al. provided a succinct summary of DG methods as applied

to water resource problems, but more extensive recent summaries are available in [40, 71,

143]. Specifically in [143], many different situations of elliptic and parabolic equations were

presented and analyzed. The linear parabolic problem given by the time-dependent diffusion

equation was analyzed for stability and error bound in the semi-discrete and fully-discrete

cases. For the transport equation, stability and error bounds were presented and proven

for the semi-discrete case while the fully-discrete case was studied numerically with a few

examples.

The fully discrete case was analyzed by Cockburn and Shu in [41]. The DG method

applied specifically to the transport problems in porous media was studied by Riviere and

Wheeler in [144]. In [82], Johnson presented the DG method for the advection-diffusion

equation. Some simple error estimates for the purely-diffusive problem with linear adsorp-

tion were presented.

In [36], Cockburn, et al. analyzed a new formulation for the stationary advection-

diffusion-reaction problem combining both discontinuous Galerkin and mixed method tech-

niques which they referred to as the local discontinuous Galerkin-hybridizable (LDG-H)

method. This unified method was first introduced in [38] in the context of a purely diffu-

sive problem (no advection or reaction terms) and then analyzed in this same situation in

[36] and [39]. The LDG-H methods cannot be applied to the purely hyperbolic case (when

there is no diffusive term) so [37] was restricted to the diffusion-dominated case. Although

convergence properties for this hybrid DG method were proven in [36] and [39] in the con-

text of the purely diffusive problem, [38] presented only numerical tests of the convergence

properties of the scheme applied to the advection-diffusion-reaction problem. Egger and

Schobel completed an analysis using this method for the advection-diffusion problem a year

later in [52]. Nonlinearity in the advection-diffusion equation was considered in [173] with

error estimates of the semi-discrete DG method stated and proven.

9

Page 21: Modeling, Analysis, and Simulation of Adsorption in ...

1.1.1.2 Upwinding

Summaries and comparisons of different upwinding schemes for advection-dominated

problems can be found in [8] and [42]. In [8], Bank, Burgler, Fichtner, and Smith summa-

rized upwinding schemes for the advection-diffusion problem. All methods were compared

and contrasted to the standard FE Galerkin discretization. In [42], Ramon Codina presented

an overview of the several finite element methods for solving the advection-diffusion-reaction

equation, particularly the advection-dominated case, including the SUPG method.

Initially the idea of upwinding was applied in centered finite difference methods to

avoid numerical oscillations in the case of small diffusion [163]. As the initial idea added too

much artificial diffusion, an improvement was made by introducing this artificial diffusion

only along the streamlines [75, 76, 92]. One more improvement was added to this method

by viewing the streamline diffusion method in the context of a weighted residual method

[111]; this streamline upwinding Petrov-Galerkin weighted residual method is sometimes

referred to as the “standard” SUPG method [42, 77].

As a weighted residual method, the upwinded test function should be applied to all

the terms (including the temporal derivative terms). However as this contributes a skew-

symmetric component to the mass matrix, many authors suggested and implemented a

modification to the classical SUPG method in which only the stationary terms are upwinded

[34, 54, 82, 85]. The idea of upwinding was more recently used in conjunction with mixed

method schemes to provide stability to methods which are inherently mass-conservative (see

Upwind-Mixed & Hybrid Mixed Methods paragraph in section 1.1.1.3).

Some analysis of the SUPG method has been completed over the years, but much

of the time simplifying assumptions were made. In [20], Brooks and Hughes showed that

when implemented as a Petrov-Galerkin weighted residual method, the SUPG method does

not add the same amount of artificial diffusion as classical upwind methods. They reviewed

information known about the phase accuracy and damping for the algorithm from work

completed by Tezduyar in [157]. They also showed the accuracy of the SUPG method on

several numerical experiments of the linear advection-diffusion equation.

10

Page 22: Modeling, Analysis, and Simulation of Adsorption in ...

Hughes and Mallet analyzed the time-dependent advection-diffusion problem with

linear approximation elements in [78]. This case is significantly simpler than that considered

in this dissertation since the second-order term arising from upwinding the diffusive term

drops out with linear elements. Also, research has been conducted to determine the best

choice for the upwinding parameters in both the case of linear elements [43] and quadratic

elements [78].

In [81, 82, 84, 85], Johnson considered the general case of elements but presented

theoretical analysis (stability and error bounds) of the stationary situation of the linear hy-

perbolic equation given by the advection-diffusion-reaction equation with linear adsorption.

The linear adsorption, in this case given by the term C, simplified the analysis somewhat

by adding stability and enabling easier writing of the analysis in terms of the H1 norm. In

[119, 118], the fully discrete case of the advection-diffusion-reaction equation with the same

adsorption term as in [82, 85, 84] was analyzed. The case of variable coefficients was also

analyzed in [118].

Franca, Frey, and Hughes compared different stabilizing techniques for approxi-

mating the stationary advective-diffusive model including SUPG [62]. Stability and er-

ror bounds were proven assuming scalar diffusion, using a slightly modified test function,

v + δ(u · ∇v + d∆v), and were written in terms of a seminorm involving ‖∇C‖0 and

‖δ1/2u · ∇C‖0. The proven error bounds did not explicitly show the dependence on the

upwinding parameter, and no reaction term was considered.

In [106], Melenk and Schwab investigated the hp-stability and error bounds for

the stationary 1D advection-dominated problem with linear adsorption. In [74], Houston,

Schwab and Suli extended this analysis to the case of two dimensions for the purely advective

problem with simple linear adsorption. The 2D advection dominated diffusion problem with

linear adsorption was then analyzed by Gerdes et al. in [66]. All analysis was completed

based on a positivity assumption on the difference between the reaction and advection

parameters.

A new stabilization method for the linear scalar advection-diffusion problem with

11

Page 23: Modeling, Analysis, and Simulation of Adsorption in ...

no adsorption was proposed by Dutra do Carmo and Alvarez [51]. The new formulation was

based on the classical SUPG formulation, and depending on the regularity of the approxi-

mate solution, the new method degenerated into either the SUPG method or the Consistent

Approximate Upwind (CAU) method. The accuracy and stability of the new formulation

was demonstrated numerically. In [64], the authors analyzed the Consistent Approximate

Upwind (CAU) Petrov-Galerkin method for the steady-state advection dominated reaction-

diffusion problem and compared it to the SUPG method. A simple linear adsorption term

written as RC was included in the analysis. Although no new analysis was completed on the

SUPG method, stability and continuity of the bilinear form associated with the steady-state

equation was recalled from previously completed analysis.

In [15], Bochev, Gunzburger, and Shadid analyzed the transient advection-diffusion

problem. A classical SUPG spatial discretization and an implicit time integration tech-

nique were analyzed for stability over a range of parameters values showing that the SUPG

method provided stability for any parameter value even when combined with an implicit

time integration method.

In [34], Chrispell, Ervin and Jenkins analyzed a fractional step θ-method for the

scalar advection-diffusion problem with a linear reaction term using SUPG to stabilize.

The solvability of the system was proven along with O(δ) error bounds. Also, work was

done to compute the optimal θ for the method used, and numerical computations verifying

the theoretical results were presented.

In [23], Burman analyzed the stability and convergence of the time-dependent ad-

vection equation (no diffusion) using the “classical” Streamline Upwind Petrov-Galerkin

method spatial discretization and a few different A -stable finite difference discretization in

time. Note that since the classical SUPG method was used, even the temporal term was

upwinded. Stability and convergence bounds were proven for both the backward Euler and

the Crank-Nicholson temporal discretizations. Because the temporal terms was upwinded,

the stability and convergence rates were O(δ) (which is to be expected), but the bound

controlled only the discrete material derivative∥∥δ1/2 (∂t(C − Ch) + u · ∇(C − Ch))

∥∥, not

12

Page 24: Modeling, Analysis, and Simulation of Adsorption in ...

the expected weighted streamline derivative∥∥δ1/2u · ∇(C − Ch)

∥∥. Similar results for BDF2

were stated without proofs. Numerical results supporting the analysis were shown. No

reaction term was considered in any of the analysis or numerical experiments.

Most recently, John and Novo considered the evolutionary advection-diffusion-reaction

equation [80], analyzing the “classical” SUPG method. As with much of the previous anal-

ysis involving SUPG (e.g. [34, 51]), the reaction was a linear term written as RC which

simplified the analysis somewhat allowing more easily for terms to be hidden in the reaction

term.

1.1.1.3 Mixed Methods

Mixed methods for second order problems such as the transport equation have been

studied extensively throughout the years (see for example [87, 108, 132, 133, 137, 138, 140,

146, 162]). In [87], Douglas and Roberts obtained global error estimates for the two- and

three-dimensional mixed finite element method for the Dirichlet problem of the general

second order elliptic equation. Although a reaction term was considered, the coefficient

function only depended on space and therefore was linear in the unknown. In [162], Vohralık

proved a posteriori error estimates for the advection-diffusion-reaction equations using the

lowest-order mixed finite element discretization. The error estimates were proven assuming

the reaction term was linear on the triangulation.

Sacco and Saleri proposed a way to stabilize the advection-diffusion equation using

a new family of mixed finite volume methods in [146]. Their formulation of the advection-

diffusion equation was stationary but allowed for a nonlinear reaction term written simply

as RC. The reaction coefficient function R was permitted to be a function of C, but

it was assumed to be nonnegative and sufficiently regular. Also positivity assumptions

were made on the interaction of the reaction coefficient function with the velocity function.

The stabilization employed the lowest-order Raviart-Thomas finite element spaces plus a

suitable quadrature formula for the mass matrix. The quadrature formula was employed as

a lumping technique to deal with the fullness of the discrete matrix instead of the alternative

13

Page 25: Modeling, Analysis, and Simulation of Adsorption in ...

technique of hybridization of the mixed formulation.

In [108], Micheletti, Sacco and Saleri studied the reaction-diffusion problem. They

analyzed a stabilization method similar to the method in [146] applied to the advection-

diffusion problem, i.e. they used lowest order Raviart-Thomas finite element spaces with a

suitable quadrature formula for the mass matrix. Here they assumed simply the adsorption

coefficient function R was constant and nonnegative. Radu and Wang analyzed a nonlinear

parabolic equation (specifically Richards’ equation) in [138] and obtained an optimal order

of convergence which was supported by numerical experiments.

In [132], Radu and Pop applied a mixed finite element method to the nonlinear

reactive transport problem taking into account both advection and diffusion. Newton’s

method was used to resolve the nonlinearity due to the adsorption isotherm. The method

was first analyzed in the context of nonlinear equilibrium adsorption when the isotherm was

nondecreasing and the derivative of the isotherm was Lipschitz continuous (as in Langmuir’s

isotherm). Then the case of non-Lipschitz adsorption (as in Freundlich’s isotherm) was

analyzed. For the case of non-Lipschitz adsorption, they first applied a regularization to

the isotherm; this regularized isotherm, which approached the Freundlich isotherm as the

regularization parameter went to zero, was then nondecreasing with a Lipschitz continuous

derivative. In both cases, they give a sufficient condition for the quadratic convergence

of Newton’s method to be preserved. Numerical examples were included to verify the

theoretical results.

Radu and Pop continued their analysis of the nonlinear reactive transport problem

in [133]. They again used a mixed finite element method with a Newton linearization but

instead considered non-equilibrium adsorption. The authors considered only the case of a

Freundlich-type adsorption. The same regularization step as in [132] was made to allow for

the convergence analysis to be completed. A sufficient condition for quadratic convergence

with Newton’s method was given and proven, and numerical verification of the theoretical

results was provided.

14

Page 26: Modeling, Analysis, and Simulation of Adsorption in ...

Time-Integrated Mixed Methods When dealing with nonlinear adsorption in the

transport equation, it is helpful to look at the analysis of Richards’ equation since it can

be written in a form that has the same mixed collection of derivatives as the advection-

diffusion-reaction equation. One technique that has been especially helpful in the analysis

of Richards’ equation is a time-integrated technique first introduced by Nochetto and Verdi

in [121]. The method first integrates in time to eliminate any temporal derivatives which

allows for less regularity on the solution. Nochetto and Verdi proved stability and error

estimates for this method applied to the degenerate parabolic problem focusing specifically

on the two-phase Stefan problem.

An extension of this time-integrated method was made by Arbogast, Wheeler and

Zhang in [6]: they analyzed this method applied to a nonlinear, possibly degenerate,

advection-diffusion equation. However, no adsorption was considered. They assumed Lip-

schitz continuity on all the nonlinear functions involved, and they obtained nearly optimal

bounds in the sense that all bounding terms reduced to approximation error except one

which involved the difference of two discrete projection (an extremely small quantity when

using RT spaces).

Woodward and Dawson analyzed Richards’ equation using this time-integrated

method in [171]. They analyzed the method over two ranges of saturation corresponding to

the possibilities of nondegeneracy and degeneracy. Their analysis considered a nonlinearity

in the volumetric water content θ which they assumed to be Lipschitz continuous. They

also assumed the absolute permeability tensor K was Lipschitz continuous in the case of

partially saturated to saturated flow. They obtained optimal error results when considering

the lowest order RT spaces.

In [129] and [130], Radu, Pop and Knabner extend the analysis of [6] and [171] on

Richards’ equation to the situation of less regularity for the coefficient functions. Specif-

ically, the function k for the conductivity of the medium was considered to be bounded,

but not necessary Lipschitz continuous. For simplicity of the analysis, they considered ho-

mogeneous Dirichlet boundary conditions and analyzed only the situation of variably- to

15

Page 27: Modeling, Analysis, and Simulation of Adsorption in ...

fully-saturated regimes. Most importantly, they showed equivalence of the time-integrated

mixed method and the original mixed method approached for Richards’ equation.

Based on the theoretical results in [129], Pop, Radu and Knabner proposed an it-

erative scheme to solve nonlinear elliptic problems (including Richards’ equation). They

proved convergence of the scheme based on a contraction argument and showed numeri-

cal support of the theoretical results. In [147], Schneid, Knabner, and Radu extended the

work done in [129] to consider the possibly degenerate case of Richards’ equation. Also,

the homogeneous Dirichlet boundary conditions were relaxed somewhat to consider possi-

bly non-homogeneous Dirichlet conditions as long as the Dirichlet function was sufficiently

smooth. Pop, Radu, and Knabner continued the analysis of this method in [136] by consid-

ering less regularity on the diffusion function; instead of Lipschitz continuity, they assumed

only Holder continuity. This low level of regularity allowed for modeling the cases of both

fast diffusion and slow diffusion.

This time-integrated method was also applied to transport of reactive solutes in

porous media although not necessarily in the context of mixed methods. Barrett and Kn-

abner looked at both the case of non-equilibrium adsorption [9] and the case of equilibrium

adsorption [10] in porous media. For simplicity in the analysis, they assumed that the pro-

cess was not advection-dominated and therefore dropped out the advection term from the

equation. Also, they focused on a formulation for the equilibrium term of the adsorption

similar to the Freundlich isotherm as this isotherm degenerates at C = 0, i.e. the func-

tion is only Holder continuous, not Lipschitz continuous at C = 0. The time-integrated

method was used so that the lower regularity of the adsorption function would not hinder

the analysis.

In [134] and [135], Radu, et al. presented and analyzed the Euler implicit-mixed

finite element scheme based on the lowest order RT elements for the reactive transport

equation in porous media coupled with Richards’ equation to describe the flux. The time-

integration strategy was employed in this case again to allow for less regularity on the

function representing water content. They proved convergence for the case of equilibrium

16

Page 28: Modeling, Analysis, and Simulation of Adsorption in ...

adsorption where they assumed Holder continuity of the adsorption equation. No upwinding

or stabilization technique was considered for the case of advection-dominate flow.

Upwind-Mixed & Hybrid Mixed Methods More recently, methods combining both

the stabilizing effects of upwinding techniques and mass-conserving effects of mixed methods

have been introduced and analyzed for the advection-diffusion equation.

Following an idea initially suggested in [31], Dawson developed and analyzed a time-

splitting method for solving an advection-dominated parabolic equation in which a Godunov

procedure was used to approximate the advection and a mixed finite element discretization

was used to approximate the diffusion [46, 47]. In [46], error estimates for the 1D case

were presented along with numerical results for the original algorithm and a few variants of

the scheme. In [47], Dawson extended the ideas to multiple dimensions and derived error

estimates.

In [4], Arbogast and Wheeler developed a new finite element method which they

called the “characteristics-mixed method” to approximate the solution to the advection

dominated transport equation. The test functions were piecewise constant in space and

approximately follow the characteristics of advection in time. They assumed the diffusion

tensor was positive definite (precluding the case of degeneracy) and considered no adsorp-

tion. They proved stability and error bounds on the approximation for the restricted case

of a rectangular domain with periodic boundary conditions.

Dawson later analyzed a scheme similar to the one presented in [4] applied to the

nonlinear advection-diffusion-reaction equation in [48]. Only the 1D situation was consid-

ered, and adsorption was assumed to be instantaneous (equilibrium). Error estimates for

the semi-discrete formulation were stated and proven, and convergence of the fully discrete

formulation was illustrated with numerical results.

In [49], Dawson and Aizinger continued analyzing the scheme used in [48] with more

relaxed assumptions on the data. The diffusion tensor was no longer assumed to be positive

definite (they considered the possibility of zero diffusion), and more realistic boundary con-

17

Page 29: Modeling, Analysis, and Simulation of Adsorption in ...

ditions than homogeneous Dirichlet were allowed. However, the reaction term was simplified

significantly in that it was merely a linear term given by qC with q a nonpositive constant.

Stability and error bounds were derived, and numerical experiments were presented that

supported the theoretical bounds.

In [21], Brunner, Radu, and Knabner analyzed an upwind-mixed hybrid finite ele-

ment method for the transport problem that differed from the one developed and analyzed

by Dawson and Aizinger in [49]. This upwind-mixed scheme was designed to improve the

stability in the case of strongly advection-dominated problems while maintaining the effi-

ciency of the classical mixed method previously analyzed [7, 19, 86, 87]. It was based on an

Euler-implicit mixed hybrid finite element discretization using RT elements of lowest order.

This method was introduced and studied numerically in the context of reactive transport

simulation in [137]. They considered the linear parabolic advection-diffusion-reaction prob-

lem where the reaction term was of the form RC with R constant and derived optimal order

convergence in time and space for the fully discrete formulation.

Radu, et al. studied the problem of multicomponent reactive transport in porous

media using a mixed hybrid finite element discretization scheme [131]. The advection-

diffusion-reaction problem was coupled with Richards’ equation for flow in possibly un-

saturated mediums and mass balance equations for each species. Although no theoretical

analysis was presented, a detailed description of the discretization and solution process was

given along with numerical results showing good convergence; Newton’s method was used

ultimately to resolve the nonlinearities in the system due to the adsorption function.

Radu and Brunner continued their work with multicomponent reactive transport in

porous media in [22]. They presented a mass conservative modified mixed finite element

scheme using BDM1 elements; the scheme was modified based on the hybrid form for

the mixed finite element scheme. Numerical results were presented that showed second

order convergence for the flux variable which was an improvement over the classical BDM1

elements which are known to be only first order accurate. The new scheme was also shown to

be more robust for high Peclet numbers using classical BDM1 elements. For the strongly

18

Page 30: Modeling, Analysis, and Simulation of Adsorption in ...

advective problem, Radu and Brunner combined the modified BDM1 approach with an

upwinding technique, the same used in [137].

1.1.2 Numerical Methods for Advection-Diffusion Equation

In addition to the numerical schemes analyzed in the literature discussed in section

1.1.1, other numerical schemes used for transport in porous media are considered with little

or no theoretical analysis. However, much numerical experimentation is included confirming

the expected behavior for the schemes. A few papers summarizing numerical techniques for

transport in porous media are available [24, 109, 137]. In [24], Celia focused on a finite dif-

ference approximation to the advection-diffusion-reaction equation and only briefly touched

on finite element methods and a few modifications of finite elements such as upwinding tech-

niques and the method of characteristics. In [137], Radu, et al. summarized the Galerkin

finite element method, finite volumes, and mixed hybrid finite elements as applied to the

contaminant transport problem. Miller et al. [109] provided a summary of many different

models pertaining to water resources systems, including the reactive transport problem,

along with numerical simulation methods pertaining to each model.

As already mentioned, the mixed-derivative form of Richards’ equation is very sim-

ilar in form to the advection-diffusion-reaction equation so solution methods for Richards’

equation can many times be applied to the transport problem. Summary papers concern-

ing numerical methods for Richards’ equation are available as well [26, 88]. In [26], Celia

and Binning presented summaries of numerical methods based on a few different forms

of Richards’ equation including both finite element and finite difference methods. Ju and

Kung [88] restricted their attention to finite element methods for Richards’ equation.

1.1.2.1 Operator Splitting

Operator splitting is very likely the most common numerical solution method for

the advection-diffusion-reaction problem [25, 56, 63, 68, 90, 91, 94, 95, 100, 103, 104, 123,

126, 128, 142, 161, 168, 174]. In the case of advection-dominated flow with no reaction,

19

Page 31: Modeling, Analysis, and Simulation of Adsorption in ...

time-splitting methods can deal with the advection dominance by splitting the advection

and diffusion operators; then the advection step is solved using a higher-order explicit

formulation and small time step while the diffusive step is solved with a lower order implicit

method and large time step [168]. If a reaction term is present, an additional step can be

taken at each time step to solve the reaction term as in [90, 91, 94, 95, 141, 142]. If prior

knowledge is known of the inherent physics in the problem, decoupling techniques can be

developed to accurately approximate the fully-coupled system [56].

One of the benefits of splitting methods is that different spatial discretizations can

be employed for each sub-step, e.g. finite volume method for diffusion term and analytical

solution for advection and reaction as in [90, 141, 142]. Another benefit is that nonlinearities

are much easier to resolve. If nonlinearity in the reaction term or diffusion/dispersion tensor

is allowed, then some linearization scheme must be implemented; however, each nonlinearity

is solved separately in a simpler problem. One of the downfalls of splitting methods is that,

depending on the interaction of the physics in the problem, an exceedingly small step size

may be necessary to ensure stability and accuracy of the method [73, 93, 145, 169]. The

theory of classical operator splitting algorithms has been well known for some time (see

[68, 103, 104, 123, 128, 161, 174]), and new splitting techniques have been proposed and

analyzed more recently [63, 91, 100, 126].

1.1.2.2 Eulerian-Lagrangian Localized Adjoint method (ELLAM)

The Eulerian-Lagrangian Localized Adjoint method (ELLAM) is another solution

method commonly applied to the advection-dominated transport equation. ELLAM, first

developed by Celia et al. and applied to the advection-diffusion equation in [28], is based

on defining test functions that satisfy the homogeneous space-time adjoint equation locally

with elements that have boundaries defined by space-time curves. This results in test

functions that are functions of both space and time. ELLAM was extended first to the

reactive transport case by Celia and Zisman in [30] and then to a set of coupled nonlinear

reaction equations designed to model contaminant transport by Ewing and Celia in [55].

20

Page 32: Modeling, Analysis, and Simulation of Adsorption in ...

More recently, Farthing et al. applied ELLAM to the advective-dispersive transport equation

with nonlinear adsorption [60].

1.1.2.3 Optimal Test Function (OTF) method

In [29] and [96], the authors investigated a new numerical solution procedure for the

simulation of reactive transport in porous media. Celia, Kindred and Herrera developed the

new method, referred to as the optimal test function (OTF) method, in [29]; the technique

used for the OTF method is derived by manipulating the test functions in the weak form

based on the homogeneous adjoint equation. The method then has the property that

it automatically adjusts to accommodating varying degrees of diffusion, advection, and

reaction dominance. In [96], Celia and Kindred developed a new conceptual model for

contaminant transport and tested the OTF method on the new model.

1.1.2.4 Richards’ Mixed-Form Methods

Another numerical solution method developed by Celia and colleagues was described

and implemented in [25, 27]. The method was originally developed for the mixed form

of Richards’ equation and therefore can be easily implemented on the reactive transport

equation because of its similarities to the mixed-form of Richards’ equation. The mixed

form was studied because of known issues with numerically solving the unmixed forms for

Richards’ equation (head-based and saturation-based). The head-based form has large mass

balance errors while the saturation-based form can degenerate in fully saturated media.

Therefore, Celia and colleagues proposed a modified numerical approach which did not

degenerate like solution techniques for the saturation-based form and had better mass-

balancing properties than solution techniques for the head-based form. The solution method

was based on a fully implicit time discretization applied to the mixed-form of Richards’

equation with an appropriate Taylor expansion of the time derivative to obtain a simple

computational algorithm. The method was derived and tested in [27] and then expanded

to two-phase flow in [25].

21

Page 33: Modeling, Analysis, and Simulation of Adsorption in ...

In [61], Richards’ equation was coupled with the non-reactive advection diffusion

equation: Celia and Forkel used the method from [27] to simplify the discretization of

the flow equation and then used a simple Picard iteration to solve the coupled flow and

transport equations. No analysis was presented with this method although the numerical

results seemed to indicate that this method, when used in conjunction with a mass-lumped

time matrix, gave reliable and robust solutions to Richards’ equation. When combined with

transport, the method had difficulty in cases that had little to no pressure-driven flow, that

is, in cases of diffusion dominance.

1.1.2.5 Method of Lines (MOL)

In [159], Kelley et al. solved the pressure-head form of Richards’ equation using

a differential algebraic implementation of the method of lines (MOL). MOL is a solution

technique that first reduces a partial differential equation to a system of ordinary differ-

ential equations (ODEs) by a standard spatial discretization (e.g. finite difference, finite

elements, etc.) and then integrates in time using a known ODE code. Kelley et al. showed

with numerical results that this method applied to Richards’ equation has good mass balance

properties, has the ability to explicitly control temporal truncation error, and is more eco-

nomical than many other solution techniques for Richards’ equation. The one-dimensional

case was considered in [159], but the method was quickly expanded to two dimensions in

[110] and a nonuniform medium in [160]. MOL can also be combined with a mixed hybrid

finite element spatial discretization to ensure local mass conservation as shown in [57].

1.1.2.6 Linearization Techniques

Considering specifically the nonlinearities that can arise in the advection-diffusion-

reaction equation, there are many numerical techniques that can be used to linearize the

equation. There has been much comparison between the normal Picard and Newton algo-

rithms when applied to either Richards’ equation or the advection-diffusion-reaction equa-

tion (see for example [165, 69, 102, 122, 127, 164]). In general, Newton’s method far

22

Page 34: Modeling, Analysis, and Simulation of Adsorption in ...

outperforms a Picard iteration for nonlinear problems in porous media. It is much more

robust so it can handle a wider range of parameters values, and it converges significantly

faster [122, 127]. Newton’s method of course does have the issue of needing a “good enough”

initial guess, but techniques have been developed to find better initial guesses for Newton’s

method [59, 127]. However, a Picard iteration is used much of the time in practice because

it is simpler to code and computationally cheap [69, 122, 127].

Because of this common usage, modifications to the Picard and Newton schemes

have been developed and studied in the hopes of finding a linearization method that is more

robust, yet still simple to implement. Paniconi and Putti gave a brief overview of a few of

these modifications in [122]. Specifically in [79], Huyakorn et al. proposed a relaxation to a

Picard iteration based on an adaptation of an empirical scheme for the solution of variably

saturated flow. They implemented this relaxed Picard method for the coupled Richards’-

Darcy-Transport system. Numerical results showed that the relaxed Picard method was

able to handle a wide range of parameter values converging faster and for a broader range

of values.

In [127], a partial Newton method was proposed for the coupled Richards’-Darcy-

Transport system with dispersion. This partial Newton method applied a Newton lineariza-

tion to the transport equation alone and then solved the system by completing a three-step

sequence similar to the Picard scheme: first the flow equation was solved using the previ-

ously calculated values for the head and concentration, then the velocities were updated,

and last the concentration values were updated. This method was motivated by the fact

that the flow equation is only weakly nonlinear and the importance of coupling and the

degree of the nonlinearity in the transport equation decreases as the dispersion becomes

dominant. The partial Newton method resulted in a system the same size as a Picard

iteration. Numerical experiments were conducted comparing a Picard iteration, a relaxed

Picard iteration, and a partial Newton iteration for a range of parameter values. The results

showed that the partial Newton method was more robust than either the Picard or relaxed

Picard iterations.

23

Page 35: Modeling, Analysis, and Simulation of Adsorption in ...

Another more recent comparison of Newton and Picard methods for solving non-

linear ground water flow problems was given in [105]. An analysis was done to compare

the effects of different nonlinearities on convergence and convergence rates. Results showed

that certain nonlinearities affected convergence more than others. They also showed that

no single strategy was effective for all problems so that a Picard iteration may be better for

some situations while a Newton iteration may be better for others.

In just the last few years, a new modification of a Picard iteration was suggested

by Walker and colleagues in [165, 102, 164]. A Picard iteration was enhanced by adding

an Anderson acceleration as described in [3] and [58]. Although Anderson acceleration has

been widely used with considerable success in electronic structure computation (in which

it is referred to as Anderson mixing), it has not been used much in other applications, nor

has it been studied extensively or numerically analyzed [58, 164]. Fang and Saad began the

analysis in [58] by showing a close and important relationship between Anderson acceleration

and secant updating methods. Walker and Ni continued the analysis in [164] by showing

that an Anderson acceleration applied to linear problems was “essentially equivalent” to the

GMRES method. Practical considerations for implementing Anderson acceleration were

presented along with numerical experiments showing its performance on many different

types of applications.

In [165] and [102], Walker, Woodward and associates studied the performance of a

Picard iteration with an Anderson acceleration in the context of variably saturated flow by

applying it to the mixed form of Richard’s equation. Numerical experiments were conducted

to compare a Newton iteration, a Picard iteration, and the Anderson-accelerated Picard

iteration over a range of parameter values. A line search algorithm was used in conjunction

with the Newton iteration to find good initial guesses for Newton’s method. The results

showed that the accelerated Picard iteration converged for more parameter values than

either the unmodified Picard iteration or the Newton iteration. In the cases when both

Newton and the accelerated Picard converges, Newton’s method was still slightly faster,

but not significantly so. The authors admitted that their numerical solver did not exploit

24

Page 36: Modeling, Analysis, and Simulation of Adsorption in ...

the symmetric nature of the accelerated Picard iteration; they therefore hypothesized that

the accelerated Picard iteration could actually be faster than Newton’s method if an efficient

symmetric solver were to be used.

Other linearizations besides Picard and Newton’s methods have been developed

over the years. In [152], Slodicka considered a nonlinear second-order elliptic boundary

value problem. The solution was found via a linearization technique originally developed

by the author in [150] and [151] which used an assumption of Lipschitz continuity on the

nonlinearity. In [150] and [151], the author proved error estimates for the linearization

scheme assuming Lipschitz continuity on the nonlinear functions. Slodicka extended this

analysis in [152] to the case of an unbounded nonlinearity.

1.2 Structure of the Dissertation

In this work, we lay out the details for the reactive transport model in addition to

describing the details of the SUPG discretization. Also, both linear and nonlinear analysis of

the discretization is presented. The nonlinear analysis considers both the case of adsorption

with an explicit representation and the case of adsorption with an implicit representation.

In the next chapter, we develop the equations used to model the protein adsorption

process along with detailed descriptions of the parameters involved and possible adsorption-

modeling functions. Chapter 3 contains a detailed derivation of the SUPG formulation be-

ginning with preliminaries necessary for the finite element approximations and the deriva-

tion of the Galerkin finite element variational formulation. In section 3.4, we recall common

inequalities that are used in the analysis. In chapter 4, we analyze the steady-state form

of the transport equation. Chapter 5 contains solvability, stability, and error bounds for

the transport problem with constant and linear adsorption. We present the nonlinear anal-

ysis in chapters 6 and 7. Chapter 6 includes solvability, stability, and error analysis valid

under the assumption of an explicit representation for the adsorption. Chapter 7 extends

the analysis to the case of an implicit representation for the adsorption by consider two

25

Page 37: Modeling, Analysis, and Simulation of Adsorption in ...

unknowns, q and C. Chaper 8 contains numerical results supporting the theoretical results

and a comparison of results from a numerical simulation to experimental data. We conclude

in Chapter 9 with a summary of work and proposals for future work.

26

Page 38: Modeling, Analysis, and Simulation of Adsorption in ...

Chapter 2

The Transport Equation

In this section, we describe the modeling equations for reactive mass transport of a

substance in a porous medium.

2.1 The Physics Involved

Consider a single fluid phase present in a porous medium having varying composition

and properties. The fluid phase is comprised of a substance in variable concentration in

a liquid (e.g. protein molecules in a salt-water solution). We will represent the physical

domain of the medium by Ω and consider the evolution of the fluid and medium over the

finite time interval [0, T ] where T represents the final time. The substance in the fluid is

transported in Ω over [0, T ] by three main mechanisms of migration: advection, diffusion,

and kinematic dispersion [50, 124].

2.1.1 Advection

Advection is the phenomenon where dissolved substances are carried along by the

movement of fluid [50, 124]. If one assumes the transport of the substance is governed only

by advection, then the resulting transport equation is easily found by using the principle of

27

Page 39: Modeling, Analysis, and Simulation of Adsorption in ...

mass balance in conjunction with the divergence theorem. This leads to the equation

−∇ · (uC) = ω∂C

∂t(2.1.1)

where u is the Darcy velocity vector and ω represents the total porosity, i.e. the ratio of

void space to total space, of the medium. See section 2.1.5 for more information on Darcy

velocity.

2.1.2 Diffusion

Molecular diffusion is the physical movement linked to molecular agitation. In a

fluid at rest, Brownian motion projects particles in all directions of space. If the chemical

potential of the fluid is uniform in space, then two neighboring points in space send out on

average the same number of particle towards each other. However, if the chemical potential

at one point is higher than at a neighboring point, the point with higher chemical potential

sends out more particles than the one with lower potential. The net result of this molecular

agitation is that particles are transferred from areas of higher chemical potential to areas

of lower potential [50, 124]. For ideal solutions, chemical potential can be replaced by

concentration.

The mass flux of particles in a fluid at rest can be modeled by Fick’s Law, which

says that the mass flux of particles in a fluid at rest is proportional to the concentration

gradient:

φ = −d∇C (2.1.2)

where φ represents the mass flux of the particles in the fluid and d is an isotropic quantity

known as the molecular diffusion coefficient [50, 124, 11]. The molecular diffusion coefficient

28

Page 40: Modeling, Analysis, and Simulation of Adsorption in ...

can be expressed by the relationship

d =RT

N

1

6πµr

where R is the constant of perfect gases, 8.32 SI units (mass length2 time−2 kelvin−1); N

is Avogadro’s number, 6.023× 1023; T is absolute temperature; µ is fluid viscosity; and r is

the mean radius of the diffusing molecular aggregates [50]. Common ions typically have a

diffusion coefficient between 10−5 and 2×10−5 cm2/s at 20C [50]; however larger molecules

such as some protein molecules can have a smaller diffusion coefficient on the order of 10−7

cm2/s [154].

If the transport of substances in a fluid at rest is due solely to Fickian diffusion, then

the principle of mass balance in conjunction with (2.1.2) leads to the following equation for

the movement of the substance:

∇ · (d∇C) =∂C

∂t, (2.1.3)

and for a fluid moving in a porous medium, the effects of both advection and diffusion on

the transported substance given in (2.1.1) and (2.1.2) can easily be combined to result in

the following governing equation [50]:

∇ · (ωd∇C)−∇ · (Cu) = ω∂C

∂t. (2.1.4)

The porosity appears on the left hand side now since diffusion is nonzero only over the

pores, but the Darcy velocity is defined over the entire area.

2.1.3 Dispersion

Kinematic dispersion is a mixing phenomenon linked mainly to the heterogeneity of

the microscopic velocities inside the porous medium [50, 124]. Inside a pore, the velocities

are not uniformly distributed. The laminar flow in the pores causes the substance to be

29

Page 41: Modeling, Analysis, and Simulation of Adsorption in ...

transported faster along the axis of the pores. Because of mixing and molecular diffusion,

this faster movement causes a progressive spreading of the substance compared to the mean

movement of advection. Differences in pore opening sizes and travel distance from one pore

to another can also cause differences in mean velocities of different pores. Additionally,

any type of heterogeneity in the porous medium, such as interlayered deposits, broken

or fractured zones, etc., also introduces a variation in the velocity field which causes the

substance to mix and spread differently in all directions [50, 11].

Consequently, the kinematic dispersion is caused by the existence of a very complex

velocity field which is not fully taken into consideration by use of the Darcy velocity used in

the advective term as Darcy velocity is an averaged quantity. The suggested mathematical

formula for dispersion is very similar to that of Fick’s Law:

dispersive flux φ = −D∇C

where D is a dispersion coefficient represented by a second-order tensor that is assumed

to be symmetric [50, 11]. The dispersion tensor’s main (and potentially only) principal

direction is in the direction of the velocity vector of the flow. If more than one dimension

is considered, then the other principle directions are normally arbitrary and always chosen

to be at right angles to the first one and each other [50]. The value(s) of the dispersion

coefficient are dependent on the flow velocity and therefore may be notated as D(u).

To take into consideration the affect of dispersion on the concentration of the trans-

ported substance, the dispersion flux D∇C is added to the diffusive flux d0∇C on the left

hand side of equation (2.1.4) which results in the transport equation:

∇ · (D∇C)−∇ · (Cu) = ω∂C

∂t(2.1.5)

where now D takes into account both the diffusion and dispersion [50, 11].

To obtain values for the dispersion tensor, it is often helpful to express the dispersion

tensor in terms of its principal directions of anisotropy as this makes it a diagonal tensor;

30

Page 42: Modeling, Analysis, and Simulation of Adsorption in ...

in three dimensions, the dispersion coefficient would look like the following:

D(u) =

∣∣∣∣∣∣∣∣∣∣DL 0 0

0 DT 0

0 0 DT

∣∣∣∣∣∣∣∣∣∣where DL is the longitudinal dispersion coefficient in the direction of the flow, and DT is

the transverse dispersion coefficient in the two directions orthogonal to the velocity. Notice

that D is always anisotropic, even if the medium is isotropic, because the spreading of

concentration in the direction of the flow is always greater than in the transverse directions

[50].

The values of DL and DT vary depending on the dispersion regime. Recall the

dimensionless Peclet number, denoted Pe, defined as

Pe =advective transport rate

diffusive flux rate.

For diffusion of matter, it can be written simply as

Pe =`|u∗|d0

(2.1.6)

where ` is a characteristic length, d0 is the coefficient of molecular diffusion, and |u∗| is the

mean microscopic velocity, related to the Darcy velocity by [50]

|u∗| = |u|ω.

Through experimentation, an empirical relationship connecting the dispersion coef-

ficients with the Peclet number has been established [50]. For larger velocities (Pe > 10),

the effects of dispersion outweigh the effects of diffusivity so the following relations are

31

Page 43: Modeling, Analysis, and Simulation of Adsorption in ...

generally admitted:

DL = αL|u|,

DT = αT |u|,

where αL and αT , which have the dimension of a length, are known as intrinsic dispersion

coefficients or dispersivities [50]. A more general expression for the dispersion coefficients

that takes into account the effect of diffusivity can also be used to extend the validity of

the model to lower Peclet numbers:

DL = ωd+ αL|u|,

DT = ωd+ αT |u|.

where ω and d are the previously mentioned total porosity and molecular diffusion respec-

tively [50, 11]. The value of αL varies greatly depending on the situation: measurements in

a laboratory on a column of sand give values for αL on the order of a few centimeters while

varying heterogeneity gives values for αL ranging from a meter to a hundred meters. The

value of αT is always much smaller than that of αL, between 15 and 1

100 of αL [50]. More

complex models for D can be considered also (see [11] for more information).

Note: in this work, we assume either that all of the fluid in the medium is mobile or

that the concentration in the mobile fraction instantaneously reaches an equilibrium with

the concentration in the immobile fraction. If either one of these assumptions is not true,

then the we must consider the fluid as two types of fluids, mobile and immobile, each with

an associated porosity and concentration. Then the transport equation is changed only

slightly from above:

∇ · (D∇C − CU) = ωc∂C

∂t+ (ω − ωc)

∂C ′

∂t

where ωc is the kinematic porosity (the porosity associated with the mobile fluid), C is

32

Page 44: Modeling, Analysis, and Simulation of Adsorption in ...

now the concentration in the mobile fluid and C ′ corresponds to the concentration in the

immobile fraction [50].

2.1.4 Adsorption

Certain mechanisms in a porous medium can turn the transport of substances into

a reactive process. These mechanisms can include any of the following:

physical mechanism in which transported substance is stopped by physical filtration

through the pores of the medium (this can occur even if the substance is much smaller

than the size of the pores);

chemical mechanism such as combining of ions, oxidation-deduction reactions, adsorption-

desorption, etc. (the adsorption-desorption can have many different causes such as

ion-exchange interaction, hydrophobic interaction, and affinity interaction);

radiological mechanisms such as radioactive decay and the creation of daughter prod-

ucts by this decay, and

biological mechanisms which can decompose or transform elements, e.g. radioactive

decay.

Most often, all of these mechanisms are modeled by one “net source or sink term”,

S, in the transport equation which accounts for the lack of mass balance:

∇ · (D∇C)−∇ · (Cu) = ω∂C

∂t+ S (2.1.7)

[50, 11]. S represents a source or sink depending on the sign of the term: positive for a

sink term and negative for a source term. Occasionally, the lack of mass balance is instead

modeled as a “retardation” term (see [124] for more information).

In this paper, we will concentrate on chemical reactions causing adsorption in the

porous medium. In the case of adsorption, solutes become attached to the solids in the

porous medium so we must consider an amount of concentration bound to the solid. We

33

Page 45: Modeling, Analysis, and Simulation of Adsorption in ...

represent this amount with a dimensionless mass concentration, q, which gives the mass of

substances adsorbed per unit mass of solid.

The mass of solid in a unit volume of porous medium is given by (1 − ω)ρs where

ω is the total porosity and ρs is the density of the solid particles being adsorbed; therefore,

the mass of substances bound to the solid is then (1− ω)ρsq. The source/sink term in the

reactive transport equation is then given by the variation of this quantity per unit time [50]:

S = (1− ω)ρs∂q

∂t. (2.1.8)

The complexity of modeling adsorption then comes with finding an appropriate

relationship between the concentration in the adsorbed phase q and the concentration in

the mobile phase C. The function for q is commonly referred to as the binding capacity,

the adsorption isotherm, or simply the isotherm. Many different functions can be used

for the isotherm depending on the type of adsorption that is occurring and the amount

of concentration in the fluid [11, 50, 124]. Most of the forms of the functions have been

determined experimentally, and all contain constants that must be obtained by data fitting.

2.1.4.1 Linear Isotherm

In cases where the transported substances are in very weak concentration, then the

ratio of adsorbed mass concentration to the volumetric concentration is constant; that is,

q

C= K, K > 0

[50, 18]. The coefficient K is called the distribution coefficient of the transported substance

in relation to the porous medium. This model assumes that adsorption is linear, reversible

and instantaneous [50, 11]. One case that warrants a linear isotherm is the case of organic

compounds in solution, such as trichloroethylene. In this case, the isotherm is generally

34

Page 46: Modeling, Analysis, and Simulation of Adsorption in ...

given by

q = K1 +K2C, K1,K2 > 0

where K1 and K2 are found by fitting to experimental data [124, 156].

2.1.4.2 Nonlinear Isotherms

Many nonlinear adsorption isotherms have been suggested to model the relationship

between q and C. The two most common isotherms used in research have been Langmuir’s

isotherm and Freundlich’s isotherm.

Langmuir’s isotherm [14, 149, 175, 148, 124, 156]:

q =qmaxKeqC

1 +KeqCfor qmax,Keq > 0. (2.1.9)

In this case, the two coefficients have physically meaningful values which is important

when analyzing the adsorptive properties of a porous medium. Keq is called the

Langmuir equilibrium constant, which indicates the concentrations in the liquid and

adsorbed phases at which the rate of adsorption equals the rate of desorption, and

qmax represents the maximum binding capacity of the porous medium.

Freundlich’s isotherm [50, 124, 9, 10, 132]:

q = KC1/n, for K > 0, n ≥ 1. (2.1.10)

Although Freundlich’s isotherm does not have the physical meaning associated with

Langmuir’s isotherm, it has more mathematical significance when it comes to analysis

of discrete methods since it is not Lipschitz continuous when C = 0. This lack of

regularity complicates the analysis and forces some type of regularization step to be

used before any analysis can be completed.

35

Page 47: Modeling, Analysis, and Simulation of Adsorption in ...

2.1.4.3 Mixed Mode Isotherms

When adsorption is caused by multiple methods at the same time, i.e. ion-exchange

and hydrophobic interactions, then a multimodal or mixed mode isotherm must be used in

order to accurately capture the transport of the substance in the fluid [120]. The Langmuir

isotherm provides more useful information than most of the other common isotherms (from

the values of the parameters Keq and qmax); however it does not capture the influence of

the adsorbent, the medium attracting the substance in the fluid, on the adsorption. An

exponentially modified nonlinear Langmuir isotherm was previously proposed by Ghose

et. al. in [67] to model the affect of pH and salt concentration on protein adsorption.

A similar isotherm was used by Gao et al. to study the adsorption of a specific type of

protein, bovine serum albumin (BSA), onto the mixed-mode adsorbents that they were

studying [65]. However, one of the downfalls of this modified Langmuir isotherm is that the

parameters lose their physical meaning.

The deficiencies in earlier models of multimodal isotherms led to the development of

a new mixed mode isotherm model by Mollerup [112, 115, 113, 116, 114]. This multi-modal

isotherm has a unique set of meaningful parameters capable of describing the adsorption

of different proteins over a wide range of operating conditions. A very detailed parameter

fitting process was described by Nfor et. al. in [120]. The multi-modal isotherm which

takes into account both interaction chromatography and ion exchange chromatography is

given by the following implicit relationship:

q

C= KeqΛ

ν+n

(1

zcs

)ν ( 1

cm

)n(1− q

qmax

)ν+n

γp (2.1.11)

where Λ is the ligand density, ν is the known stoichiometric coefficient of the salt counter-

ion, n is the stoichiometric coefficient of the ligand, z is the known charge on the salt

counter-ion, cs is the known salt concentration, cm is the known molar concentration of

solution in the pore volume, qmax is the known maximum binding capacity for the proteins

36

Page 48: Modeling, Analysis, and Simulation of Adsorption in ...

onto the porous medium, γp is the normalized activity coefficient

γpγ∞,wp

with γp the known activity coefficient of the protein, and Keq is given by

Keq = Keqγ∞,wp .

The normalized activity coefficient is expressed by a suitable activity coefficient model as

in [112]:

γp =γpγ∞,wp

= exp(Kscs +KpC)

where Ks and Kp are interaction constants that are specific for the salt and the protein

respectively.

2.1.4.4 Equilibrium vs. Non-Equilibrium Adsorption

Depending on the rate of the reaction with respect to the flow, the chemical reaction

describing the adsorption may be either fast (equilibrium) or slow (non-equilibrium) [11,

156, 9, 10, 48]. In the case of equilibrium adsorption reactions, the adsorbed concentration q

in (2.1.8) is simply replaced by one of the isotherm relationships described above [50, 10, 48].

When non-equilibrium adsorption is considered, the adsorbed concentration q is

related to the equilibrium concentration q∗ and the term ∂q∂t is written as

∂q

∂t= km(q∗ − q) (2.1.12)

where km is the mass transfer coefficient or rate parameter [156, 9, 48]. Now the equilibrium

concentration q∗ is given by one of the isotherm relationships described above. Notice if we

consider the case as km →∞, we end up with the case of equilibrium absorption [48]. For

37

Page 49: Modeling, Analysis, and Simulation of Adsorption in ...

most of the work in this paper, we assume equilibrium adsorption although a few numerical

results using non-equilibrium adsorption are shown in Chapter 8.

2.1.5 Coupling with Fluid Movement

The velocity u in (2.1.7) is calculated using Darcy’s law as

u = −kµ

(∇p+ ρg∇z) (2.1.13)

where k is the intrinsic permeability, µ is the dynamic viscosity of the fluid, p is pressure,

ρ is the fluid density, g is the gravitational constant, ∇z is a vector of coordinates 〈0, 0, 1〉

and the axis z is vertical and oriented upward [50, 11]. Recall Darcy’s law is an empirical

relationship between the averaged Darcy velocity u and the permeability of the porous

medium which is an indication of the ability of a fluid to flow through the medium. Notice

the generalized form of Darcy’s equation is written here since ρ may vary with C.

Because of the variation of ρ, to complete the transport-velocity coupling, we must

add the continuity equation for the fluid and two state equations relating ρ and µ to C and

p:

∇ · (ρu) +∂

∂t(ρω) = 0, (2.1.14)

ρ = ρ(C, p), (2.1.15)

µ = µ(C, p), (2.1.16)

[50].

For the purposes of this paper, we consider a simplified version for the velocity to be

more similar to the experimental setup which we are trying to model. In the experimental

setup, ρ and µ are assumed to be constant, the effect of gravity on the velocity is negligible

and ∇p is held constant by the machinery; therefore, u is a prescribed function that is

independent of time and concentration.

38

Page 50: Modeling, Analysis, and Simulation of Adsorption in ...

2.1.6 Boundary Conditions

Let Ω be described by a bounded polygonal domain in Rd, d = 1, 2 or 3 with Γ = ∂Ω.

We will divide the boundary Γ into two to three regions: the inflow, the outflow, and the

(possibly nonexistent) no flux boundary. More formally, we will denote the inflow boundary

by Γ+, that is

Γ+ = x ∈ Γ : n(x) · u(x) > 0,

where n(x) is the outward unit normal to Γ at x ∈ Γ and we will denote the outflow

boundary by Γ− where

Γ− = x ∈ Γ : n(x) · u(x) < 0.

The boundaries comprising the no-flow hydraulic zone(s) will be denoted by Γn; specifically,

we have Γn = Γ\(Γ+ ∪ Γ−). Note that Γn exists only in the cases when d = 2 or d = 3

and may not exist even in those cases if the boundary is comprised entirely of inflow and

outflow regions.

For the inflow boundary, Γ+, the concentration is fixed:

C = g1 (2.1.17)

[11, 50]. A more complex inflow boundary condition, such as Danckwerts’ boundary con-

ditions, can be applied in certain cases of reactive transport [154, 175, 11, 45]; however, in

this work we restrict out attention to a fixed Dirichlet condition.

For the no-flow hydraulic zone(s) of the boundary, Γn, we assume the velocity u

satisfies

u · n = 0, (2.1.18)

where n is the outward unit normal vector for Γn, so the advective flux will always be zero.

39

Page 51: Modeling, Analysis, and Simulation of Adsorption in ...

We also assume that no solute flow is coming in or out by diffusion, i.e.

(D∇C) · n = 0. (2.1.19)

Again, a more general boundary condition, (D∇C) ·n = g2, can be used which applies even

in the case when there is known diffusion across the boundary [50].

For the outflow boundary, we enforce a computational “do nothing” or natural

boundary condition [98]. Based on the variational formulation we use in this work, the

natural boundary condition on the outflow is equivalent to a homogeneous Neumann con-

dition as given in (2.1.19) [98]. Note that in the case of the time-integrated mixed method

analyzed in Section 6.1, the natural boundary condition results in a no flux condition:

(uC − D∇C) · n = 0. For that case, we take no flux conditions an the entire Neumann

boundary, Γn ∪ Γ−.

2.2 The Reactive Transport Problem

Combining (2.1.7)-(2.1.16) with boundary conditions given by (2.1.17)-(2.1.19) and

the stated assumptions, we obtain the reactive transport problem given as follows: for a

forcing function f ∈ L2(0, T ;L2(Ω)) and initial concentrations C0 ∈ L2(Ω) and q0 ∈ L2(Ω),

we consider the following problem

ω∂C

∂t+ (1− ω)ρs

∂q

∂t+∇ · (uC)−∇ · (D∇C) = f, x ∈ Ω, t > 0, (2.2.1)

q = g(q, C), x ∈ Ω, t > 0 (2.2.2)

C(x, t) = Cin, x ∈ Γ+, t > 0, (2.2.3)

(D∇C) · n(x, t) = 0, x ∈ Γn ∪ Γ−, t > 0, (2.2.4)

u · n(x, t) = 0, x ∈ Γn, t > 0 (2.2.5)

C(x, 0) = C0, x ∈ Ω, (2.2.6)

q(x, 0) = q0, x ∈ Ω (2.2.7)

40

Page 52: Modeling, Analysis, and Simulation of Adsorption in ...

where ω is the total porosity, ρs is the density of solid particles in the fluid, u is the Darcy

velocity of the fluid, D is the diffusion tensor, and C and q are the concentration in the

liquid and adsorbed phases respectively. Note that most often we will have f = 0 and

C0 = 0.

41

Page 53: Modeling, Analysis, and Simulation of Adsorption in ...

Chapter 3

SUPG Formulation

In this section we formulate a streamline upwinding Petrov-Galerkin (SUPG) dis-

cretization for (2.2.1)-(2.2.4). As was stated in the introduction, the standard Galerkin

method produces an oscillating solution for advection-dominated transport problems (that

is, if ‖D‖ << h) when the exact solution is non-smooth [17, 82, 70]. In some cases, the

solution loses stability especially in the presence of nonlinearity [17]. Several strategies, a

few of which were described in section 1.1.1, have been employed to correct this oscillatory

behavior. Each solution method has downfalls; however, the streamline diffusion method

and discontinuous Galerkin method have high order accuracy and good stability characteris-

tics while performing significantly better than the conventional solution methods (standard

Galerkin and classical artificial diffusion) [82]. As was shown in section 1.1.1, there is little

analysis involving nonlinear adsorption with the SUPG method, and the analysis that is

available is restricted to the SUPG method being using with mixed methods. We therefore

will develop and analyze an SUPG method applied to solve the reactive transport equation.

3.1 Variational Formulation

To aid in the derivation of the SUPG formulation, we will derive the standard

Galerkin variational formulation for (2.2.1)-(2.2.7). First we introduce notation that will

42

Page 54: Modeling, Analysis, and Simulation of Adsorption in ...

be used throughout the paper.

Recall the L2 inner product over Ω is defined as

(u, v) =

∫ΩuvdΩ, ∀u, v ∈ L2(Ω)

and the L2 norm over Ω is defined by

‖u‖L2(Ω) = (u, u)1/2 ∀u ∈ L2(Ω).

The L2 inner product over the boundary ∂Ω = Γ will be denoted by

〈u, v〉 =

∫ΓuvdΩ.

Recall also the Hk norm of u over a domain Ω is defined as

‖u‖Hk(Ω) =(‖u‖2L2(Ω) + ‖∇u‖2L2(Ω) + · · ·+ ‖∇ku‖2L2(Ω)

)1/2.

Last recall the L2 time-space norm ‖ · ‖L2(0,T ;L2(Ω)) is defined as

‖u‖L2(0,T ;L2(Ω)) = ‖u‖L2(L2) =

(∫ T

0‖u(·, t)‖2L2(Ω)dt

)1/2

,

and the L∞ time-space norm ‖ · ‖L∞(0,T ;L2(Ω)) is defined as

‖u‖L∞(0,T ;L2(Ω)) = ‖u‖L∞(L2) = sup0≤t≤T

‖u(·, t)‖L2(Ω).

For simplicity, we will denote the Hk norm over Ω by ‖ · ‖k, and consequently ‖ · ‖0

will denote the L2 norm on Ω. If a norm over a space other than Ω is used, it will be

explicitly written; for example, the Hk norm and L2 norm over E ⊂ Ω will be denoted by

‖ · ‖k,E and‖ · ‖0,E respectively.

We also make use of the following discrete norm in the time-dependent error analysis:

43

Page 55: Modeling, Analysis, and Simulation of Adsorption in ...

|||v|||∞,k := max1≤n≤N

‖vn‖k. (3.1.1)

In the derivation of the variational formulations, we will use Green’s Theorem.

Theorem 1 (Green’s Theorem). Given Ω a bounded domain and nE the outward normal

vector to Γ, we have for all v ∈ H2(Ω) and u ∈ H1(Ω),

∫Ωu∆v dΩ =

∫∂E

(∇v · nE)u dΩ−∫E∇v · ∇u dΩ. (3.1.2)

To simplify the notation of the boundary pieces in the analysis, we let ΓD denote the

portion of the boundary having a Dirichlet boundary condition and ΓN denote the portion

with a Neumann boundary condition. According to the boundary conditions in the reactive

transport problem (2.2.1)-(2.2.7), we have

ΓD = Γ+ and ΓN = Γn ∪ Γ−.

Because of the Dirichlet boundary condition on the inflow boundary, we will use

V = H10,D(Ω) = v ∈ H1(Ω) : v = 0 on ΓD

and

W = w ∈ L2(Ω)

for the mobile and adsorbed concentration spaces respectively. The derivation of the vari-

ational formulation of (2.2.1)-(2.2.2) proceeds in the normal manner. Multiplying (2.2.1)-

(2.2.2) by arbitrary functions v ∈ V and w ∈W respectively and integrating over Ω gives

(ω∂C

∂t, v

)+

((1− ω)ρs

∂q

∂t, v

)+ (∇ · (uC), v)− (∇ · (D∇C, v) = (f, v) , (3.1.3)

(q, w) = (g(q, C), w) . (3.1.4)

44

Page 56: Modeling, Analysis, and Simulation of Adsorption in ...

Since u is incompressible, we rewrite the convective term as follows:

∇ · (uC) = (∇ · u)C + u · ∇C = u · ∇C. (3.1.5)

Also we apply Green’s Theorem to the diffusive term to obtain

−(∇ · (D∇C), v) = −〈(D∇C) · n, v〉+ (D∇C,∇v) = (D∇C,∇v) (3.1.6)

where the boundary term is zero based on boundary conditions and the definition of V .

Combining (3.1.3)-(3.1.6), we obtain the following variational formulation: Find (C,w) ∈

V ×W such that

(ω∂C

∂t, v

)+

((1− ω)ρs

∂q

∂t, v

)+ (u · ∇C, v) + (D∇C,∇v) = (f, v) (3.1.7)

(q, w) = (g(q, C), w) (3.1.8)

for all v ∈ V and w ∈W .

3.2 SUPG Derivation

For the SUPG derivation, we assume that D is essentially bounded, that is ‖D‖∞ ≤

∞. The beginning of the SUPG derivation parallels the derivation for the variational for-

mulation. We begin by rewriting (3.1.3) using the incompressibility of u to obtain

(ω∂C

∂t, v

)+

((1− ω)ρs

∂q

∂t, v

)+ (u · ∇C, v)− (∇ · (D∇C, v) = (f, v) . (3.2.1)

To stabilize the solution without affecting the mass matrix, we will upwind the test

functions for all but the temporal terms as in [54, 34, 82] with the test function v+ δu ·∇v,

45

Page 57: Modeling, Analysis, and Simulation of Adsorption in ...

where v ∈ V and δ is defined as

δ =

Kh, ‖D‖∞ < h

0, ‖D‖∞ ≥ h. (3.2.2)

The value for K is chosen to be sufficiently small to stabilize the diffusive term (see future

analysis for specific bounds on K).

Then the upwinded form of (3.2.1) is given by

(ω∂C

∂t, v

)+

((1− ω)ρs

∂q

∂t, v

)+ (u · ∇C, v + δu · ∇v)

− (∇ · (D∇C), v + δ(u · ∇v)) = (f, v + δ(u · ∇v)) .

(3.2.3)

Just as with the variational formulation, we apply Green’s Theorem to the diffusive term.

However, in this case we first split off the upwinded term as follows:

− (∇ · (D∇C), v + δ(u · ∇v)) = −(∇ · (D∇C), v)− (∇ · (D∇C), δ(u · ∇v))

= −〈(D∇C) · n, v〉+ (D∇C,∇v)− (∇ · (D∇C), δ(u · ∇v))

= (D∇C,∇v)− (∇ · (D∇C), δ(u · ∇v))

where again the boundary integral goes to zero based on boundary conditions and the

definition of V . Then (3.2.3) becomes

(ω∂C

∂t, v

)+

((1− ω)ρs

∂q

∂t, v

)+ (u · ∇C, v) + (u · ∇C, δ(u · ∇v))

+ (D∇C,∇v)− (∇ · (D∇C), δ(u · ∇v)) = (f, v) + (f, δ(u · ∇v)) .

(3.2.4)

Therefore, the SUPG formulation of the convection-diffusion equation is as follows: Find

46

Page 58: Modeling, Analysis, and Simulation of Adsorption in ...

(C, q) ∈ V ×W such that

(ω∂C

∂t, v

)+

((1− ω)ρs

∂q

∂t, v

)+ (u · ∇C, v) + δ (u · ∇C,u · ∇v)

+ (D∇C,∇v)− δ(∇ · (D∇C),u · ∇v) = (f, v) + δ (f,u · ∇v)

(3.2.5)

(q, w) = (g(q, C), w) (3.2.6)

for all v ∈ V and w ∈W .

To formulate a discrete analogue to (3.2.3), we must appropriately define the term

(∇ · (D∇C),u · ∇v) as C ∈ H1(Ω) does not guarantee that ∇2C is well-defined. However

we must talk about some finite element preliminaries before continuing. Let Th = E be

a triangulation of Ω so that

Ω =⋃E∈Th

E.

Denote by Pk(E) the space of polynomials of degree less than or equal to k:

Pk(E) = spanxi11 xi22 · · ·x

idd : i1 + i2 + · · ·+ id ≤ k,x ∈ E.

We define the finite element space for the mobile and adsorbed concentrations as

Vh := v ∈ V : v|E ∈ Pk(E), ∀E ∈ Th

and

Wh := w ∈W : v|E ∈ Pn(E), ∀E ∈ TH.

Although we allow for higher order polynomials in Vh in the analysis, we normally use only

linear functions in computation, that is P1(E), to simplify the upwinding terms in the case

of constant dispersion. Also, the regularity of the spaces V = H1(Ω) and W = L2(Ω) allows

for a lower degree polynomials to be used for the adsorbed concentration space, Wh; that

is, we can allow n = k− 1. However, when linear polynomials are used for Vh, as is the case

normally, we use the same polynomials for Wh because constant adsorbed concentration is

47

Page 59: Modeling, Analysis, and Simulation of Adsorption in ...

unrealistic.

We then define (∇ · (D∇Ch),u · ∇v) as in [34, 82, 143]

(∇ · (D∇Ch),u · ∇v

)≡∑E∈Th

∫E∇ · (D∇Ch) (u · ∇v)dE, (3.2.7)

where Th = E is the finite element triangulation of Ω. In other words, we simply sum

over the interior of each triangle E, where ∇2Ch and u · ∇v are well-defined.

It is assumed that Vh satisfies the following local approximation property [80]: For

each v ∈ V ∩Hk(Ω) there exists vh ∈ Vh such that

‖v − vh‖0,E + hE‖v − vh‖∇(v − vh)‖0,E + h2E‖v − vh‖0,E ≤ KhkE‖v‖k,E .

3.3 Finite Element Approximations

In this section, we formulate the finite element approximations for solving the SUPG

formulation of the transport equation. We state the approximations for both the semi-

discrete (continuous-in-time) and fully-discrete cases

3.3.1 Semi-Discrete (Continuous-in-Time) Approximation

Let the spatial discretization of the isotherm be denoted by

qh = q(Ch).

Then the semi-discrete streamline diffusion method is formulated as follows: Find (Ch, qh) ∈

Vh ×Wh such that

(ω∂Ch∂t

, vh

)+

((1− ω)ρs

∂qh∂t

, vh

)+ (u · ∇Ch, vh) + δ (u · ∇Ch,u · ∇vh)

+ (D∇Ch,∇vh)− δ(∇ · (D∇Ch),u · ∇vh) = (f, vh) + δ (f,u · ∇vh)

(3.3.1)

(qh, wh) = (g(qh, Ch), wh) (3.3.2)

48

Page 60: Modeling, Analysis, and Simulation of Adsorption in ...

for all vh ∈ Vh and wh ∈Wh.

3.3.2 Fully-Discrete Approximation

For the fully-discrete approximation, we partition the time interval [0, T ] as

t0 = 0 < t1 < · · · < tN = T

and let ∆t = tn − tn−1 denote the step size for t where tn = n∆t. Also, we let fn = f(tn),

Cn = C(tn), and qn = q(Cn), and we denote the approximation to the time derivative at

time tn for any function h by

∂hn

∂t≈ dt(hn) =

hn − hn−1

∆t.

We will write the finite element approximation for C at time tn as Cnh , and we will assume

C(tn) ≈ Cnh . Then for any θ ∈ [0, 1], the fully-discrete streamline diffusion method is

formulated as follows: For n = 1, ..., N , find (Cnh , qnh) ∈ Vh ×Wh such that

(ωdt(Cnh ), v) + ((1− ω)ρsdt(q

nh), vh) + θ (u · ∇Cnh , vh) + θδ (u · ∇Cnh ,u · ∇vh)

+ θ(D∇Cnh ,∇vh)− θδ(∇ · (D∇Cnh ),u · ∇vh)− θ(fn, vh)− θδ(fn,u · ∇vh)

= (1− θ)(fn−1, vh

)+ (1− θ)δ(fn−1,u · ∇vh)− (1− θ)

(u · ∇Cn−1

h , vh)

− (1− θ)δ(u · ∇Cn−1

h ,u · ∇vh)− (1− θ)(D∇Cn−1

h ,∇vh)

+ (1− θ)δ(∇ · (D∇Cn−1h ),u · ∇vh)

(3.3.3)

θ (qnh , wh)− θ (g(qnh , Cnh ), wh) = −(1− θ)

(qn−1h , qh

)+ (1− θ)

(g(qn−1

h , Cn−1h ), wh

)(3.3.4)

for all vh ∈ Vh and wh ∈Wh.

We will focus on two fully-discrete approximations in the analysis: fully-explicit and

fully-implicit.

49

Page 61: Modeling, Analysis, and Simulation of Adsorption in ...

3.3.2.1 Fully-Explicit Approximation

For the fully explicit case, we take θ = 0 in (3.3.3) to obtain the following variational

formulation For n = 1, ..., N , find Cnh ∈ Vh such that

(ω dt(Cnh ), vh) + ((1− ω)ρsdt(q

nh)) +

(u · ∇Cn−1

h , vh))

+ δ(u · ∇Cn−1

h ,u · ∇vh)

+(D∇Cn−1

h ,∇vh)− δ

(∇ · (D∇Cn−1

h ),u · ∇vh)

=(fn−1, vh

)+ δ

(fn−1,u · ∇vh

) (3.3.5)

(qn−1h , wh

)=(g(qn−1

h , Cn−1h ), wh

)(3.3.6)

for all vh ∈ Vh and wh ∈Wh.

3.3.2.2 Fully-Implicit Approximation

For the fully implicit case, we take θ = 1 in (3.3.3) to obtain the following variational

formulation For n = 1, ..., N , find Cnh ∈ Vh such that

(ω dt(Cnh ), vh) + ((1− ω)ρsdt(q

nh)) + (u · ∇Cnh , vh) + δ (u · ∇Cnh ,u · ∇vh)

+ (D∇Cnh ,∇vh)− δ (∇ · (D∇Cnh ),u · ∇vh) = (fn, vh) + δ (fn,u · ∇vh)(3.3.7)

(qnh , wh) = (g(qnh , Cnh ), wh) (3.3.8)

for all vh ∈ Vh and wh ∈Wh.

3.4 Useful Inequalities

We conclude this chapter by recalling some well-known inequalities that will be used

often in the analysis in Chapters 4 - 7.

3.4.1 Young’s, Holder’s, and Cauchy-Schwarz Inequalities

Young’s, Holder’s, and Cauchy-Schwarz inequalities [97] will be used often in the

analysis to obtain upper and lower bounds on the terms in the variational formulations.

50

Page 62: Modeling, Analysis, and Simulation of Adsorption in ...

Lemma 3.4.1. (Young’s Inequalities) If a and b are nonnegative real numbers and p

and q are positive real numbers such that 1p + 1

q = 1, then

ab ≤ ap

p+bq

q. (3.4.1)

Specifically, for p = q = 2, we have

ab ≤ 1

2a2 +

1

2b2, (3.4.2)

or more generally, for any ε > 0,

ab ≤ 1

2εa2 +

1

2εb2. (3.4.3)

Lemma 3.4.2. (Holder’s Inequality) Let 1 ≤ p, q ≤ ∞ such that 1p + 1

q = 1. Then if

u ∈ Lp(Ω) and v ∈ Lq(Ω), the product uv belongs to L1(Ω), and

∫Ω|uv| ≤ ‖u‖Lp(Ω)‖v‖Lq(Ω). (3.4.4)

Lemma 3.4.3. (Cauchy-Schwarz inequality) If we choose p = q = 2 in Lemma 3.4.2,

then we obtain the Cauchy-Schwarz inequality

∫Ω|uv| ≤ ‖u‖L2(Ω)‖v‖L2(Ω). (3.4.5)

3.4.2 Gronwall’s Inequalities

Both the continuous discrete forms of Gronwall’s Inequality are important when

analyzing the time-dependent formulations [143].

Lemma 3.4.4. (Continuous Gronwall’s Inequality) Let f , g, h be piecewise continuous

non-negative functions defined on (a, b). Assume that g is nondecreasing. Assume that there

51

Page 63: Modeling, Analysis, and Simulation of Adsorption in ...

is a positive constant K independent of t such that

∀t ∈ (a, b), f(t) + h(t) ≤ g(t) +K

∫ t

af(s)ds.

Then

∀t ∈ (a, b), f(t) + h(t) ≤ eK(t−a)g(t). (3.4.6)

Lemma 3.4.5. (Discrete Gronwall’s Inequality) Let ∆t, B, K > 0 and let an, bn,

cn, and dn be sequences of nonnegative numbers satisfying

∀n ≥ 0, an + ∆tn∑i=0

bi ≤ B +K∆tn∑i=0

ai + ∆tn∑i=0

ci.

Then, if K∆t < 1,

∀n ≥ 0, an + ∆tn∑i=0

bi ≤ eK(n+1)∆t

(B + ∆t

n∑i=0

ci

). (3.4.7)

3.4.3 Poincare Inequalities

The classical Poincare-Friedrichs Inequality will be used in the steady-state analysis

[143].

Lemma 3.4.6. (Poincare Inequality) For all v ∈ H1(Ω), there exists a constant KPF

such that

‖v‖0 ≤ KPF

(‖∇v‖0 +

∣∣∣∣∫Γv

∣∣∣∣) .Consequently, for all v ∈ H1

0 (Ω)

‖v‖0 ≤ KPF ‖∇v‖0. (3.4.8)

We note that (3.4.8) is still applicable in the case where v = 0 on a subset of the

boundary provided the subset has nonzero measure. The requirement for v to be zero on

a nonzero-measure set is to account for the possibility of a constant function. In general,

52

Page 64: Modeling, Analysis, and Simulation of Adsorption in ...

the Poincare Inequality does not hold for constant functions as they have zero derivatives.

However, assuming v = 0 on a subset of the boundary ensures that the only way for v to

be constant is if v = 0 in all of Ω for which the Poincare Inequality holds. Hence in this

work, we apply the Poincare Inequality simply by assuming C = 0 on ΓD.

We also use the following relationship which is a direct consequence of Lemma

(3.4.6).

Lemma 3.4.7. For all v ∈ H10 (Ω), there exists a constant KPF such that

‖∇v‖0 ≥ KPF ‖v‖1. (3.4.9)

3.4.4 Interpolation and Inverse Inequalities

The interpolation and inverse inequalities will be used often throughout the anal-

ysis. The following lemma contains an interpolation error involving the weighted elliptic

projection which is a direct consequence of the well-known interpolation error involving the

elliptic projection as stated in [80].

Lemma 3.4.8. (Interpolation Inequality) The weighted elliptic projection z of the exact

solution z is defined as

∀ t ≥ 0, ∀ v ∈ Pk(Ω) a(z(t)− z(t), v) = 0,

for

a(z, v) = (D∇z,∇v).

Then if z belongs to L2(0, T ;Hk(Ω)), the following interpolation error estimate holds:

∀t ≥ 0, ‖z(t)− z(t)‖0 ≤ Chk‖z(t)‖k.

The inverse inequality as stated in [143] is recalled below for convenience.

53

Page 65: Modeling, Analysis, and Simulation of Adsorption in ...

Lemma 3.4.9. Let E be a bounded domain in Rd with diameter hE:

hE = supx,y∈E

‖x− y‖.

Then, there is a constant Ki independent of hE such that for any polynomial function v of

degree k defined on E, we have

∀ 0 ≤ j ≤ k, ‖∇jv‖0,E ≤ Kih−jE ‖v‖0,E .

54

Page 66: Modeling, Analysis, and Simulation of Adsorption in ...

Chapter 4

Steady State Analysis

In this section, we analyze the finite element SUPG approximation for the steady

state transport equation.

4.1 Finite Element Approximation & Assumptions

The steady state form of the reactive transport problem is given by

∇ · (uC)−∇ · (D∇C) = f in Ω, (4.1.1)

∇ · u = 0 in Ω, (4.1.2)

C = Cin on Γ+, (4.1.3)

∇C · n = 0 on Γn ∪ Γ−, (4.1.4)

u · n = 0 on Γn. (4.1.5)

Therefore, the corresponding SUPG variational formulation is as follows: Find C ∈ H20,D(Ω)

such that

(u · ∇C, v) + δ (u · ∇C,u · ∇v) + (D∇C,∇v)− δ(∇ · (D∇C),u · ∇v

)= (f, v) + δ

(f,u · ∇v

) (4.1.6)

55

Page 67: Modeling, Analysis, and Simulation of Adsorption in ...

for all v ∈ V . Then the finite element streamline diffusion approximation for the steady-

state equation is formulated as the following: Find Ch ∈ Vh such that

(u · ∇Ch, vh) + δ (u · ∇Ch,u · ∇vh) + (D∇Ch,∇vh)− δ(∇ · (D∇Ch),u · ∇vh)

= (f, vh) + δ (f,u · ∇vh)(4.1.7)

for all vh ∈ Vh.

We make the following assumptions for the steady-state analysis.

(A1) u is nonzero, independent of time [34] and essentially bounded in space [147, 34] with

∇ · u = 0 [34].

(A2) D = [dij ]i,j=1...n is symmetric positive definite [4, 50], independent of time (since u

is assumed independent of time), and it and its derivatives are essentially bounded

in space [4, 147, 34]; specifically, we notate the boundedness of D using ‖D‖∞ ≤ β1,∣∣∣ ∂∂xidij∣∣∣ ≤ β2, for all i, j.

(A3) C is nonnegative [50, 9, 48], and C = 0 on ΓD.

The assumption that C = 0 on ΓD is not always true in practice. However, the assumption

is still reasonable as a transformation can be applied to C in any case when a nonzero

Dirichlet boundary condition arises to give a solution that is zero on ΓD.

4.2 Boundedness and Coercivity of Bilinear Form

In this section, we consider the bilinear form defined by

aδ(C, v) = (u · ∇C, v) + δ(u · ∇C,u · ∇v) + (D∇C,∇v)− δ(∇ · (D∇C),u · ∇v). (4.2.1)

We begin by stating and proving the boundedness of the bilinear form. The bound

below is stated over the discrete space Vh in terms of the H1 norm on C, but it can also be

shown to be true in a continuous sense in terms of the H2 norm on C.

56

Page 68: Modeling, Analysis, and Simulation of Adsorption in ...

Theorem 4.2.1. (Boundedness of aδ) Assume that (A1)-(A3) are satisfied. Then for

Ch, vh ∈ Vh, there exists a positive constant κB independent of h such that

aδ(C, v) ≤ (1 + δ)κB‖C‖1‖v‖1,

i.e. aδ is bounded.

Proof. For notational simplicity we drop the subscript h from the variables. We use the

assumptions along with Cauchy-Schwarz Inequality to obtain upper bounds on the terms

in aδ as follows:

(u · ∇C, v) ≤ ‖u‖∞‖∇C‖0‖v‖0 ≤ ‖u‖∞‖C‖1‖v‖1

δ(u · ∇C,u · ∇v) ≤ δ‖u‖2∞‖∇C‖0‖∇v‖0 ≤ δ‖u‖2∞‖C‖1‖v‖1

(D∇C,∇v) ≤ β1‖∇C‖0‖∇v‖0 ≤ β1‖C‖1‖v‖1

−δ (∇ · (D∇C),u · ∇v) ≤δ |(∇ ·D : ∇C,u · ∇v)|+∣∣(D : ∇2C,u · ∇v

)∣∣≤δβ2 ‖∇C‖0 ‖u · ∇v‖0 + δβ1

∥∥∇2C∥∥

0‖u · ∇v‖0

≤δβ2‖u‖∞ ‖∇C‖0 ‖∇v‖0 + δβ1Kih−1‖u‖∞ ‖∇C‖0 ‖∇v‖0

≤‖u‖∞(δβ2 + Kβ1Ki

)‖C‖1 ‖v‖1

≤(1 + δ)‖u‖∞(β2 + Kβ1Ki

)‖C‖1 ‖v‖1 .

Note that we use the definition of δ = Kh when bounding the upwinded diffusion term to

57

Page 69: Modeling, Analysis, and Simulation of Adsorption in ...

eliminate the h−1 arising from the application of the inverse inequality. Therefore,

aδ(C, v) ≤(‖u‖∞ + δ‖u‖2∞ + β1 + (1 + δ)‖u‖∞

(β2 + Kβ1Ki

))‖C‖1 ‖v‖1

≤(1 + δ)(‖u‖∞ + ‖u‖2∞ + β1 + ‖u‖∞

(β2 + Kβ1Ki

))‖C‖1 ‖v‖1

=(1 + δ)κB‖C‖1‖v‖1

where κB = ‖u‖∞ + ‖u‖2∞ + β1 + ‖u‖∞(β2 + Kβ1Ki

).

Next we state and prove the coercivity of aδ. Note that the coercivity is stated over

the discrete space Vh as in [80, 153].

Theorem 4.2.2. (Coercivity of aδ) Assume that (A1)-(A3) are satisfied. Then for Ch ∈

Vh, there exists a positive constant κC independent of h such that

aδ(Ch, Ch) ≥ κC ||Ch||21,

i.e. the bilinear form aδ is coercive.

Proof. Again for notational simplicity we drop the subscript h from the variables. Lower

bounds on the terms in aδ are found using the assumptions along with the Cauchy-Schwarz,

Young’s, and Poincare’s Inequalities and are given below.

(u · ∇C,C) =

∫Ω

u · (∇C)CdΩ

=1

2

∫Ω

u · ∇(C2)dΩ

=1

2

∫ΓC2 (u · n) ds−

∫Ω

(∇ · u)︸ ︷︷ ︸=0

C2dΩ

=

1

2

∫ΓD

(C)2︸︷︷︸=0

(u · n)ds+1

2

∫ΓN

(C)2 (u · n)︸ ︷︷ ︸≥0

ds

≥ 0

58

Page 70: Modeling, Analysis, and Simulation of Adsorption in ...

(D∇C,∇C) = (D1/2∇C,D1/2∇C)

= ‖D1/2∇C‖20

≥ λ‖∇C‖20

δ (∇ · (D∇C),u · ∇C) ≤((∇ ·D) · ∇C + D : ∇2C, δu · ∇C

)≤β2δ ‖∇C‖0 ‖u · ∇C‖0 + β1δ

∥∥∇2C∥∥

0‖u · ∇C‖0

≤β2δ‖u‖∞ ‖∇C‖20 + β1δKih−1 ‖∇C‖0 ‖u · ∇C‖0

≤β2δ‖u‖∞ ‖∇C‖20 +1

2‖∇C‖20 +

δ2β21K

2i h−2

2‖u · ∇C‖20

≤β2δ‖u‖∞ ‖∇C‖20 +λ

2‖∇C‖20 +

δ2β21K

2i h−2

2λ‖u · ∇C‖20

⇒ −δ (∇ · (D∇C),u · ∇C) ≥− β2δ‖u‖∞ ‖∇C‖20 −λ

2‖∇C‖20 −

δ2β21K

2i h−2

2λ‖u · ∇C‖20

Therefore,

aδ(C,C) =(u · ∇C,C) + δ(u · ∇C,u · ∇C) + (D∇C,∇C)− δ(∇ · (D∇C,u · ∇v)

≥δ‖u · ∇C‖20 + λ‖∇C‖20 − β2δ‖u‖∞ ‖∇C‖20 −λ

2‖∇C‖20 −

δ2β21K

2i h−2

2λ‖u · ∇C‖20

=

2− β2δ‖u‖∞

)‖∇C‖20 + δ

(1− δβ2

1K2i h−2

)‖u · ∇C‖20 (4.2.2)

To ensure positivity of the components in (4.2.2), we choose δ appropriately. For

the first term on the left, we assume h ≤ 1 and choose δ = Kh such that

K <λ

2β2‖u‖∞.

Consequently

β2δ‖u‖∞ = β2Kh‖u‖∞ <λ

2h ≤ λ

2.

For the second component, it seems that we must choose δ ∼ O(h2). However,

recall that δ is defined to be 0 when β1 > h so that the terms contributing to the second

59

Page 71: Modeling, Analysis, and Simulation of Adsorption in ...

component do not arise [50]. Consequently, we assume β1 ≤ h and use this assumption to

help control the h−2 in the second component. Then we choose δ = Kh such that

K <2λ

β1K2i

so that

δβ21K

2i h−2

2λ≤ δβ1K

2i h−1

2λ= K

β1K2i

2λ< 1.

We note the process used to obtain the above bounds closely follows the work by

Johnson on page 186 of [82]. We use similar reasoning consistently throughout this disser-

tation.

Therefore, we choose δ = Kh such that

K ≤ min

λ

2β2‖u‖∞,

β1K2i

to ensure positivity of both terms.

Last, we use the homogeneous Dirichlet boundary condition to apply Poincare’s

Inequality to (4.2.2) as follows:

aδ(C,C) ≥ K(‖∇C‖20 + ‖u · ∇C‖20

)≥ K

(K−2PF ‖C‖

21 + ‖u · ∇C‖20

)≥ κC‖C‖21

where κC = KK−2PF .

4.3 Error Estimates

In this section we develop a priori error estimates for the steady state transport

equation.

Theorem 4.3.1. Suppose that the assumptions of Theorems 4.2.1 and 4.2.2 are satisfied so

60

Page 72: Modeling, Analysis, and Simulation of Adsorption in ...

that (4.1.6) has an exact solution C ∈ Hk+1(Ω), and assume Ch solves the discrete steady

state SUPG formulation (4.1.7). Then there exists a positive constant K independent of h

such that

‖C − Ch‖1 ≤ Khk(1 + δ)‖C‖k+1.

Proof. Subtracting (4.1.7) from (4.1.6) gives

(u · ∇e, vh) + δ(u · ∇e,u · ∇vh) + (D∇e,∇vh)− δ(∇ · (D∇e),u · ∇vh) = 0 (4.3.1)

where e = C − Ch. Now let C be the elliptic projection of the exact solution C as defined

in Lemma 3.4.8 and define

η := C − C, χ := Ch − C.

Notice then that e = η − χ, and (4.3.1) implies

(u · ∇χ, vh) + δ(u · ∇χ,u · ∇vh) + (D∇χ,∇vh)− δ(∇ · (D∇χ),u · ∇vh)

= (u · ∇η, vh) + δ(u · ∇η,u · ∇vh) + (D∇η,∇vh)− δ(∇ · (D∇η),u · ∇vh).(4.3.2)

Choosing vh = χ and using the definition of the elliptic projection, we rewrite (4.3.2) as

(u · ∇χ, χ) + δ‖u · ∇χ‖20 + (D∇χ,∇χ)− δ(∇ · (D∇χ),u · ∇χ)

= (u · ∇η, χ) + δ(u · ∇η,u · ∇χ)− δ(∇ · (D∇η),u · ∇χ).(4.3.3)

We obtain lower bounds for the terms on the left of 4.3.3 by again using the as-

61

Page 73: Modeling, Analysis, and Simulation of Adsorption in ...

sumptions and Young’s and Cauchy-Schwarz Inequalities.

(u · ∇χ, χ) =

∫Ω

u · (∇χ)χdΩ

=1

2

∫Ω

u · ∇(χ2)dΩ

=1

2

∫Γχ2 (u · n) ds−

∫Ω

(∇ · u)︸ ︷︷ ︸=0

χ2dΩ

=

1

2

∫ΓD

(χ)2︸︷︷︸=0

(u · n)ds+1

2

∫ΓN

(χ)2 (u · n)︸ ︷︷ ︸≥0

ds

≥ 0

(D∇χ,∇χ) = (D1/2∇χ,D1/2∇χ)

= ‖D1/2∇χ‖20

≥ λ‖∇χ‖20

δ (∇ · (D∇χ),u · ∇χ) ≤(∇ ·D : ∇χ+ D : ∇2χ, δu · ∇χ

)≤β2δ ‖∇χ‖0 ‖u · ∇χ‖0 + β1δ

∥∥∇2χ∥∥

0‖u · ∇χ‖0

≤β2δ‖u‖∞ ‖∇χ‖20 + β1δKih−1 ‖∇χ‖0 ‖u · ∇χ‖0

≤β2δ‖u‖∞ ‖∇χ‖20 +1

2ε‖∇χ‖20 +

δ2β21K

2i h−2

2ε‖u · ∇χ‖20

≤β2δ‖u‖∞ ‖∇χ‖20 +λ

10‖∇χ‖20 +

5δ2β21K

2i h−2

2λ‖u · ∇χ‖20

⇒ −δ (∇ · (D∇χ),u · ∇χ) ≥− β2δ‖u‖∞ ‖∇χ‖20 −λ

10‖∇χ‖20 −

5δ2β21K

2i h−2

2λ‖u · ∇χ‖20

62

Page 74: Modeling, Analysis, and Simulation of Adsorption in ...

Upper bounds on the terms on the right are found similarly and are shown below.

(u · ∇η, χ) ≤ ‖u · ∇η‖0‖χ‖0

≤ 1

2ε‖u · ∇η‖20 +

1

2ε‖χ‖20

≤ 1

2ε‖u · ∇η‖20 +

1

2εKPF ‖∇χ‖20

=5KPF

2λ‖u · ∇η‖20 +

λ

10‖∇χ‖20

δ(u · ∇η,u · ∇χ) ≤ δ‖u · ∇η‖0‖u · ∇χ‖0

≤ δ2

4ε‖u · ∇η‖20 + ε‖u‖2∞‖∇χ‖20

=5‖u‖2∞δ2

2λ‖u · ∇η‖20 +

λ

10‖∇χ‖20

δ (∇ · (D∇η),u · ∇χ) ≤(∇ ·D : ∇η + D : ∇2η, δu · ∇χ

)≤β2δ‖u‖∞ ‖∇η‖0 ‖∇χ‖0 + β1δ‖u‖∞

∥∥∇2η∥∥

0‖∇χ‖0

≤β22δ

2‖u‖2∞2ε1

‖∇η‖20 +1

2ε1‖∇χ||20 +

β21δ

2‖u‖2∞2ε2

‖∇2η‖20 +1

2ε2‖∇χ‖20

≤5β22δ

2‖u‖2∞2λ

‖∇η‖20 +λ

10‖∇χ||20 +

5β21δ

2‖u‖2∞2λ

‖∇2η‖20 +λ

10|∇χ‖20

Combining the above bounds with (4.3.3) and adding like terms gives

2− β2δ‖u‖∞

)‖∇χ‖20 + δ

(1− 5δβ2

1K2i h−2

)‖u · ∇χ‖20

≤(

5KPF

λ+

5‖u‖2∞δ2

)‖u · ∇η‖20 +

5β22δ

2‖u‖2∞2λ

‖∇η‖20 +5β2

1δ2‖u‖2∞2λ

‖∇2η‖20.(4.3.4)

To ensure positivity of the coefficients on the left, we choose δ appropriately. For

the first term on the left, we assume h ≤ 1 and choose δ = Kh such that

K <λ

2β2‖u‖∞.

63

Page 75: Modeling, Analysis, and Simulation of Adsorption in ...

Consequently

β2δ‖u‖∞ = β2Kh‖u‖∞ <λ

2h ≤ λ

2.

For the second term on the left, it seems that we must choose δ ∼ O(h2). However

as in the proof of coercivity, we again assume β1 ≤ h to help control the h2 since δ = 0

when β1 > h so this term does not arise in that case [50]. Then we choose δ = Kh such

that

K <2λ

5β1K2i

so that

5δβ21K

2i h−2

2λ≤ δ5β1K

2i h−1

2λ= K

5β1K2i

2λ< 1.

Therefore, we choose δ = Kh such that

K ≤ min

λ

2β2‖u‖∞,

5β1K2i

to ensure positivity of both terms.

We note that the second derivative term on the right of (4.3.4) can be bounded by its

gradient. This terms is associated with the upwinded diffusion term which is defined locally

over the interior of the mesh triangulation [34, 82]. Thus a local approximation property

can be applied as in [80], and the associated δ2 can serve to bound the h−2 appearing from

the application of the approximation property.

Consequently, (4.3.4) implies

Km

(‖∇χ||20 + ‖u · ∇χ‖20

)≤ (1 + δ2)KM

(‖u · ∇η‖20 + ‖∇η‖20

), (4.3.5)

where

Km = min

λ

2− β2δ‖u‖∞, δ

(1− 5δβ2

1K2i h−2

)and

KM = max

5

λ

(KPF +

‖u‖2∞2

),5‖u‖2∞

(β2

2 + KK2i

).

64

Page 76: Modeling, Analysis, and Simulation of Adsorption in ...

Then by the boundedness of u,

‖u · ∇η‖20 + ‖∇η‖20 ≤ (‖u‖2∞ + 1)‖η‖21, (4.3.6)

and by Poincare’s Inequality, we have

‖∇χ‖20 + ‖u · ∇χ‖20 ≥ ‖∇χ‖20 ≥ K−2PF ‖χ‖

21. (4.3.7)

Combining (4.3.5) with (4.3.6) and (4.3.7), we obtain

Km‖χ‖21 ≤ δ2KM‖η‖21

⇒ ‖χ‖21 ≤ δ2 KM

Km

‖η‖21

⇒ ‖χ‖1 ≤ δK‖η‖1,

and using the Triangle Inequality and the Projection Inequality, we have

‖C − Ch‖1 ≤ ‖χ‖1 + ‖η||1

≤ δK‖η‖1 + ‖η‖1

≤ K(1 + δ)‖η‖1

≤ Khk(1 + δ)‖C‖k+1.

65

Page 77: Modeling, Analysis, and Simulation of Adsorption in ...

Chapter 5

Time-Dependent Linear Analysis

In the linear analysis, we consider similar assumptions as with the steady state case

with additional assumptions on the parameters associated with the time derivative terms:

ω and ρs.

(B1) ω and ρs are constant in time and space. [50]

(B2) u is nonzero, independent of time [34] and essentially bounded in space [147, 34] with

∇ · u = 0 [34].

(B3) D = [dij ]i,j=1...n is symmetric positive definite [4, 50], independent of time (since u

is assumed independent of time), and it and its derivatives are essentially bounded

in space [4, 147, 34]; specifically, we notate the boundedness of D using ‖D‖∞ ≤ β1,∣∣∣ ∂∂xidij∣∣∣ ≤ β2, for all i, j.

(B4) C is nonnegative [50, 9, 48], and C = 0 on ΓD.

Again, the assumption that C = 0 on ΓD is not always true in practice. However, the

assumption is still reasonable as a transformation can be applied to C in any case when a

nonzero Dirichlet boundary condition arises to give a solution that is zero on ΓD.

66

Page 78: Modeling, Analysis, and Simulation of Adsorption in ...

5.1 Case 1: Constant Isotherm

We first state and prove a priori stability bounds and a priori error estimates for

the case of constant adsorption. In the case of constant adsorption, q(C) = K with K ≥ 0.

Although the sign of K does not change the error analysis, we assume K is nonnegative

since we wish the isotherm to represent a sink term. Notice that K may be zero which

would correspond to the case in which no isotherm is present. Since q is constant, then

∂q∂t = 0, and therefore the variational formulation given in (3.2.5)-(3.2.6) simplifies to the

following: Find C ∈ H2(Ω) such that

(ω∂C

∂t, v

)+ (u · ∇C, v) + δ (u · ∇C,u · ∇v) + (D∇C,∇v)

− δ(∇ · (D∇C),u · ∇v) = (f, v) + δ (f,u · ∇v)

(5.1.1)

for all v ∈ V .

5.1.1 Continuous in Time Analysis

The semi-discrete streamline diffusion formulation with constant adsorption is as

follows: Find Ch ∈ Vh such that

(ω∂Ch∂t

, vh

)+ (u · ∇Ch, vh) + δ (u · ∇Ch,u · ∇vh) + (D∇Ch,∇vh)

− δ(∇ · (D∇Ch),u · ∇vh) = (f, vh) + δ (f,u · ∇vh)

(5.1.2)

for all vh ∈ Vh.

The first theorem in this section gives the a priori stability bound for the case of

constant adsorption.

Theorem 5.1.1. (Stability Bound) Assume that (B1) - (B4) are satisfied and the

semi-discrete SUPG formulation with constant adsorption given by (5.1.2) has a solution

Ch ∈ L∞(0, T,H1(Ω)) with f ∈ L2(0, T ;L2(Ω)). Then there exists a positive constant K

67

Page 79: Modeling, Analysis, and Simulation of Adsorption in ...

independent of h such that

‖Ch‖2L∞(0,T ;L2(Ω)) +

∫ T

0

(‖u · ∇Ch‖20 + ‖∇Ch‖20

)dt ≤ K

[‖C0‖20 + (1 + δ2)‖f‖2L2(0,T ;L2(Ω))

].

Proof. Taking vh = Ch in (5.1.2), we obtain

(ω∂Ch∂t

, Ch

)+ aδ(Ch, Ch) = (f, Ch) + δ (f,u · ∇Ch) (5.1.3)

where for simplicity we have written the the upwinded convective and diffusive terms as the

bilinear form aδ defined in (4.2.1). We rewrite the first term in (5.1.3) as

(ω∂Ch∂t

, Ch

)= ω

(∂Ch∂t

, Ch

)=ω

2

d

dt‖Ch‖20 ,

and rewrite and obtain lower bounds on the terms in aδ using the assumptions along with

the Cauchy-Schwarz, Young’s, and Poincare’s Inequalities as follows:

(u · ∇Ch, Ch) =

∫Ω

u · (∇Ch)ChdΩ

=1

2

∫Ω

u · ∇(Ch2)dΩ

=1

2

∫Γ(Ch)2 (u · n) ds−

∫Ω

(∇ · u)︸ ︷︷ ︸=0

(Ch)2dΩ

=

1

2

∫ΓD

(Ch)2︸ ︷︷ ︸=0

(u · n)ds+1

2

∫ΓN

(Ch)2 (u · n)︸ ︷︷ ︸≥0

ds

≥ 0

(u · ∇Ch, δu · ∇Ch) = δ‖u · ∇Ch‖20

68

Page 80: Modeling, Analysis, and Simulation of Adsorption in ...

(D∇Ch,∇Ch) = (D1/2∇Ch,D1/2∇Ch)

= ‖D1/2∇Ch‖20

≥ λ‖∇Ch‖20

δ (∇ · (D∇Ch),u · ∇Ch) ≤(∇ ·D : ∇Ch + D : ∇2Ch, δu · ∇Ch

)≤β2δ ‖∇Ch‖0 ‖u · ∇Ch‖0 + β1δ

∥∥∇2Ch∥∥

0‖u · ∇Ch‖0

≤β2δKih−1 ‖Ch‖0 ‖u · ∇Ch‖0 + β1δKih

−1 ‖∇Ch‖0 ‖u · ∇Ch‖0

≤ β22

2ε1‖Ch‖20 +

δ2K2i h−2

2ε1‖u · ∇Ch‖20

+1

2ε2‖∇Ch‖20 +

δ2β21K

2i h−2

2ε2‖u · ∇Ch‖20

≤β22λ

4β21

‖Ch‖20 +δ2β2

1K2i h−2

λ‖u · ∇Ch‖20

4‖∇Ch‖20 +

δ2β21K

2i h−2

λ‖u · ∇Ch‖20

⇒ −δ (∇ · (D∇Ch),u · ∇Ch) ≥− β22λ

4β21

‖Ch‖20 −δ2β2

1K2i h−2

λ‖u · ∇Ch‖20

− λ

4‖∇Ch‖20 −

δ2β21K

2i h−2

λ‖u · ∇Ch‖20

We also obtain upper bounds on the terms on the right in (5.1.3) using Cauchy-Schwarz

and Young’s Inequalities:

(f, Ch + δu · ∇Ch) ≤ ‖f‖0‖Ch + δu · ∇Ch‖0

≤ ‖f‖0‖Ch‖0 + δ‖f‖0‖u · ∇Ch‖0

≤ ‖f‖20 +1

4‖Ch‖20 +

δ2

2ε‖f‖20 +

1

2ε‖u · ∇Ch‖20

≤ ‖f‖20 +1

4‖Ch‖20 +

δ2

2ε‖f‖20 +

‖u‖2∞2

ε‖∇Ch‖20

≤ ‖f‖20 +1

4‖Ch‖20 +

δ2‖u‖2∞λ

‖f‖20 +λ

4‖∇Ch‖20.

69

Page 81: Modeling, Analysis, and Simulation of Adsorption in ...

Using the above bounds in (5.1.3) and combining like terms gives

ω

2

d

dt‖Ch‖20 + δ

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇Ch‖20 +

λ

2‖∇Ch‖20

≤ 1

4

(1 +

β22λ

β21

)‖Ch‖20 +

(1 +

δ2‖u‖2∞λ

)‖f‖20.

(5.1.4)

Multiplying through by 2ω and integrating from 0 to t we obtain

‖Ch(t)‖20 +2

ω

[∫ t

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇Ch(s)‖20ds+

∫ t

0

λ

4‖∇Ch(s)‖20ds

]≤ 1

ω

∫ t

0

1

2

(1 +

β22λ

β21

)‖Ch(s)‖20ds+ ‖Ch(0)‖20 +

2

ω

∫ t

0

(1 +

δ2‖u‖2∞λ

)‖f‖20ds.

(5.1.5)

To ensure positivity of the terms on the left, we choose δ appropriately. The choice

of δ = O(h2) seems necessary to control the second term on the left. However recall that δ

is defined to be 0 when β1 > h so that the terms contributing to the second component do

not arise [50]. Consequently, we assume β1 ≤ h and use this assumption when choosing δ.

Specifically, we choose δ = Kh such that

K <λ

2β1K2i

,

then

2δβ21K

2i h−2

λ≤ 2Kβ2

1K2i h−1

λ≤ 2β1K

2i

λK < 1.

Then Continuous Gronwall’s Inequality (Lemma 3.4.4) applied to (5.1.5) implies

‖Ch(t)‖20 +2K

ω

[∫ t

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇Ch(s)‖20ds+

∫ t

0

λ

4‖∇Ch(s)‖20ds

]≤ K

[‖Ch(0)‖20 +

2

ω

∫ t

0

(1 +

δ2‖u‖2∞λ

)‖f‖20ds

](5.1.6)

where

K = max0≤t≤T

exp

(t

(1 +

β22λ

β21

))= exp

(T

(1 +

β22λ

β21

)).

Since the coefficients on the left are positive, the coefficients on the right are finite, and

70

Page 82: Modeling, Analysis, and Simulation of Adsorption in ...

Ch(0) = C0, (5.1.6) implies

‖Ch(t)‖20 +

∫ t

0‖u · ∇Ch(s)‖20ds+

∫ t

0‖∇Ch(s)‖20ds

≤ K[‖C(0)‖20 + (1 + δ2)

∫ t

0‖f‖20ds

].

(5.1.7)

As (5.1.7) is true for any 0 ≤ t ≤ T , then we have

‖Ch‖2L∞(0,T ;L2(Ω)) +

∫ T

0‖u · ∇Ch‖20dt+

∫ T

0‖∇Ch‖20dt

≤ K[‖C(0)‖20 + (1 + δ2)‖f‖2L2(0,T ;L2(Ω))

].

where K is a positive constant independent of h.

The next theorem gives an a priori error estimate for the case of constant adsorption.

Theorem 5.1.2. (Error Estimate) Suppose that (B1) - (B4) are satisfied. Assume also

that the variational formulation with constant adsorption given by (5.1.1) has an exact

solution C ∈ H1(0, T,Hk+1(Ω)) and Ch solves the semi-discrete SUPG formulation with

constant adsorption given by (5.1.2). Then there exists a positive constant K independent

of h such that

‖C − Ch‖L2(0,T ;H1(Ω)) ≤ K

(h

∥∥∥∥∂C∂t∥∥∥∥L2(0,T ;H1(Ω))

+ hk(1 + δ)‖C‖L2(0,T ;Hk+1(Ω))

).

Proof. Subtracting (5.1.2) from (5.1.1), we obtain

ω

(∂e

∂t, vh

)+ (u · ∇e, vh) + δ(u · ∇e,u · ∇vh) + (D∇e,∇vh)

− δ(∇ · (D∇e),u · ∇vh) = 0

(5.1.8)

where e = C − Ch. Now let C be the elliptic projection the exact solution C as defined in

71

Page 83: Modeling, Analysis, and Simulation of Adsorption in ...

Lemma 3.4.8, and define

η := C − C, χ := Ch − C.

Then (5.1.8) can be written as

ω

(∂χ

∂t, vh

)+ (u · ∇χ, vh) + δ(u · ∇χ, vh) + (D∇χ,∇vh)− δ(D∆χ,u · ∇vh)

= ω

(∂η

∂t, vh

)+ (u · ∇η, vh) + δ(u · ∇η, vh) + (D∇η,∇vh)− δ(D∆η,u · ∇vh).

(5.1.9)

Now using the definition of the elliptic projection, (D∇η,∇vh) = 0 and letting vh = χ in

(5.1.9), we obtain

ω

(∂χ

∂t, χ

)+ (u · ∇χ, χ) + δ(u · ∇χ,u · ∇χ) + (D∇χ,∇χ)

− δ(∇ · (D∇χ),u · ∇χ) = ω

(∂η

∂t, χ

)+ (u · ∇η, χ)

+ δ(u · ∇η,u · ∇χ)− δ(∇ · (D∇η),u · ∇χ).

Again, we rewrite the first terms as

ω

(∂χ

∂t, χ

)=ω

2

d

dt‖χ‖20.

Lower bounds for the terms on the left, obtained in a similar manner to the bounds in the

proof of stability, are given below.

(u · ∇χ, χ) ≥ 0

δ(u · ∇χ,u · ∇χ) = δ‖u · ∇χ‖20

(D∇χ,∇χ) ≥ λ‖∇χ‖2

72

Page 84: Modeling, Analysis, and Simulation of Adsorption in ...

−δ(∇ · (D∇χ),u · ∇χ) ≥− β22λ

2β21

‖χ|20 −δ2β2

1K2i h−2

2λ‖u · ∇χ‖20

− λ

6‖∇χ‖20 −

3δ2β21K

2i h−2

2λ‖u · ∇χ‖20

We obtain upper bounds for the terms on the right in the expected manner with

the use of Cauchy-Schwarz and Young’s Inequalities:

ω

(∂η

∂t, χ

)≤ ω

∥∥∥∥∂η∂t∥∥∥∥

0

‖χ‖0

≤ ω

2

∥∥∥∥∂η∂t∥∥∥∥2

0

2‖χ‖20

(u · ∇η, χ) ≤ ‖u · ∇η‖0‖χ‖0

≤ 1

2‖u · ∇η‖20 +

1

2‖χ‖20

δ(u · ∇η,u · ∇χ) ≤ δ‖u · ∇η‖0‖u · ∇χ‖0

≤ δ2

2ε1‖u · ∇η‖20 +

1

2ε1‖u · ∇χ‖20

≤ δ2

4ε1‖u · ∇η‖20 + ε1‖u‖2∞‖u · ∇χ‖20

≤ 3δ2‖u‖2∞2λ

‖u · ∇η‖20 +λ

6‖∇χ‖20

−δ(∇ · (D∇η),u · ∇χ) ≤ |(δ∇ · (D∇η) ,u · ∇χ)|

≤∥∥δ ((∇ ·D) · (∇η) +

(D∇2η

))∥∥0‖u · ∇χ‖0

≤ δ2

2ε2

∥∥(∇ ·D) · (∇η) +(D∇2η

)∥∥2

0+

1

2ε2 ‖u · ∇χ‖20

≤ 3δ2‖u‖2∞2λ

[β2

2 ‖∇η‖20 + β2

1

∥∥∇2η∥∥2

0

]+λ

6‖∇χ‖20

73

Page 85: Modeling, Analysis, and Simulation of Adsorption in ...

Using these bounds in (5.1.10) and combining like terms gives

ω

2

d

dt‖χ‖20+δ

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇χ‖20 +

λ

2‖∇χ‖20

≤1

2

(ω + 1 +

β22λ

β21

)‖χ‖20 +

ω

2

∥∥∥∥∂η∂t∥∥∥∥2

0

+1

2

(1 +

3δ2‖u‖2∞λ

)‖u · ∇η‖20 (5.1.10)

+3δ2‖u‖2∞

[β2

2 ‖∇η‖20 + β2

1

∥∥∇2η∥∥2

0

].

We then multiply (5.1.10) by 2ω and integrate from zero to T to obtain

‖χ(t)‖20+2

ω

[ ∫ T

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇χ(s)‖20ds+

∫ T

0

λ

2‖∇χ(s)‖20ds

]

≤ 1

ω

∫ T

0

(ω + 1 +

β22λ

β21

)‖χ(s)‖20ds+

1

ω

[∫ T

∥∥∥∥∂η∂t (s)

∥∥∥∥2

0

ds (5.1.11)

+

∫ T

0

(1 +

3δ2‖u‖2∞λ

)‖u · ∇η(s)‖20ds

+

∫ T

0

3δ2‖u‖2∞λ

[β2

2 ‖∇η(s)‖20 + β21

∥∥∇2η(s)∥∥2

0

]ds

]+ ‖χ(0)‖20.

To ensure positivity of the terms on the left, we choose δ appropriately. As with the

proof of stability, the choice of δ = O(h2) seems necessary to control the second term on the

left. However recall that δ is defined to be 0 when β1 > h so that the terms contributing

to the second component do not arise [50]. Consequently, we assume β1 ≤ h and use this

assumption when choosing δ. Specifically, we choose δ = Kh such that

K <λ

2β1K2i

,

then

2δβ21K

2i h−2

λ≤ 2Kβ2

1K2i h−1

λ≤ 2β1K

2i

λK < 1.

Then using the fact that ‖χ(0)‖0 = 0 and applying Continuous Gronwall’s Inequality

74

Page 86: Modeling, Analysis, and Simulation of Adsorption in ...

(Lemma 3.4.4) to (5.1.11) gives us

‖χ(t)‖20+2K

ω

[ ∫ T

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇χ(s)‖20ds+

∫ T

0

λ

2‖∇χ(s)‖20ds

]

≤Kω

[∫ T

∥∥∥∥∂η∂t (s)

∥∥∥∥2

0

ds+

∫ T

0

(1 +

3δ2‖u‖2∞λ

)‖u · ∇η(s)‖20ds

+

∫ T

0

3δ2‖u‖2∞λ

[β2

2 ‖∇η(s)‖20 + β21

∥∥∇2η(s)∥∥2

0

]ds

].

Consequently, we have

‖χ(t)‖20 +Km

∫ T

0

(‖u · ∇χ(s)‖20 + ‖∇χ(s)‖20

)ds

≤ K∫ T

0

∥∥∥∥∂η∂t (s)

∥∥∥∥2

0

ds+KM

((1 + δ2)

∫ T

0‖u · ∇η(s)‖20ds

+ δ2

∫ T

0

(‖∇η(s)‖20 + ‖∇2η(s)‖20

)ds

).

(5.1.12)

Note that the second derivative term on the right can be bounded by its gradient. This terms

is associated with the upwinded diffusion term which is defined locally over the interior of

the mesh triangulation [34, 82]. Thus a local approximation property can be applied as in

[80], and the associated δ2 can serve to bound the h−2 appearing from the application of

the approximation property. Therefore, (5.1.12) implies

Km

∫ T

0

(‖u · ∇χ‖20 + ‖∇χ‖20

)dt ≤K

∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt (5.1.13)

+ (1 + δ2)KM

∫ T

0

(‖u · ∇η‖20 + ‖∇η‖20

)dt

⇒∫ T

0‖∇χ‖20dt ≤K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+ (1 + δ2)

∫ T

0‖∇η‖20dt

)(5.1.14)

75

Page 87: Modeling, Analysis, and Simulation of Adsorption in ...

and also

‖χ‖20 ≤K∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+ (1 + δ2)KM

∫ T

0

(‖u · ∇η‖20 + ‖∇η‖20

)dt

⇒∫ T

0‖χ‖20dt ≤K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+ (1 + δ2)

∫ T

0‖∇η‖20dt

). (5.1.15)

where K is dependent on T . Consequently,

∫ T

0‖χ‖21dt ≤ K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+ (1 + δ2)

∫ T

0‖∇η‖20dt

). (5.1.16)

By the triangle inequality,

‖e‖1 =‖η − χ‖21

≤ (‖η‖1 + ‖χ‖1)2

≤K(‖η‖21 + ‖χ‖21

)⇒∫ T

0‖e‖21dt ≤K

(∫ T

0‖η‖21dt+

∫ T

0‖χ‖21dt

). (5.1.17)

Combining (5.1.16) with (5.1.17), we have

∫ T

0‖e‖21dt ≤K

(∫ T

0‖η‖21dt+

∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+ (1 + δ2)

∫ T

0‖∇η‖20dt

). (5.1.18)

Since C ∈ H1(0, T ;H1(Ω)) implies ∂C∂t ∈ L2(0, T ;H1(Ω)), we can apply the Projection

Inequality to ‖η‖0, ‖∇η‖0, and∥∥∥∂η∂t ∥∥∥0

:

‖η‖20 ≤K2ph

2k‖C‖2k (5.1.19)

‖∇η‖20 ≤K2ph

2k‖∇C‖2k (5.1.20)∥∥∥∥∂η∂t∥∥∥∥2

0

≤K2ph

2

∥∥∥∥∂C∂t∥∥∥∥2

1

(5.1.21)

76

Page 88: Modeling, Analysis, and Simulation of Adsorption in ...

Therefore, we have

∫ T

0‖e‖21dt ≤K

(h2

∫ T

0

∥∥∥∥∂C∂t∥∥∥∥2

1

dt+ h2k

∫ T

0‖C‖2kdt+ (1 + δ2)h2k

∫ T

0‖∇C‖2kdt

)

≤K

(h2

∫ T

0

∥∥∥∥∂C∂t∥∥∥∥2

1

dt+ h2k(1 + δ2)

∫ T

0‖C‖2k+1dt

)

which implies

‖C − Ch‖2L2(0,T ;H1(Ω)) ≤K

(h2

∥∥∥∥∂C∂t∥∥∥∥2

L2(0,T ;H1(Ω))

+ h2k(1 + δ2)‖C‖2L2(0,T ;Hk+1(Ω))

).

Consequently, we have

‖C − Ch‖L2(0,T ;H1(Ω)) ≤K

(h

∥∥∥∥∂C∂t∥∥∥∥L2(0,T ;H1(Ω))

+ hk(1 + δ)‖C‖L2(0,T ;Hk+1(Ω))

)

where K is a positive constant independent of h.

5.1.2 Fully Explicit Analysis

For the fully explicit formulation, we show the existence of a solution is simple given

the assumptions on the parameters. We also show that, as penalty for the fully explicit

evaluation, strict constraints on the time step are required for obtaining an a priori stability

bound and an a priori error estimate in the case of a constant adsorption.

In the case of constant adsorption, the fully explicit discrete formulation in (3.3.5)-

(3.3.6) simplifies to the following: For n = 1, ..., N , find Cnh ∈ Vh such that

(ω dt(Cnh ), vh) +

(u · ∇Cn−1

h , vh))

+ δ(u · ∇Cn−1

h ,u · ∇vh)

+(D∇Cn−1

h ,∇vh)

− δ(∇ · (D∇Cn−1

h ),u · ∇vh)

=(fn−1, vh

)+ δ

(fn−1,u · ∇vh

) (5.1.22)

for all vh ∈ Vh. The exact variational formulation is again given by (5.1.1).

We begin by showing that (5.1.22) is uniquely solvable for Cnh at each time step n.

Lemma 5.1.3. (Solvability) Assume (B1) - (B4) are satisfied. Then there exists a unique

77

Page 89: Modeling, Analysis, and Simulation of Adsorption in ...

solution Cnh ∈ Vh satisfying (5.1.22).

Proof. By assumption, ω, u, and D are not dependent on Cnh . Consequently, (5.1.22) is a

fully explicit linear equation, and the existence and uniqueness of Cnh is trivial. At each

n, the system arising from (5.1.22) is a square diagonal system of linear equations which

implies the existence of a unique solution Cnh ∈ Vh.

We next state and prove an a priori stability bound.

Theorem 5.1.4. (Stability Bound) Suppose the assumptions of Lemma 5.1.3 are satisfied

so that the fully explicit SUPG formulation with constant adsorption given by (5.1.22) has

a solution Cnh ∈ L2(0, T ;H1(Ω)) with f ∈ L2(0, T ;L2(Ω)). Also assume that ∆t is on the

order of h2. Then, there exists a positive constant K independent of h and ∆t such that for

all N > 0

‖CNh ‖20 + ∆t

N∑n=1

(‖u · ∇Cnh‖

20 + ‖∇Cnh‖

20

)≤ K

(‖C0

h‖20 + (1 + δ2)∆t

N∑n=1

‖fn−1‖20

)

Proof. To show the stability of (5.1.22), we will let v = Cnh which gives

(ω dt(Cnh ), Cnh ) + aδ(C

n−1h , Cnh ) =

(fn−1, Cnh + δu · ∇Cnh

)(5.1.23)

where for simplicity we have written the the upwinded convective and diffusive terms as the

bilinear form aδ defined in (4.2.1). Using

(x− y)x =1

2(x2 − y2 + (x− y)2),

we obtain

(ω dt(Cnh ), Cnh ) =

ω

∆t

(Cnh − C

n−1h , Cnh

)=

ω

2∆t

[(Cnh , C

nh )−

(Cn−1h , Cn−1

h

)+(Cnh − C

n−1h , Cnh − C

n−1h

)]=

ω

2∆t

[‖Cnh‖

20 − ‖Cn−1

h ‖20 + ‖Cnh − Cn−1h ‖20

]. (5.1.24)

78

Page 90: Modeling, Analysis, and Simulation of Adsorption in ...

Combining (5.1.23) and (5.1.24) and adding and subtracting aδ(Cnh , C

nh ) gives

ω

2∆t

[‖Cnh‖

20 − ‖Cn−1

h ‖20 + ‖Cnh − Cn−1h ‖20

]+ aδ(C

nh , C

nh )

= aδ(Cnh − C

n−1h , Cnh ) + (fn−1, Cnh + δu · ∇Cnh )

(5.1.25)

Lower bounds on the terms in aδ(Cnh , C

nh ) are obtained using the assumptions, Cauchy-

Schwarz Inequality, and Young’s Inequality and are shown below.

(u · ∇Cnh , Cnh ) =

∫Ω

u · (∇Cnh )CnhdΩ

=1

2

∫Ω

u · ∇(Cnh2)dΩ

=1

2

∫Γ(Cnh )2 (u · n) ds−

∫Ω

(∇ · u)︸ ︷︷ ︸=0

(Cnh )2dΩ

=

1

2

∫ΓD

(Cnh )2︸ ︷︷ ︸=0

(u · n)ds+1

2

∫ΓN

(Cnh )2 (u · n)︸ ︷︷ ︸≥0

ds

≥ 0

(u · ∇Cnh , δu · ∇Cnh ) = δ‖u · ∇Cnh‖20

(D∇Cnh ,∇Cnh ) = (D1/2∇Cnh ,D1/2∇Cnh )

= ‖D1/2∇Cnh‖20

≥ λ‖∇Cnh‖20

79

Page 91: Modeling, Analysis, and Simulation of Adsorption in ...

δ (∇ · (D∇Cnh ),u · ∇Cnh ) ≤(∇ ·D : ∇Cnh + D : ∇2Cnh , δu · ∇Cnh

)≤β2δ ‖∇Cnh‖0 ‖u · ∇C

nh‖0 + β1δ

∥∥∇2Cnh∥∥

0‖u · ∇Cnh‖0

≤β2δKih−1 ‖Cnh‖0 ‖u · ∇C

nh‖0 + β1δKih

−1 ‖∇Cnh‖0 ‖u · ∇Cnh‖0

≤ β22

2ε1‖Cnh‖

20 +

δ2K2i h−2

2ε1‖u · ∇Cnh‖

20

+1

2ε2‖∇Cnh‖

20 +

δ2β21K

2i h−2

2ε2‖u · ∇Cnh‖

20

≤β22λ

2β21

‖Cnh‖20 +

δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

10‖∇Cnh‖

20 +

5δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

⇒ −δ (∇ · (D∇Cnh ),u · ∇Cnh ) ≥− β22λ

2β21

‖Cnh‖20 −

δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

− λ

10‖∇Cnh‖

20 −

5δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

Upper bounds on the terms in aδ(Cnh −C

n−1h , Cnh ) are found using similar techniques

and are shown below.

(u · ∇(Cnh − Cn−1h ), Cnh ) ≤ ‖u‖∞‖‖∇(Cnh − C

n−1h )‖0‖Cnh‖0

≤ 1

2‖u‖2∞‖∇(Cnh − C

n−1h )‖20 +

1

2‖Cnh‖

20

≤ 1

2‖u‖2∞K2

i h−2‖Cnh − C

n−1h ‖20 +

1

2‖Cnh‖

20

δ(u · ∇(Cnh − Cn−1h ),u · ∇Cnh ) ≤ δ‖u‖2∞‖∇(Cnh − C

n−1h )‖0‖∇Cnh‖0

≤ δ2‖u‖4∞2ε

‖∇(Cnh − Cn−1h )‖20 +

1

2ε‖∇Cnh‖

20

=5δ2‖u‖4∞

2λ‖∇(Cnh − C

n−1h )‖20 +

λ

10‖∇Cnh‖

20

≤ 5δ2‖u‖4∞K2i h−2

2λ‖Cnh − C

n−1h ‖20 +

λ

10‖∇Cnh‖

20

80

Page 92: Modeling, Analysis, and Simulation of Adsorption in ...

(D∇(Cnh − Cn−1h ),∇Cnh ) ≤ β1‖∇(Cnh − C

n−1h )‖0‖∇Cnh‖0

≤ β21

2ε‖∇(Cnh − C

n−1h )‖20 +

1

2ε‖∇Cnh‖

20

=5β2

1

2λ‖∇(Cnh − C

n−1h )‖20 +

λ

10‖∇Cnh‖

20

≤ 5β21K

2i h−2

2λ‖Cnh − C

n−1h ‖20 +

λ

10‖∇Cnh‖

20

−δ(∇ · (D∇(Cnh − Cn−1h )),u · ∇Cnh ) ≤δβ2‖∇(Cnh − C

n−1h )‖0‖u · ∇Cnh‖0

+ δβ1‖∇2(Cnh − Cn−1h )‖0‖u · ∇Cnh‖0

≤δβ2‖u‖∞‖∇(Cnh − Cn−1h )‖0‖∇Cnh‖0

+ δβ1Kih−1‖∇(Cnh − C

n−1h )‖0‖u · ∇Cnh‖0

≤δ2β2

2‖u‖2∞2ε1

‖∇(Cnh − Cn−1h )‖20 +

1

2ε1‖∇Cnh‖

20

+1

2ε2‖∇(Cnh − C

n−1h )‖20 +

δ2β21K

2i h−2

2ε2‖u · ∇Cnh‖

20

≤5δ2β22‖u‖2∞2λ

‖∇(Cnh − Cn−1h )‖20 +

λ

10‖∇Cnh‖

20

2‖∇(Cnh − C

n−1h )‖20 +

δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

≤5δ2β22‖u‖2∞K2

i h−2

2λ‖Cnh − C

n−1h ‖20 +

λ

10‖∇Cnh‖

20

+λK2

i h−2

2‖Cnh − C

n−1h ‖20 +

δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

The terms involving f are bounded similarly as follows:

(fn−1, Cnh + δu · ∇Cnh ) ≤‖fn−1‖0‖Cnh‖0 + δ‖fn−1‖0‖u · ∇Cnh‖0

≤1

2‖fn−1‖20 +

1

2‖Cnh‖

20 +

δ2

2ε‖fn−1‖20 +

1

2ε‖u · ∇Cnh‖

20

≤1

2‖fn−1‖20 +

1

2‖Cnh‖

20 +

δ2

2ε‖fn−1‖20 +

1

2ε‖u‖2∞‖∇Cnh‖

20

=1

2‖fn−1‖20 +

1

2‖Cnh‖

20 +

5δ2‖u‖2∞2λ

‖fn−1‖20 +λ

10‖∇Cnh‖

20.

81

Page 93: Modeling, Analysis, and Simulation of Adsorption in ...

Putting all these bounds together with (5.1.25) and combining like terms gives

ω

2∆t

(‖Cnh‖

20 − ‖Cn−1

h ‖20)

+

2∆t− 5β2

1 + λ(‖u‖2∞ + λ) + 5δ2‖u‖2∞(‖u‖2∞ + β22)

2λK2i h−2

]‖Cnh − C

n−1h ‖20

2‖∇Cnh‖

20 + δ

(1− 7δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

≤(

1 +β2

2β21

)‖Cnh‖

20 +

1

2

(1 +

5δ2‖u‖2∞λ

)‖fn−1‖20

Multiplying through by 2∆tω and summing from n = 1 to n = N , we obtain

‖CNh ‖20 + ∆tN∑n=1

2

ω

2‖∇Cnh‖

20 + δ

(1− 7δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

)+ ∆t

N∑n=1

[1

∆t− 5β2

1 + λ(‖u‖2∞ + λ) + 5δ2‖u‖2∞(‖u‖2∞ + β22)

ωλK2i h−2

]‖Cnh − C

n−1h ‖20

≤ ∆tN∑n=1

2

ω

(1 +

β22λ

2β21

)‖Cnh‖

20 + ∆t

N∑n=1

1

ω

(1 +

5δ2‖u‖2∞λ

)‖fn−1‖20 + ‖C0

h‖20

(5.1.26)

To ensure positivity of the terms on the left, we choose δ and ∆t appropriately. For

the third term on the left, we choose ∆t ∼ O(h2); specifically we choose

∆t <ωλ(

5β21 + λ(‖u‖2∞ + λ) + 5δ2‖u‖2∞(‖u‖2∞ + β2

2))K2i

h2

so that

1

∆t− 5β2

1 + λ(‖u‖2∞ + λ) + 5δ2‖u‖2∞(‖u‖2∞ + β22)

ωλK2i h−2 > 0.

The choice of δ = O(h2) seems necessary to control the second term on the left.

However recall that δ is defined to be 0 when β1 > h so that the terms contributing to

the second component do not arise [50]. Consequently, we assume β1 ≤ h and use this

assumption to help control the h2. Specifically, we choose δ = Kh such that

K <2λ

7β1K2i

,

82

Page 94: Modeling, Analysis, and Simulation of Adsorption in ...

then

7δβ21K

2i h−2

2λ≤ 7Kβ2

1K2i h−1

2λ≤ 7β1K

2i

2λK < 1.

Then applying Discrete Gronwall’s Inequality (Lemma 3.4.5) to (5.1.26) gives

‖CNh ‖20 + ∆tN∑n=1

2

ω

2‖∇Cnh‖

20 + δ

(1− 7δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

)+ ∆t

N∑n=1

[1

∆t− 5β2

1 + λ(‖u‖2∞ + λ) + 5δ2‖u‖2∞(‖u‖2∞ + β22)

ωλK2i h−2

]‖Cnh − C

n−1h ‖20

≤ K

(‖C0

h‖20 + ∆t

N∑n=1

1

ω

(1 +

5δ2‖u‖2∞λ

)‖fn−1‖20

)(5.1.27)

where K = expλT/ω. Since the coefficients on the left are positive and the coefficients

on the right are finite, we drop the second term on the left and manipulate the constants

in (5.1.27) to obtain for any N > 0

‖CNh ‖20 + ∆tN∑n=1

(‖u · ∇Cnh‖

20 + ‖∇Cnh‖

20

)≤ K

(‖C0

h‖20 + (1 + δ2)∆tN∑n=1

‖fn−1‖20

)

where K is a positive constant independent of h and ∆t.

Next, we obtain an a priori error estimate in a similar fashion as the a priori stability

bound and the error estimate for the semi-continuous case. For the error analysis, we use

the following notation:

Cn denotes the exact solution at time tn, i.e. Cn = C(tn),

Cnh denotes the approximate solution at time tn, i.e. Cnh = Ch(tn),

Cn denotes the elliptic projection of Cn as defined by Lemma 3.4.8.

Theorem 5.1.5. (Error Estimate) Suppose the assumptions of Lemma 5.1.3 are satisfied

so that the fully explicit SUPG formulation with constant adsorption given by (5.1.22) has

a solution Cnh . Assume also that the variational formulation with constant adsorption given

83

Page 95: Modeling, Analysis, and Simulation of Adsorption in ...

by (5.1.1) has an exact solution C ∈ H2(0, T,Hk+1(Ω)). In addition assume that ∆t is on

the order of h2. Then there exists a positive constant K independent of h and ∆t such that

‖C − Ch‖L2(0,T ;H1(Ω)) ≤ K

(hk(1 + δ) ‖C‖H1(0,T ;Hk+1(Ω)) + ∆t

∥∥∥∥∂2C

∂t2

∥∥∥∥L2(0,T ;L2(Ω))

).

Proof. Subtracting (5.1.22) from (5.1.1) we obtain

ω

(∂C

∂t− dt(Cnh ), vh

)+(u · ∇en−1, vh

)+ δ

(u · ∇en−1,u · ∇vh

)+(D∇en−1,∇vh

)− δ

(∇ · (D∇en−1),u · ∇vh

)= 0

where en = Cn − Cnh . Consequently,

ω (dt(χn), vh) +aδ(χ

n−1, vh) = ω (rn, vh) + ω (dt(ηn), vh) + aδ(η

n−1, vh) (5.1.28)

where rn is the residual in the derivative approximation at time tn

rn =∂Cn

∂t− dt(Cn)

and χn and ηn are defined as

ηn := Cn − Cn, χn := Cnh − Cn

with Cn the elliptic projection of Cn defined in Lemma 3.4.8.

Now choosing vh = χn, we use

(x− y)x =1

2

(x2 − y2 + (x− y)2

)to rewrite the first term on the left hand side of (5.1.28):

ω (dt(χn), χn) =

ω

2∆t

(‖χn‖20 − ‖χn−1‖20 + ‖χn − χn−1‖20

).

84

Page 96: Modeling, Analysis, and Simulation of Adsorption in ...

Therefore, (5.1.28) becomes

ω

2∆t

(‖χn‖20 − ‖χn−1‖20 + ‖|χn − χn−1‖20

)+ aδ(χ

n−1, χn)

= ω (rn, χn) + ω (dt(ηn), χn) + aδ(η

n−1, χn),

and by adding and subtracting aδ(χn, χn) we obtain

ω

2∆t

(‖χn‖20 − ‖χn−1‖20 + ‖|χn − χn−1‖20

)+ aδ(χ

n, χn)

≤ ω (rn, χn) + ω (dt(ηn), χn) + aδ(χ

n − χn−1, χn) + aδ(ηn−1, χn).

(5.1.29)

The lower bound for the terms on the left in (5.1.29) are obtained in a manner

similar to the stability proof and are listed below.

(u · ∇χn, χn) ≥ 0

(u · ∇χn, δu · ∇χn) = δ‖u · ∇χn‖20

(D∇χn,∇χn) ≥ λ‖∇χn‖20

−δ (∇ · (D∇χn),u · ∇χn) ≥− β22λ

2β21

‖χn‖20 −δ2β2

1K2i h−2

2λ‖u · ∇χn‖20

− λ

14‖∇χn‖20 −

7δ2β21K

2i h−2

2λ‖u · ∇χn‖20

Upper bounds on the terms in aδ(χn − χn−1, χn) are also found similarly and are shown

below.

(u · ∇(χn − χn−1), χn) ≤ 1

2‖u‖2∞K2

i h−2‖χn − χn−1‖20 +

1

2‖χn‖20

85

Page 97: Modeling, Analysis, and Simulation of Adsorption in ...

δ(u · ∇(χn − χn−1),u · ∇Cnh ) ≤ 7δ2‖u‖4∞K2i h−2

2λ‖χn − χn−1‖20 +

λ

14‖∇χn‖20

(D∇(χn − χn−1),∇χn) ≤ 7β21K

2i h−2

2λ‖χn − χn−1‖20 +

λ

14‖∇χn‖20

−δ(∇ · (D∇(χn − χn−1)),u · ∇Cnh ) ≤7δ2β22‖u‖2∞K2

i h−2

2λ‖χn − χn−1‖20 +

λ

14‖∇χn‖20

+λK2

i h−2

2‖χn − χn−1‖20 +

δ2β21K

2i h−2

2λ‖u · ∇χn‖20

The rest of the terms on the right of (5.1.29) are bounded using the assumptions,

Cauchy-Schwarz Inequality (3.4.5) and Young’s Inequality (3.4.2) as follows:

ω (rn, χn) + ω (dt(ηn), χn) ≤ ω‖χn‖20 +

ω

2‖dt(ηn)‖20 +

ω

2‖rn‖20,

(u · ∇ηn−1, χn) ≤ 1

2‖u · ∇ηn−1‖20 +

1

2‖χn‖20,

(u · ∇ηn−1, δu · ∇χn) ≤ δ‖u‖∞‖u · ∇ηn−1‖0‖∇χn‖0

≤ δ2‖u‖2∞2ε

‖u · ∇ηn−1‖20 +1

2ε‖∇χn‖20

≤ 7δ2‖u‖2∞2λ

‖u · ∇ηn−1‖20 +λ

14‖∇χn‖20,

86

Page 98: Modeling, Analysis, and Simulation of Adsorption in ...

−(∇ · (D∇ηn−1), δu · ∇χn) ≤δβ2‖∇ηn−1‖0‖u · ∇χn‖0 + δβ1‖∇2ηn−1‖0‖u · ∇χn‖0

≤δβ2‖u‖∞‖∇ηn−1‖0‖∇χn‖0

+ δβ1‖u‖∞‖∇2ηn−1‖0‖∇χn‖0

≤δ2β2

2‖u‖2∞2ε1

‖∇ηn−1‖20 +1

2ε1‖∇χn‖20

+δ2β2

1‖u‖2∞2ε2

‖∇2ηn−1‖20 +1

2ε2‖∇χn‖20

≤7δ2β22‖u‖2∞2λ

‖∇ηn−1‖20 +λ

14‖∇χn‖20

+7δ2β2

1‖u‖2∞2λ

‖∇2ηn−1‖20 +λ

14‖∇χn‖20.

Note by the definition of the elliptic projection (D∇ηn−1,∇χn) = 0.

Then combining the above bounds with (5.1.29) and adding like terms gives

ω

2∆t

(‖χn‖20 − ‖χn−1‖20

)+ δ

(1− 9δ2β2

1K2i h−2

)‖u · ∇χn‖20 +

λ

2‖∇χn‖20

+

2∆t−(

7β21 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β2

2)

)K2i h−2

]‖χn − χn−1‖20

≤(ω + 1 +

β22λ

2β21

)‖χn‖20 +

ω

2‖dt(ηn)‖20 +

ω

2‖rn‖20 (5.1.30)

+1

2

(1 +

7δ2‖u‖2∞λ

)‖u · ∇ηn−1‖20 +

7δ2β22‖u‖2∞2λ

‖∇ηn−1‖20

+7δ2β2

1‖u‖2∞2λ

‖∇2ηn−1‖20

Before continuing, we bound the second and third terms on the right hand side of

(5.1.30) using standard expansions of differentiable functions. Note for any function f ,

f(ti) = f(ti−1) +

∫ ti

ti−1

ft(t)dt (5.1.31)

= f(ti−1) + ∆tft(ti−1) +

∫ ti

ti−1

(ti − t)ftt(t)dt (5.1.32)

87

Page 99: Modeling, Analysis, and Simulation of Adsorption in ...

where ∆t = ti − ti−1. Then (5.1.31) implies

ηn − ηn−1 =

∫ tntn−1

ηt(t)dt ≤

(∫ tntn−1

dt

)1/2(∫ tntn−1

(ηt(t))2 dt

)1/2

= (∆t)1/2

(∫ tntn−1

(ηt(t))2 dt

)1/2

,

which in turn implies

∥∥ηn − ηn−1∥∥2

0≤∫

Ω

(∆t)1/2

(∫ tntn−1

(ηt(t))2 dt

)1/22

= ∆t

∫Ω

(∫ tntn−1

(ηt(t))2 dt

)dΩ

= ∆t

∫ tntn−1

(∫Ω

(ηt(t))2 dΩ

)dt

= ∆t

∫ tntn−1

‖ηt‖20 dt,

and consequently

‖dt(ηn)‖20 ≤1

∆t

∫ tntn−1

‖ηt‖20 dt. (5.1.33)

Also, (5.1.32) implies

rn =C(tn)− C(tn−1)

∆t− Ct(tn−1) =

1

∆t

∫ tntn−1

(tn − t)Ctt(t)dt

88

Page 100: Modeling, Analysis, and Simulation of Adsorption in ...

which implies

‖rn‖20 =

∫Ω

(1

∆t

∫ tntn−1

(tn − t)Ctt(t)dt

)2

≤∫

Ω

1

∆t

(∫ tntn−1

(tn − t)2dt

)1/2(∫ tntn−1

(Ctt(t))2 dt

)1/22

=∆t

3

∫Ω

(∫ tntn−1

(Ctt(t))2 dt

)dΩ

=∆t

3

∫ tntn−1

‖Ctt‖20 dt. (5.1.34)

If we combine (5.1.33) and (5.1.34) with (5.1.30), we obtain

ω

2∆t

(‖χn‖20 − ‖χn−1‖20

)+ δ

(1− 9δ2β2

1K2i h−2

)‖u · ∇χn‖20 +

λ

2‖∇χn‖20

+

2∆t−(

7β21 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β2

2)

)K2i h−2

]‖χn − χn−1‖20

≤(ω + 1 +

β22λ

2β21

)‖χn‖20 +

ω

2

(1

∆t

∫ tntn−1

‖ηt‖20

)(5.1.35)

2

(∆t

3

∫ tntn−1

‖Ctt‖20

)+

1

2

(1 +

7δ2‖u‖2∞λ

)‖u · ∇ηn−1‖20

+7δ2β2

2‖u‖2∞2λ

‖∇ηn−1‖20 +7δ2β2

1‖u‖2∞2λ

‖∇2ηn−1‖20.

89

Page 101: Modeling, Analysis, and Simulation of Adsorption in ...

Then multiplying through by 2∆tω and summing (5.1.35) from n = 1 to n = N gives

‖χN‖20 + ∆tN∑n=1

2

ω

(1− 9δ2β2

1K2i h−2

)‖u · ∇χn‖20 +

λ

2‖∇χn‖20

]+ ∆t

N∑n=1

[1

∆t−(

7β21 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β2

2)

ωλ

)K2i h−2

]‖χn − χn−1‖20

≤ ∆tN∑n=1

2

ω

(ω + 1 +

β22λ

2β21

)‖χn‖20 + ∆t

(1

∆t

∫ T

0‖ηt‖20 dt

)

+ ∆t

(∆t

3

∫ T

0‖Ctt‖20 dt

)+ ∆t

N∑n=1

1

ω

(1 +

7δ2‖u‖2∞λ

)‖u · ∇ηn−1‖20

+ ∆t

N∑n=1

7δ2β22‖u‖2∞ωλ

‖∇ηn−1‖20 + ∆t

N∑n=1

7δ2β21‖u‖2∞ωλ

‖∇2ηn−1‖20 + ‖χ0‖20.

To ensure positivity of the terms on the left, we choose δ and ∆t appropriately. As

with the proof of stability, the choice of δ = O(h2) seems necessary to control the second

term on the left. However recall that δ is defined to be 0 when β1 > h so that the terms

contributing to the second component do not arise [50]. Consequently, we assume β1 ≤ h

and use this assumption when choosing δ. Specifically, we choose δ = Kh such that

K <2λ

9β1K2i

,

then

9δβ21K

2i h−2

2λ≤ 9Kβ2

1K2i h−1

2λ≤ 9β1K

2i

2λK < 1.

For the last term on the left, we choose ∆t ∼ O(h2); specifically we choose

∆t <ωλ(

7β21 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β2

2))K2i

h2

so that

1

∆t− 7β2

1 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β22)

ωλK2i h−2 > 0.

Then applying Discrete Gronwall’s Inequality (Lemma 3.4.5) and using the fact that

90

Page 102: Modeling, Analysis, and Simulation of Adsorption in ...

‖χ0‖0 = 0, we see

‖χN‖20 + ∆tN∑n=1

2

ω

(1− 9δ2β2

1K2i h−2

)‖u · ∇χn‖20 +

λ

2‖∇χn‖20

]+ ∆t

N∑n=1

[1

∆t−(

7β21 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β2

2)

ωλ

)K2i h−2

]‖χn − χn−1‖20

≤ K

[∆t

(1

∆t

∫ T

0‖ηt‖20

)+ ∆t

(∆t

3

∫ T

0‖Ctt‖20

)(5.1.36)

+ ∆tN∑n=1

1

ω

(1 +

7δ2‖u‖2∞λ

)‖u · ∇ηn−1‖20

+ ∆t

N∑n=1

7δ2‖u‖2∞ωλ

(β2

2‖∇ηn−1‖20 + β21‖∇2ηn−1‖20

).

Note that the second derivative term on the right in (5.1.36) can be bounded by its

gradient. This terms is associated with the upwinded diffusion term which is defined locally

over the interior of the mesh triangulation [34, 80, 82]. Thus a local approximation property

can be applied as in [80], and the associated δ2 can serve to bound the h−2 appearing from

the application of the approximation property. Therefore, since the coefficients on the left

are positive and the coefficients on the right are finite, (5.1.36) implies

∆t

N∑n=1

‖∇χn‖20 ≤ K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

+ ∆t(1 + δ2)

N∑n=1

(‖u · ∇ηn−1‖20 + ‖∇ηn−1‖20

)).

(5.1.37)

Applying Lemma 3.4.7 (the consequence of Poincare’s Inequality) to (5.1.37) gives

∆t

N∑n=1

‖∇χn‖21 ≤ K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

+ ∆t(1 + δ2)N∑n=1

(‖u · ∇ηn−1‖20 + ‖∇ηn−1‖20

)) (5.1.38)

Since C ∈ H2(0, T,Hk+1(Ω)) implies ∂C∂t ∈ L

2(0, T ;H1(Ω)), we can apply the pro-

91

Page 103: Modeling, Analysis, and Simulation of Adsorption in ...

jection inequality (Lemma 3.4.8) to ‖η‖0, ‖∇η‖0,∥∥∥∂η∂t ∥∥∥0

as follows:

‖η‖20 ≤ K2ph

2k‖C‖2k (5.1.39)

‖∇η‖20 ≤ K2ph

2k‖∇C‖2k (5.1.40)∥∥∥∥∂η∂t∥∥∥∥2

0

≤ K2ph

2

∥∥∥∥∂C∂t∥∥∥∥2

1

. (5.1.41)

Also by the Triangle Inequality,

∆t

N∑n=1

‖en‖1 =∆t

N∑n=1

‖ηn − χn‖21

≤∆t

N∑n=1

(‖ηn‖1 + ‖χn‖1)2

≤∆t

N∑n=1

K(‖ηn‖21 + ‖χn‖21

)≤∆t

N∑n=1

K(K2ph

2k‖Cn‖2k+1 + 2‖χn‖21). (5.1.42)

Combining (5.1.37) - (5.1.42) gives us

∆tN∑n=1

‖en‖21 ≤ K

(h2k∆t

N∑n=1

‖Cn‖2k+1 + h2

∫ T

0

∥∥∥∥∂C∂t∥∥∥∥2

0

+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

+ h2k(1 + δ2)∆tN∑n=1

‖∇Cn−1‖2k

)

≤ K

(h2k

∫ T

0‖C‖2k+1 + h2

∫ T

0

∥∥∥∥∂C∂t∥∥∥∥2

0

+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

+ h2k(1 + δ2)

∫ T

0‖∇C‖2k

)

≤ K

(h2k(1 + δ2) ‖C‖2H1(0,T ;Hk+1(Ω)) + (∆t)2

∥∥∥∥∂2C

∂t2

∥∥∥∥2

L2(0,T ;L2(Ω))

)(5.1.43)

92

Page 104: Modeling, Analysis, and Simulation of Adsorption in ...

and consequently

‖C − Ch‖L2(0,T ;H1(Ω)) ≤ K

(hk(1 + δ)‖C‖H1(0,T ;Hk+1(Ω) + ∆t

∥∥∥∥∂2C

∂t2

∥∥∥∥L2(0,T ;L2(Ω))

)

where K is a positive constant independent of h and ∆t.

5.1.3 Fully Implicit Analysis

For the fully implicit formulation, we show the existence of a unique solution is only

marginally more difficult than in the fully explicit case. We also show that in contrast to

the fully explicit formulation there is no restriction on the time step for either the stability

bound or error estimate in the cases of constant adsorption.

In the case of constant adsorption, the fully implicit, discrete formulation in (3.3.7)-

(3.3.8) simplifies to the following: For n = 1, ..., N , find Cnh ∈ Vh such that

(ω dt(Cnh ), vh) + (u · ∇Cnh , vh) + δ (u · ∇Cnh ,u · ∇vh) +

(D∇Cn−1

h ,∇vh)

− δ (∇ · (D∇Cnh ),u · ∇vh) = (fn, vh) + δ (fn,u · ∇vh)(5.1.44)

for all vh ∈ Vh. The exact variational formulation is again given by (5.1.1).

We begin by showing that (5.1.44) is uniquely solvable for Cnh at each time step n.

Lemma 5.1.6. (Solvability) Assume (B1) - (B4) are satisfied. Then there exists a unique

solution Cnh ∈ Vh satisfying (5.1.44).

Proof. Choosing vh = Cnh in (5.1.44) and rearranging, we see

a(Cnh , Cnh ) = (fn, Cnh ) + δ (fn, Cnh ) +

ω

∆t

(Cn−1h , Cnh

)where

a(Cnh , Cnh ) =

ω

∆t‖Cnh‖

20 + (u · ∇Cnh , Cnh ) + δ‖u · ∇Cnh‖

20 + (D∇Cnh ,∇Cnh )

− δ (∇ · (D∇Cnh ),u · ∇Cnh ) .(5.1.45)

93

Page 105: Modeling, Analysis, and Simulation of Adsorption in ...

As a(·, ·) is a bilinear form, we need only show positivity for the existence of a unique

solution. We obtain lower bounds for each term in a(·, ·) as follows:

(u · ∇Cnh , Cnh ) =

∫Ω

u · (∇Cnh )CnhdΩ

=1

2

∫Ω

u · ∇(Cnh2)dΩ

=1

2

∫Γ(Cnh )2 (u · n) ds−

∫Ω

(∇ · u)︸ ︷︷ ︸=0

(Cnh )2dΩ

=

1

2

∫ΓD

(Cnh )2︸ ︷︷ ︸=0

(u · n)ds+1

2

∫ΓN

(Cnh )2 (u · n)︸ ︷︷ ︸≥0

ds

≥ 0,

(D∇Cnh ,∇Cnh ) = (D1/2∇Cnh ,D1/2∇Cnh )

= ‖D1/2∇Cnh‖20

≥ λ‖∇Cnh‖20,

δ (∇ · (D∇Cnh ),u · ∇Cnh ) ≤(∇ ·D : ∇Cnh + D : ∇2Cnh , δu · ∇Cnh

)≤β2δ ‖∇Cnh‖0 ‖u · ∇C

nh‖0 + β1δ

∥∥∇2Cnh∥∥

0‖u · ∇Cnh‖0

≤β2δ‖u‖∞ ‖∇Cnh‖20 + β1δKih

−1 ‖∇Cnh‖0 ‖u · ∇Cnh‖0

≤β2δ‖u‖∞ ‖∇Cnh‖20 +

1

2ε‖∇Cnh‖

20 +

δ2β21K

2i h−2

2ε‖u · ∇Cnh‖

20

≤β2δ‖u‖∞ ‖∇Cnh‖20 +

λ

2‖∇Cnh‖

20 +

δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

⇒ −δ (∇ · (D∇Cnh ),u · ∇Cnh ) ≥− β2δ‖u‖∞ ‖∇Cnh‖20 −

λ

2‖∇Cnh‖

20 −

δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20.

94

Page 106: Modeling, Analysis, and Simulation of Adsorption in ...

Combining the bounds above with (5.1.45) gives

a(Cnh , Cnh ) ≥ ω

∆t‖Cnh‖

20 +

2− β2δ‖u‖∞

)‖∇Cnh‖

20

+ δ

(1− δβ2

1K2i h−2

)‖u · ∇Cnh‖

20.

(5.1.46)

To ensure positivity of the second and third terms in (5.1.46), we choose δ appro-

priately. For the first term on the left, we assume h ≤ 1 and choose δ = Kh such that

K <λ

2β2‖u‖∞.

Consequently

β2δ‖u‖∞ = β2Kh‖u‖∞ <λ

2h ≤ λ

2.

For the second term on the left, it seems that we must choose δ ∼ O(h2). However,

recall that δ is defined to be 0 when β1 > h so that the terms contributing to the second

component do not arise [50]. Consequently, we assume β1 ≤ h and use this assumption to

help control the h−2 in the second component. Then we choose δ = Kh such that

K <2λ

β1K2i

so that

δβ21K

2i h−2

2λ≤ δβ1K

2i h−1

2λ= K

β1K2i

2λ< 1.

Therefore, we choose δ = Kh such that

K < min

λ

2β2‖u‖∞,

β1K2i

to ensure positivity a(·, ·).

Thus Ker(a) = 0 and since (5.1.44) represents a square system of linear equation,

a unique solution exists.

95

Page 107: Modeling, Analysis, and Simulation of Adsorption in ...

We now state and prove the a priori stability bound.

Theorem 5.1.7. (Stability Bound) Suppose the assumptions of Lemma 5.1.6 are satisfied

so that the fully implicit SUPG formulation with constant adsorption given by (5.1.44) has a

solution Ch ∈ L2(0, T,H1(Ω)) with f ∈ L2(0, T ;L2(Ω). Then there exists a positive constant

K independent of h and ∆t such that for all N > 0

‖CNh ‖20 + ∆tN∑n=1

(‖u · ∇Cnh‖

20 + ‖∇Cnh‖

20

)≤ K

(‖C0

h‖20 + (1 + δ2)∆t

N∑n=1

‖fn‖20

).

Proof. To show the boundedness of the solution of (5.1.44), we choose vh = Cnh to obtain

(ω dt(Cnh ), Cnh ) + aδ(C

nh , C

nh ) = (fn, Cnh + δu · ∇Cnh ) (5.1.47)

The first term on the left is bounded below using

(x− y)x =1

2(x2 − y2 + (x− y)2) ≥ 1

2(x2 − y2),

to obtain

(ω dt(Cnh ), Cnh ) =

ω

∆t

(Cnh − C

n−1h , Cnh

)=

ω

2∆t

[(Cnh , C

nh )−

(Cn−1h , Cn−1

h

)+(Cnh − C

n−1h , Cnh − C

n−1h

)]=

ω

2∆t

[‖Cnh‖

20 − ‖Cn−1

h ‖20 + ‖Cnh − Cn−1h ‖20

]≥ ω

2∆t

[‖Cnh‖

20 − ‖Cn−1

h ‖20].

Lower bounds for the terms in aδ(Cnh , C

nh ) are obtained using the assumptions, Cauchy-

96

Page 108: Modeling, Analysis, and Simulation of Adsorption in ...

Schwarz Inequality, and Young’s Inequality as shown below.

(u · ∇Cnh , Cnh ) =

∫Ω

u · (∇Cnh )CnhdΩ

=1

2

∫Ω

u · ∇(Cnh2)dΩ

=1

2

∫Γ(Cnh )2 (u · n) ds−

∫Ω

(∇ · u)︸ ︷︷ ︸=0

(Cnh )2dΩ

=

1

2

∫ΓD

(Cnh )2︸ ︷︷ ︸=0

(u · n)ds+1

2

∫ΓN

(Cnh )2 (u · n)︸ ︷︷ ︸≥0

ds

≥ 0

(u · ∇Cnh , δu · ∇Cnh ) = δ‖u · ∇Cnh‖20

(D∇Cnh ,∇Cnh ) = (D1/2∇Cnh ,D1/2∇Cnh )

= ‖D1/2∇Cnh‖20

≥ λ‖∇Cnh‖20

97

Page 109: Modeling, Analysis, and Simulation of Adsorption in ...

δ (∇ · (D∇Cnh ),u · ∇Cnh ) ≤(∇ ·D : ∇Cnh + D : ∇2Cnh , δu · ∇Cnh

)≤β2δ ‖∇Cnh‖0 ‖u · ∇C

nh‖0 + β1δ

∥∥∇2Cnh∥∥

0‖u · ∇Cnh‖0

≤β2δKih−1 ‖Cnh‖0 ‖u · ∇C

nh‖0 + β1δKih

−1 ‖∇Cnh‖0 ‖u · ∇Cnh‖0

≤ β22

2ε1‖Cnh‖

20 +

δ2K2i h−2

2ε1‖u · ∇Cnh‖

20

+1

2ε2‖∇Cnh‖

20 +

δ2β21K

2i h−2

2ε2‖u · ∇Cnh‖

20

≤β22λ

2β21

‖Cnh‖20 +

δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

4‖∇Cnh‖

20 +

δ2β21K

2i h−2

λ‖u · ∇Cnh‖

20

⇒ −δ (∇ · (D∇Cnh ),u · ∇Cnh ) ≥− β22λ

2β21

‖Cnh‖20 −

δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

− λ

4‖∇Cnh‖

20 −

δ2β21K

2i h−2

λ‖u · ∇Cnh‖

20

An upper bound for the term involving f is obtained similarly:

(fn, Cnh + δu · ∇Cnh ) ≤‖fn‖0‖Cnh‖0 + δ‖fn‖0‖u · ∇Cnh‖0

≤1

2‖fn‖20 +

1

2‖Cnh‖

20 +

δ2

2ε‖fn‖20 +

1

2ε‖u · ∇Cnh‖

20

≤1

2‖fn‖20 +

1

2‖Cnh‖

20 +

δ2

2ε‖fn‖20 +

1

2ε‖u‖2∞‖∇Cnh‖

20

=1

2‖fn‖20 +

1

2‖Cnh‖

20 +

5δ2‖u‖2∞λ

‖fn‖20 +λ

4‖∇Cnh‖

20.

Combining the above bounds with (5.1.47) and collecting terms, we obtain

ω

2∆t

(‖Cnh‖

20 − ‖Cn−1

h ‖20)

2‖∇Cnh‖

20 + δ

(1− 3δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

≤ 1

2

(1 +

β22λ

β21

)‖Cnh‖

20 +

(1

2+

5δ2‖u‖2∞λ

)‖fn‖20.

98

Page 110: Modeling, Analysis, and Simulation of Adsorption in ...

Multiplying through by 2∆tω and summing from n = 1 to n = N gives us

‖CNh ‖20 + ∆tN∑n=1

2

ω

2‖∇Cnh‖

20 + δ

(1− 3δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

)≤ ∆t

ω

N∑n=0

(1 +

β22λ

β21

)‖Cnh‖

20 + ∆t

N∑n=1

2

ω

(1

2+

5δ2‖u‖2∞λ

)‖fn‖20 + ‖C0

h‖20.

(5.1.48)

To ensure positivity of the terms on the left, we choose δ appropriately. The choice of

δ = O(h2) seems necessary to control the last term on the left. However recall that δ is

defined to be 0 when β1 > h so that the terms contributing to the second component do

not arise [50]. Consequently, we assume β1 ≤ h and use this assumption to help control the

h2. Specifically, we choose δ = Kh such that

K <2λ

3β1K2i

,

then

3δβ21K

2i h−2

2λ≤ 3Kβ2

1K2i h−1

2λ≤ 3β1K

2i

2λK < 1.

Then applying Discrete Gronwall’s Inequality (Lemma 3.4.5) to (5.1.48) gives

‖CNh ‖20 + ∆t

N∑n=1

2

ω

2‖∇Cnh‖

20 + δ

(1− 3δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

)≤ K

(‖C0

h‖20 + ∆t

N∑n=1

2

ω

(1

2+

5δ2‖u‖2∞λ

)‖fn‖20

).

Since the coefficients on the left are positive and the coefficients on the right are finite, then

we have for any N > 0

⇒ ‖CNh ‖20 + ∆tN∑n=1

(‖u · ∇Cnh‖

20 + ‖∇Cnh‖

20

)≤ K

(‖C0

h‖20 + (1 + δ2)∆t

N∑n=1

‖fn‖20

)

where K is a positive constant independent of h and ∆t.

Next, we obtain an a priori error estimate in a similar fashion to the a priori stability

99

Page 111: Modeling, Analysis, and Simulation of Adsorption in ...

bound and the error estimates for the semi-continuous case. We use the same notation as

before.

Theorem 5.1.8. (Error Estimate) Suppose the assumptions of Lemma 5.1.6 are satisfied

so that the fully implicit SUPG formulation with constant adsorption given by (5.1.44) has a

solution Ch. Assume also that the variational formulation with constant adsorption given by

(5.1.1) has an exact solution C ∈ H2(0, T,Hk+1(Ω)). Then there exists a positive constant

K such that

‖C − Ch‖L2(0,T ;H1(Ω)) ≤ K

(hk(1 + δ)‖C‖H1(0,T ;Hk+1(Ω) + ∆t

∥∥∥∥∂2C

∂t2

∥∥∥∥L2(0,T ;L2(Ω))

)

Proof. Subtracting (5.1.44) from (5.1.1) gives

ω

(∂C

∂t− dt(Cnh ), v

)+ (u · ∇en, v) + δ (u · ∇en,u · ∇v) + (D∇en,∇v)

− δ (∇ · (D∇en),u · ∇v) = 0.

Consequently,

ω (dt(χn), v) + (u · ∇χn, v) + δ (u · ∇χn,u · ∇v) + (D∇(χn),∇v)

− δ (∇ · (D∇χn),u · ∇v) = ω (rn, v) + ω (dt(ηn), v) + (u · ∇ηn, v)

+ δ (u · ∇ηn,u · ∇v) + (D∇(ηn),∇v)− δ (∇ · (D∇ηn),u · ∇v)

(5.1.49)

where rn is the residual in the derivative approximation at time tn,

rn =∂C

∂t− dt(Cn),

and χn and ηn are defined as

ηn := Cn − Cn, χn := Cnh − Cn

where Cn is the elliptic projection of Cn defined in Lemma 3.4.8. Choosing v = χn, we use

100

Page 112: Modeling, Analysis, and Simulation of Adsorption in ...

the fact that

1

2(x2 − y2) ≤ 1

2(x2 − y2 + (x− y)2) = (x− y)x,

to bound the first term on the left as follows:

ω (dt(χn), χn) ≥ ω

2∆t

(‖χn‖20 − ‖χn−1‖20

).

Lower bounds for the rest of the terms on the left are bounded using the assumptions,

Cauchy-Scwarz Inequality, and Young’s Inequality in a similar manner to the proof of

stability and are shown below.

(u · ∇χn, χn) ≥ 0,

(u · ∇χn, δu · ∇χn) = δ‖u · ∇χn‖20,

(D∇χn,∇χn) ≥ λ‖∇χn‖20,

−δ (∇ · (D∇χn),u · ∇χn) ≥− β22λ

2β21

‖χn‖20 −δ2β2

1K2i h−2

2λ‖u · ∇χn‖20

− λ

8‖∇χn‖20 −

2δ2β21K

2i h−2

λ‖u · ∇χn‖20

We obtain upper bounds for the terms on the right using similar techniques as shown below.

ω (rn, χn) + ω (dt(ηn), χn) ≤ ω‖χn‖20 +

ω

2‖dt(ηn)‖20 +

ω

2‖rn‖20

(u · ∇ηn, χn) ≤ 1

2‖u · ∇ηn‖20 +

1

2‖χn‖20

101

Page 113: Modeling, Analysis, and Simulation of Adsorption in ...

(u · ∇ηn, δu · ∇χn) ≤ 2δ2‖u‖2∞λ

‖u · ∇ηn‖20 +λ

8‖∇χn‖20

−(∇ · (D∇ηn), δu · ∇χn) ≤2δ2β22‖u‖2∞λ

‖∇ηn‖20 +λ

8‖∇χn‖20

+2δ2β2

1‖u‖2∞λ

‖∇2ηn‖20 +λ

8‖∇χn‖20

Note by the definition of the elliptic projection (D∇ηn,∇χn) = 0. Using the bounds above

along with (5.1.49) and combining like terms gives

ω

2∆t

(‖χn‖20 − ‖χn−1‖20

)+λ

2‖∇χn‖20 + δ

(1− 5δβ2

1K2i h−2

)‖u · ∇χn‖20

≤(ω +

1

2+β2

2β21

)‖χn‖20 +

ω

2‖dt(ηn)‖20 +

ω

2‖rn‖20

+

(1

2+

2δ2‖u‖2∞λ

)‖u · ∇ηn‖20 +

2δ2‖u‖2∞λ

(β2

2‖∇ηn‖20 + β21‖∇2ηn‖20

)We bound the second and third terms on the right hand side of (5.1.3) in a similar

manner to the fully explicit formulation:

‖dt(ηn)‖20 ≤1

∆t

∫ tntn−1

∥∥∥∥∂η∂t∥∥∥∥2

0

dt, (5.1.50)

and

‖rn‖20 =∆t

3

∫ tntn−1

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt (5.1.51)

Combining (5.1.3) with (5.1.50) and (5.1.51), we obtain

ω

2∆t

(‖χn‖20 − ‖χn−1‖20

)+λ

2‖∇χn‖20 + δ

(1− 5δβ2

1K2i h−2

)‖u · ∇χn‖20

≤(ω +

1

2+β2

2β21

)‖χn‖20 +

ω

2

(1

∆t

∫ tntn−1

∥∥∥∥∂η∂t∥∥∥∥2

0

dt

)+ω

2

(∆t

3

∫ tntn−1

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

)

+

(1

2+

2δ2‖u‖2∞λ

)‖u · ∇ηn‖20 +

2δ2‖u‖2∞λ

(β2

2‖∇ηn‖20 + β21‖∇2ηn‖20

),

102

Page 114: Modeling, Analysis, and Simulation of Adsorption in ...

and multiplying by 2∆tω and summing from n = 1 to n = N gives

‖χN‖20 + ∆tN∑n=1

2

ω

2‖∇χn‖20 + δ

(1− 5δβ2

1K2i h−2

)‖u · ∇χn‖20

)≤ ∆t

N∑n=1

2

ω

(ω +

1

2+β2

2β21

)‖χn‖20 + ∆t

(1

∆t

∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt

)

+ ∆t

(∆t

3

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

)+ ∆t

N∑n=1

2

ω

(1

2+

2δ2‖u‖2∞λ

)‖u · ∇ηn‖20

+ ∆tN∑n=1

4δ2‖u‖2∞ωλ

(β2

2‖∇ηn‖20 + β21‖∇2ηn‖20

)+ ‖χ0‖20

(5.1.52)

where T = tN .

To ensure positivity of the terms on the left, we choose δ. As with the proof of

stability, the choice of δ = O(h2) seems necessary to control the last term on the left.

However recall that δ is defined to be 0 when β1 > h so that the terms contributing to

the second component do not arise [50]. Consequently, we assume β1 ≤ h and use this

assumption when choosing δ. Specifically, we choose δ = Kh such that

K <2λ

5β1K2i

,

then

5δβ21K

2i h−2

2λ≤ 5Kβ2

1K2i h−1

2λ≤ 5β1K

2i

2λK < 1.

Applying Discrete Gronwall’s Inequality (Lemma 3.4.5) to (5.1.52) and using the

103

Page 115: Modeling, Analysis, and Simulation of Adsorption in ...

fact that ‖χ0‖20 = 0, we obtain

‖χN‖20 + ∆tN∑n=1

2

ω

2‖∇χn‖20 + δ

(1− 5δβ2

1K2i h−2

)‖u · ∇χn‖20

)≤ K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+(∆t)2

3

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

+ ∆tN∑n=1

2

ω

(1

2+

2δ2‖u‖2∞λ

)‖u · ∇ηn‖20

+ ∆tN∑n=1

4δ2‖u‖2∞ωλ

(β2

2‖∇ηn‖20 + β21‖∇2ηn‖20

)).

(5.1.53)

Note that the second derivative term on the right in (5.1.53) can be bounded by its

gradient. This terms is associated with the upwinded diffusion term which is defined locally

over the interior of the mesh triangulation [34, 82]. Thus a local approximation property

can be applied as in [80], and the associated δ2 can serve to bound the h−2 appearing from

the application of the approximation property. Therefore, since the coefficients on the left

are positive and the coefficients on the right are finite, (5.1.53) implies

∆tN∑n=1

‖∇χn‖20 ≤ K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

+ ∆t(1 + δ2)

N∑n=1

(‖u · ∇ηn‖20 + ‖∇ηn‖20

)).

(5.1.54)

Since u is bounded,

(‖∇ηn‖20 + ‖u · ∇ηn‖20

)≤ (1 + ‖u‖2∞)‖∇ηn‖20

and (5.1.54) implies

∆tN∑n=1

‖∇χn‖20 ≤ K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

+ ∆t(1 + δ2)

N∑n=1

K‖∇ηn‖20

).

(5.1.55)

104

Page 116: Modeling, Analysis, and Simulation of Adsorption in ...

Applying Lemma 3.4.7 (the consequence of Poincare’s Inequality) to (5.1.55) gives

∆tN∑n=1

‖∇χn‖21 ≤ K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

+ ∆t(1 + δ2)

N∑n=1

K‖∇ηn‖20

).

(5.1.56)

By the Triangle Inequality

∆tN∑n=1

‖en‖1 =∆tN∑n=1

‖ηn − χn‖21

≤∆tN∑n=1

(‖ηn‖1 + ‖χn‖1)2

≤∆tN∑n=1

K(‖ηn‖21 + ‖χn‖21

)and consequently

∆t

N∑n=1

‖en‖21 ≤ K

(∆t

N∑n=1

‖ηn‖21 +

∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

+ ∆t(1 + δ2)

N∑n=1

‖∇ηn‖20

). (5.1.57)

Since C ∈ H2(0, T,Hk+1(Ω)), we can apply the projection inequality (Lemma 3.4.8) to

‖η‖0, ‖∇η‖0, and∥∥∥∂η∂t ∥∥∥0

as follows:

‖η‖20 ≤ K2ph

2k‖C‖2k, (5.1.58)

‖∇η‖20 ≤ K2ph

2k‖∇C‖2k, (5.1.59)∥∥∥∥∂η∂t∥∥∥∥2

0

≤ K2ph

2

∥∥∥∥∂C∂t∥∥∥∥2

1

. (5.1.60)

105

Page 117: Modeling, Analysis, and Simulation of Adsorption in ...

Combining (5.1.57)-(5.1.60) gives us

∆tN∑n=1

‖en‖21 ≤K

(h2k∆t

N∑n=1

‖Cn‖2k+1dt+ h2

∫ T

0

∥∥∥∥∂C∂t∥∥∥∥2

1

dt+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

+ h2k(1 + δ2)∆tN∑n=1

‖∇Cn‖2k

)

≤K

(h2

∫ T

0

∥∥∥∥∂C∂t∥∥∥∥2

1

+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

+ h2k(1 + δ2)

∫ T

0‖C‖2k+1

)

≤K

(h2k(1 + δ2)‖C‖2H1(0,T ;Hk+1(Ω) + (∆t)2

∥∥∥∥∂2C

∂t2

∥∥∥∥2

L2(0,T ;L2(Ω))

),

and consequently

‖C − Ch‖L2(0,T ;H1(Ω)) ≤ K

(h2k(1 + δ)‖C‖H1(0,T ;Hk+1(Ω) + ∆t

∥∥∥∥∂2C

∂t2

∥∥∥∥L2(0,T ;L2(Ω))

)

where K is a positive constant independent of h and ∆t.

5.2 Case 2: Linear Isotherm

We now state and prove a priori stability bounds and a priori error estimates for

the case of linear adsorption. In the case of a linear isotherm, q(C) = K1 + K2C with

K1,K2 > 0. Therefore, ∂q∂t = K2∂C∂t , and consequently the upwinded variational formulation

in (3.2.5)-(3.2.6) becomes the following: Find C ∈ H2(Ω) such that

((ω + (1− ω)ρsK2)

∂C

∂t, v

)+ (u · ∇C, v) + δ (u · ∇C,u · ∇v) + (D∇C,∇v)

− δ(∇ · (D∇C),u · ∇v) = (f, v) + δ (f,u · ∇v)

(5.2.1)

for all v ∈ V .

Since the time-derivative of the adsorption is now simply a constant multiple of the

time derivative of the concentration, the proofs below are very similar to the case of linear

adsorption. In fact, the only modification is the constant in front of the time derivative is

changed from simply the porosity, ω, to a quantity taking into account the porosity and the

106

Page 118: Modeling, Analysis, and Simulation of Adsorption in ...

adsorption constant, ω + (1− ω)ρsK2.

5.2.1 Continuous In Time Analysis

The upwinded semi-discrete formulation with linear adsorption is given by the fol-

lowing: Find Ch ∈ Vh such that

((ω + (1− ω)ρsK2)

∂Ch∂t

, vh

)+ (u · ∇Ch, vh) + δ (u · ∇Ch,u · ∇vh)

+ (D∇Ch,∇vh)− δ(∇ · (D∇Ch),u · ∇vh) = (f, vh) + δ (f,u · ∇vh)

(5.2.2)

for all vh ∈ Vh. The exact variational formulation is given by (5.2.1).

We begin with stating and proving an a priori stability bound.

Theorem 5.2.1. (Stability Bound) Assume that (B1) - (B4) are satisfied and that the

semi-discrete SUPG formulation with linear adsorption given by (5.2.2) has a solution

Ch ∈ L∞(0, T,H1(Ω)) with f ∈ L2(0, T ;L2(Ω)). Then there exists a positive constant

K independent of h such that

‖Ch‖2L∞(0,T ;L2(Ω)) +

∫ T

0

(‖u · ∇Ch‖20 + ‖∇Ch‖20

)dt ≤ K

[‖C0‖20 + (1 + δ2)‖f‖2L2(0,T ;L2(Ω))

].

Proof. As the only difference between this case and the constant isotherm case is the con-

stant in front of ∂C∂t , the stability proof carries through similarly. Taking v = Ch in the

semi-discrete formulation, bounding and rewriting the terms in a similar manner, we obtain

an equation similar to (5.1.4):

ω + (1− ω)ρsK2

2

d

dt‖Ch‖20 + δ

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇Ch‖20 +

λ

2‖∇Ch‖20

≤ 1

4

(1 +

β22λ

β21

)‖Ch‖20 +

(1 +

δ2‖u‖2∞λ

)‖f‖20. (5.2.3)

107

Page 119: Modeling, Analysis, and Simulation of Adsorption in ...

Multiplying (5.2.3) through by 2ω+(1−ω)ρsK2

and integrating from 0 to t we obtain

‖Ch(t)‖20 +2

ω + (1− ω)ρsK2

∫ t

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇Ch(s)‖20ds

+2

ω + (1− ω)ρsK2

∫ t

0

λ

4‖∇Ch(s)‖20ds

≤ 1

ω + (1− ω)ρsK2

∫ t

0

1

2

(1 +

β22λ

β21

)‖Ch(s)‖20ds+ ‖Ch(0)‖20

+2

ω + (1− ω)ρsK2

∫ t

0

(1 +

δ2‖u‖2∞λ

)‖f‖20ds.

(5.2.4)

To ensure positivity of the terms on the left, we have the same requirement on δ

as in the case of constant adsorption, namely choose δ = Kh such that K < λ2β1K2

i. Then

Continuous Gronwall’s Inequality (Lemma 3.4.4) applied to (5.2.4) implies

‖Ch(t)‖20 +2K

ω + (1− ω)ρsK2

∫ t

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇Ch(s)‖20ds

+2K

ω + (1− ω)ρsK2

∫ t

0

λ

4‖∇Ch(s)‖20ds

]≤ K

[‖Ch(0)‖20 +

1

ω + (1− ω)ρsK2

∫ t

0

(1 +

δ2‖u‖2∞λ

)‖f‖20ds

] (5.2.5)

where

K = max0≤s≤t

exp

(s

ω + (1− ω)ρsK2

)= exp

(t

ω + (1− ω)ρsK2

).

Since the coefficients on the left are positive, the coefficients on the right are finite, Ch(0) =

C0, and (5.2.5) is true for any 0 ≤ t ≤ T , then we have

‖Ch‖2L∞(0,T ;L2(Ω)) +

∫ T

0‖u · ∇Ch‖20 +

∫ T

0‖∇Ch‖20 ≤ K

[‖C0‖20 + (1 + δ2)‖f‖2L2(0,T ;L2(Ω))

]

where K is a positive constant independent of h.

We now state and prove an a priori error estimate for the case of linear adsorption.

Theorem 5.2.2. (Error Estimate) Suppose that (B1) - (B4) are satisfied. Assume also

that the variational formulation with linear adsorption given by (5.2.1) has an exact solution

108

Page 120: Modeling, Analysis, and Simulation of Adsorption in ...

C ∈ H1(0, T,Hk+1(Ω)), and Ch solves the semi-discrete SUPG formulation with linear

adsorption given by (5.2.2). Then there exists a positive constant K independent of h such

that

‖C − Ch‖L2(0,T ;H1(Ω)) ≤ K

(h

∥∥∥∥∂C∂t∥∥∥∥L2(0,T ;H1(Ω))

+ hk(1 + δ)‖C‖L2(0,T ;Hk+1(Ω))

)

Proof. As the only difference between this case and the constant isotherm case is the con-

stant in front of ∂C∂t , the error analysis carries through similarly. All steps completed for the

constant isotherm case can be repeated in this case, and this process leads to an equation

similar to (5.1.11):

‖χ(t)‖20 +

∫ T

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇χ(s)‖20ds

+2

ω + (1− ω)ρsK2

∫ T

0

λ

2‖∇χ(s)‖20ds

≤ 1

ω + (1− ω)ρsK2

∫ T

0

(ω + (1− ω)ρsK2 + 1 +

β22λ

β21

)‖χ(s)‖20ds

+1

ω + (1− ω)ρsK2

∫ T

0(ω + (1− ω)ρsK2)

∥∥∥∥∂η∂t (s)

∥∥∥∥2

0

ds

+1

ω + (1− ω)ρsK2

∫ T

0

(1 +

3δ2‖u‖2∞λ

)‖u · ∇η(s)‖20ds

+1

ω + (1− ω)ρsK2

∫ T

0

3δ2‖u‖2∞λ

[β2

2 ‖∇η(s)‖20 + β21

∥∥∇2η(s)∥∥2

0

]ds+ ‖χ(0)‖20.

(5.2.6)

Again, we ensure positivity of the coefficients on the left by choosing δ = Kh

appropriate; specifically we choose K < λ2β1K2

i. Then we use the fact that χ(0) = 0 and

109

Page 121: Modeling, Analysis, and Simulation of Adsorption in ...

apply continuous Gronwall’s Inequality (Lemma 3.4.4) to (5.2.6) to obtain

‖χ(t)‖20 +2K

ω + (1− ω)ρsK2

∫ T

(1− 2δβ2

1K2i h−2

λ

)‖u · ∇χ(s)‖20ds

+2K

ω + (1− ω)ρsK2

∫ T

0

λ

2‖∇χ(s)‖20ds

]

≤ K

ω + (1− ω)ρsK2

∫ T

0(ω + (1− ω)ρsK2)

∥∥∥∥∂η∂t (s)

∥∥∥∥2

0

ds

+K

ω + (1− ω)ρsK2

∫ T

0

(1 +

3δ2‖u‖2∞λ

)‖u · ∇η(s)‖20ds

+K

ω + (1− ω)ρsK2

∫ T

0

3δ2‖u‖2∞λ

[β2

2 ‖∇η(s)‖20 + β21

∥∥∇2η(s)∥∥2

0

]ds.

Consequently, we have

‖χ(t)‖20 +Km

∫ T

0

(‖u · ∇χ(s)‖20 + ‖∇χ(s)‖20

)ds

≤ K∫ T

0

∥∥∥∥∂η∂t (s)

∥∥∥∥2

0

ds+ (1 + δ2)KM

(∫ T

0

(‖u · ∇η(s)‖20 + ‖∇η(s)‖20 + ‖∇2η(s)‖20

)ds

).

Then using the same process as in the constant case, we obtain the following bound

‖C − Ch‖L2(0,T ;H1(Ω)) ≤K

(h

∥∥∥∥∂C∂t∥∥∥∥L2(0,T ;H1(Ω))

+ hk(1 + δ)‖C‖L2(0,T ;Hk+1(Ω))

)

where K is a positive constant independent of h.

5.2.2 Fully Explicit Analysis

As in the fully explicit formulation with constant adsorption, the theorems and

proofs below show the solvability is trivial but a strict constraint on the time step is required

to obtain the a priori stability bound and error estimate.

In the case of linear adsorption, the fully explicit, discrete formulation in (3.3.5)-

110

Page 122: Modeling, Analysis, and Simulation of Adsorption in ...

(3.3.6) simplifies to the following: For n = 1, ..., N , find Cnh ∈ Vh such that

((ω + ρs(1− ω)K2)dt(Cnh ), vh) +

(u · ∇Cn−1

h , vh))

+ δ(u · ∇Cn−1

h ,u · ∇vh)

+(D∇Cn−1

h ,∇vh)− δ

(∇ · (D∇Cn−1

h ),u · ∇vh)

=(fn−1, vh

)+ δ

(fn−1,u · ∇vh

) (5.2.7)

for all vh ∈ Vh. The exact variational formulation is again given by (5.2.1).

We begin by showing that (5.2.7) is uniquely solvable for Cnh at each time step n.

Lemma 5.2.3. (Solvability) Assume (B1) - (B4) are satisfied. Then there exist a unique

solution Cnh ∈ Vh satisfying (5.2.7).

Proof. As in the case of constant adsorption, we assume ω, u, and D are not dependent

on Cnh . In addition now, we assume ρs and K2 are also independent of Cnh . Consequently,

(5.2.7) is a fully explicit linear equation, and the existence and uniqueness of Cnh is trivial.

At each n, the system arising from (5.2.7) is a square diagonal system of linear equations

which implies the existence of a unique solution Cnh ∈ Vh.

We now state and prove an a priori stability bound for the case of linear adsorption.

Theorem 5.2.4. (Stability Bound) Suppose the assumption of Lemma 5.2.3 are satisfied

so that the fully explicit SUPG formulation with linear adsorption given by (5.2.7) has a

solution Ch ∈ L2(0, T,H1(Ω)) with f ∈ L2(0, T ;L2(Ω)). In addition assume that ∆t is on

the order of h2. Then, there exists a positive constant K independent of h and ∆t such that

for all N > 0

‖CNh ‖20 + ∆t

N∑n=1

(‖u · ∇Cnh‖

20 + ‖∇Cnh‖

20

)≤ K

(‖C0

h‖20 + (1 + δ2)∆tN∑n=1

‖fn−1‖20

).

Proof. To show the stability of (5.2.7), we will let vh = Cnh which gives

((ω + ρs(1− ω)K2)dt(Cnh ), Cnh ) + aδ(C

n−1h , Cnh ) =

(fn−1, Cnh + δu · ∇Cnh

)(5.2.8)

111

Page 123: Modeling, Analysis, and Simulation of Adsorption in ...

where aδ is the bilinear form defined in (4.2.1). We add and subtract aδ(Cnh , C

n−1h ), rewrite

and bound all all terms in a similar manner to the case of constant adsorption, and com-

bining like terms to obtain

ω + (1− ω)ρsK2

2∆t

(‖Cnh‖

20 − ‖Cn−1

h ‖20)

2‖∇Cnh‖

20 + δ

(1− 7δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

+

[ω + (1− ω)ρsK2

2∆t− 5β2

1 + λ(‖u‖2∞ + λ) + 5δ2‖u‖2∞(‖u‖2∞ + β22)

2λK2i h−2

]‖Cnh − C

n−1h ‖20

≤(

1 +β2

2β21

)‖Cnh‖

20 +

1

2

(1 +

5δ2‖u‖2∞λ

)‖fn−1‖20.

Multiplying through by 2∆tω+ρs(1−ω)K2

and summing from n = 1 to n = N gives us

‖CNh ‖20 + ∆tN∑n=1

2

ω + (1− ω)ρsK2

2‖∇Cnh‖

20 + δ

(1− 7δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

)+ ∆t

N∑n=1

[1

∆t− 5β2

1 + λ(‖u‖2∞ + λ) + 5δ2‖u‖2∞(‖u‖2∞ + β22)

(ω + (1− ω)ρsK2)λK2i h−2

]‖Cnh − C

n−1h ‖20

≤ ∆t

N∑n=1

2

ω + (1− ω)ρsK2

(1 +

β22λ

2β21

)‖Cnh‖

20 + ‖C0

h‖20 (5.2.9)

+ ∆t

N∑n=1

1

ω + (1− ω)ρsK2

(1 +

5δ2‖u‖2∞λ

)‖fn−1‖20.

To ensure positivity of the terms on the left, we assume conditions on δ = Kh and

∆t similar to the case of constant adsorption, namely

∆t ≤ (ω + (1− ω)ρsK2)λ(5β2

1 + λ(‖u‖2∞ + λ) + 5δ2‖u‖2∞(‖u‖2∞ + β22))K2i

h2

and

K <2λ

7β1K2i

.

We therefore apply Discrete Gronwall’s Inequality (Lemma 3.4.5) to (5.2.9) to obtain

112

Page 124: Modeling, Analysis, and Simulation of Adsorption in ...

‖CNh ‖20 + ∆tN∑n=1

2

ω + (1− ω)ρsK2

2‖∇Cnh‖

20 + δ

(1− 7δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

)+ ∆t

N∑n=1

[1

∆t− 5β2

1 + λ(‖u‖2∞ + λ) + 5δ2‖u‖2∞(‖u‖2∞ + β22)

(ω + (1− ω)ρsK2)λK2i h−2

]‖Cnh − C

n−1h ‖20

≤ K

(‖C0

h‖20 + ∆t

N∑n=1

1

ω + (1− ω)ρsK2

(1 +

5δ2‖u‖2∞λ

)‖fn−1‖20

).

Continuing as in the case of constant adsorption, we obtain for any N > 0

‖CNh ‖20 + ∆tN∑n=1

(‖u · ∇Cnh‖

20 + ‖∇Cnh‖

20

)≤ K

(‖C0

h‖20 + (1 + δ2)∆tN∑n=1

‖fn−1‖20

)

where K is a positive constant independent of h and ∆t.

Next, we obtain and an a priori error estimate in a similar fashion as the a pri-

ori stability bound and the a priori error estimate for the constant adsorption case. We

maintain the same notation.

Theorem 5.2.5. (Error Estimate) Suppose the assumption of Lemma 5.2.3 are satisfied

so that the fully explicit SUPG formulation with linear adsorption given by (5.2.7) has a

solution Ch. Assume also that the variational formulation with linear adsorption given by

(5.2.1) has an exact solution C ∈ H2(0, T,Hk+1(Ω)). In addition assume that ∆t is on the

order of h2. Then there exists a positive constant K independent of h and ∆t such that

‖C − Ch‖L2(0,T ;H1(Ω)) ≤ K

(hk(1 + δ) ‖C‖H1(0,T ;Hk+1(Ω)) + ∆t

∥∥∥∥∂2C

∂t2

∥∥∥∥L2(0,T ;L2(Ω))

).

Proof. Subtracting (5.2.7) from (5.2.1) we obtain

(ω + ρs(1− ω)K2)

(∂C

∂t− dt(Cnh ), vh

)+(u · ∇en−1, vh

)+ δ

(u · ∇en−1,u · ∇vh

)+(D∇en−1,∇vh

)− δ

(∇ · (D∇en−1),u · ∇vh

)= 0

113

Page 125: Modeling, Analysis, and Simulation of Adsorption in ...

where en = Cn − Cnh . Consequently,

(ω + ρs(1− ω)K2) (dt(χn), vh) + aδ(χ

n−1, vh)

= (ω + ρs(1− ω)K2) (rn, vh) + (ω + ρs(1− ω)K2) (dt(ηn), vh) + aδ(η

n−1, vh)(5.2.10)

where rn is the residual in the derivative approximation at time tn

rn =∂C

∂t− dt(Cn)

and χn and ηn−1 are defined as

ηn := Cn − Cn, χn := Cnh − Cn

with Cn the elliptic projection of Cn defined in Lemma 3.4.8.

Now choosing v = χn and bounding the terms similarly to the case of constant

adsorption, we obtain

ω + ρs(1− ω)K2

2∆t

(‖χn‖20 − ‖χn−1‖20

)+ δ

(1− 9δ2β2

1K2i h−2

)‖u · ∇χn‖20

2‖∇χn‖20 +

ω + ρs(1− ω)K2

2∆t‖χn − χn−1‖20

−(

7β21 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β2

2)

)K2i h−2‖χn − χn−1‖20

≤(ω + ρs(1− ω)K2 + 1 +

β22λ

2β21

)‖χn‖20 +

ω + ρs(1− ω)K2

2‖dt(ηn)‖20

+ω + ρs(1− ω)K2

2‖rn‖20 +

1

2

(1 +

7δ2‖u‖2∞λ

)‖u · ∇ηn−1‖20

+7δ2β2

2‖u‖2∞2λ

‖∇ηn−1‖20 +7δ2β2

1‖u‖2∞2λ

‖∇2ηn−1‖20.

(5.2.11)

We bound the second and third terms on the right hand side of (5.2.11) the same as with

114

Page 126: Modeling, Analysis, and Simulation of Adsorption in ...

the case of constant adsorption so (5.2.11) becomes

ω + ρs(1− ω)K2

2∆t

(‖χn‖20 − ‖χn−1‖20

)+ δ

(1− 9δ2β2

1K2i h−2

)‖u · ∇χn‖20

2‖∇χn‖20 +

ω + ρs(1− ω)K2

2∆t‖χn − χn−1‖20

−(

7β21 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β2

2)

)K2i h−2‖χn − χn−1‖20

≤(ω + ρs(1− ω)K2 + 1 +

β22λ

2β21

)‖χn‖20 +

ω + ρs(1− ω)K2

2

(1

∆t

∫ tntn−1

‖ηt‖20

)

+ω + ρs(1− ω)K2

2

(∆t

3

∫ tntn−1

‖Ctt‖20

)+

1

2

(1 +

7δ2‖u‖2∞λ

)‖u · ∇ηn−1‖20

+7δ2β2

2‖u‖2∞2λ

‖∇ηn−1‖20 +7δ2β2

1‖u‖2∞2λ

‖∇2ηn−1‖20.

Then multiplying through by 2∆tω+ρs(1−ω)K2

and summing (5.1.35) from n = 1 to n = N gives

‖χN‖20 + ∆t

N∑n=1

2

ω + ρs(1− ω)K2

(1− 9δ2β2

1K2i h−2

)‖u · ∇χn‖20 +

λ

2‖∇χn‖20

]+ ∆t

N∑n=1

[1

∆t−(

7β21 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β2

2)

ω + ρs(1− ω)K2λ

)K2i h−2

]‖χn − χn−1‖20

≤ ∆tN∑n=1

2

ω + ρs(1− ω)K2

(ω + ρs(1− ω)K2 + 1 +

β22λ

2β21

)‖χn‖20 + ‖χ0‖20

+ ∆t

(1

∆t

∫ T

0‖ηt‖20

)+ ∆t

(∆t

3

∫ T

0‖Ctt‖20

)+ ∆t

N∑n=1

1

ω + ρs(1− ω)K2

(1 +

7δ2‖u‖2∞λ

)‖u · ∇ηn−1‖20

+ ∆tN∑n=1

7δ2β22‖u‖2∞

(ω + ρs(1− ω)K2)λ‖∇ηn−1‖20 + ∆t

N∑n=1

7δ2β21‖u‖2∞

(ω + ρs(1− ω)K2)λ‖∇2ηn−1‖20.

To ensure positivity of the terms on the left, we have similar constraints on δ = Kh

and ∆t as in the case of constant adsorption, namely

K <2λ

9β1K2i

,

115

Page 127: Modeling, Analysis, and Simulation of Adsorption in ...

and

∆t ≤ ω + ρs(1− ω)K2λ(7β2

1 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β22))K2i

h2.

Then applying Discrete Gronwall’s Inequality (Lemma 3.4.5) and using the fact that

‖χ0‖0 = 0, we see

‖χN‖20 + ∆t

N∑n=1

2

ω + ρs(1− ω)K2

(1− 9δ2β2

1K2i h−2

)‖u · ∇χn‖20 +

λ

2‖∇χn‖20

]+ ∆t

N∑n=1

[1

∆t−(

7β21 + λ(‖u‖2∞ + λ) + 7δ2‖u‖2∞(‖u‖2∞ + β2

2)

ω + ρs(1− ω)K2λ

)K2i h−2

]‖χn − χn−1‖20

≤ K

[∆t

(1

∆t

∫ T

0‖ηt‖20

)+ ∆t

(∆t

3

∫ T

0‖Ctt‖20

)

+ ∆t

N∑n=1

1

ω + ρs(1− ω)K2

(1 +

7δ2‖u‖2∞λ

)‖u · ∇ηn−1‖20

+ ∆t

N∑n=1

7δ2‖u‖2∞(ω + ρs(1− ω)K2)λ

(β2

2‖∇ηn−1‖20 + β21‖∇2ηn−1‖20

).

Continuing in a similar manner as the case of constant adsorption, we have

∆tN∑n=1

‖en‖21 ≤ K

(h2k(1 + δ2) ‖C‖2H1(0,T ;Hk+1(Ω)) + (∆t)2

∥∥∥∥∂2C

∂t2

∥∥∥∥2

L2(0,T ;L2(Ω))

),

and consequently

‖C − Ch‖L2(0,T ;H1(Ω)) ≤ K

(hk(1 + δ)‖C‖H1(0,T ;Hk+1(Ω) + ∆t

∥∥∥∥∂2C

∂t2

∥∥∥∥L2(0,T ;L2(Ω))

)

where K is a positive constant independent of h and ∆t.

5.2.3 Fully Implicit Analysis

As with the case of constant adsorption, the fully-implicit formulation is solvable

and has no restrictions on the time step for either the stability bound or error estimate

which we show below.

With linear adsorption, the fully implicit, discrete formulation in (3.3.7)-(3.3.8)

116

Page 128: Modeling, Analysis, and Simulation of Adsorption in ...

simplifies to the following: For n = 0, 1, ..., N − 1, find Cnh ∈ Vh such that

((ω + ρs(1− ω)K2) dt(C

nh ), vh

)+ (u · ∇Cnh , vh) + δ (u · ∇Cnh ,u · ∇vh)

+ (D∇Cnh ,∇vh)− δ (∇ · (D∇Cnh ),u · ∇vh) = (fn, vh) + δ (fn,u · ∇vh)(5.2.12)

for all vh ∈ Vh. The exact variational formulation is again given by (5.2.1).

We begin by showing (5.2.12) is uniquely solvable for Cnh at each time step n.

Lemma 5.2.6. (Solvability) Assume (B1) - (B4) are satisfied. Then there exists a unique

solution Cnh ∈ Vh satisfying (5.2.12).

Proof. Choosing vh = Cnh in (5.2.12) and rearranging, we see

a(Cnh , Cnh ) =

(fn, Cn−1

h

)+ δ

(fn, Cn−1

h

)+ω + (1− ω)ρsK2

∆t

(Cn−1h , Cnh

)where

a(Cnh , Cnh ) =

ω + (1− ω)ρsK2

∆t‖Cnh‖

20 + (u · ∇Cnh , Cnh ) + δ‖u · ∇Cnh‖

20

+ (D∇Cnh ,∇Cnh )− δ (∇ · (D∇Cnh ),u · ∇Cnh ) .

Note that the bilinear form a(·, ·) is modified from the constant adsorption case only by the

coefficient of the first term, ω+(1−ω)ρsK2

∆t . By assumption ω + (1− ω)ρsK2 > 0, and conse-

quently we have positivity of a(·, ·) under the same requirements of the constant adsorption

case, namely that we choose δ = Kh such that

K < min

λ

2β2‖u‖∞,

β1K2i

.

Therefore, a(·, ·) is positive. Thus Ker(a) = 0 and since (5.2.12) represents a square

system of linear equation, a unique solution exists.

We next state and prove an a priori stability bound for the case of linear adsorption.

Theorem 5.2.7. (Stability Bound) Suppose the assumptions of Lemma 5.2.6 are satisfied

117

Page 129: Modeling, Analysis, and Simulation of Adsorption in ...

so that the fully implicit SUPG formulation with linear adsorption given in (5.2.12) has

a solution Ch ∈ L2(0, T,H1(Ω)) with f ∈ L2(0, T ;L2(Ω)). Then there exists a positive

constant K independent of h and ∆t such that for all N > 0

‖CNh ‖20 + ∆tN∑n=1

(‖u · ∇Cnh‖

20 + ‖Cnh‖

20

)≤ K

(‖C0

h‖20 + ∆t(1 + δ2)

N∑n=1

‖fn‖20

).

Proof. To show the boundedness of the solution of (5.2.12), we choose vh = Cnh to obtain

((ω + ρs(1− ω)K2) dt(C

nh ), Cnh

)+ aδ(C

nh , C

nh ) = (fn, Cnh + δ,u · ∇Cnh ) . (5.2.13)

Bounding the terms in (5.2.13) in a similar manner to the case of constant adsorption, we

obtain

ω + ρs(1− ω)K2

2∆t

(‖Cnh‖

20 − ‖Cn−1

h ‖20)

2‖∇Cnh‖

20 + δ

(1− 3δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

≤ 1

2

(1 +

β22λ

β21

)‖Cnh‖

20 +

(1

2+

5δ2‖u‖2∞λ

)‖fn‖20.

Multiplying through by 2∆tω+ρs(1−ω)K2

and summing from n = 1 to n = N then gives us

‖CNh ‖20 + ∆tN∑n=1

2

ω + ρs(1− ω)K2

λ

2‖∇Cnh‖

20

+ ∆tN∑n=1

2

ω + ρs(1− ω)K2δ

(1− 3δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

≤ ∆t

ω + ρs(1− ω)K2

N∑n=0

(1 +

β22λ

β21

)‖Cnh‖

20 + ‖C0

h‖20

+ ∆tN∑n=1

2

ω + ρs(1− ω)K2

(1

2+

5δ2‖u‖2∞λ

)‖fn‖20.

(5.2.14)

To ensure positivity of the terms on the left, we assume conditions on δ = Kh

similar to the case of constant adsorption, namely

K <2λ

3β1K2i

.

118

Page 130: Modeling, Analysis, and Simulation of Adsorption in ...

Then Discrete Gronwall’s Inequality (Lemma 3.4.7) applied to (5.2.14) implies

‖CNh ‖20 + ∆tN∑n=1

2

ω + ρs(1− ω)K2

2‖∇Cnh‖

20 + δ

(1− 3δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

)≤ K

(‖C0

h‖20 + ∆t

N∑n=1

2

ω + ρs(1− ω)K2

(1

2+

5δ2‖u‖2∞λ

)‖fn‖20

).

Consequently for any N > 0

‖CNh ‖20 + ∆tN∑n=1

(‖u · ∇Cnh‖

20 + ‖∇Cnh‖

20

)≤ K

(‖C0

h‖20 + ∆t(1 + δ2)

N∑n=1

‖fn‖20

)

where K is a positive constant independent of h and ∆t.

Next, we obtain an a priori error estimate in a similar fashion to the a priori stability

bound and the error estimates for the case of constant adsorption. We maintain the same

notation.

Theorem 5.2.8. (Error Estimate) Suppose the assumptions of Lemma 5.2.6 are satisfied

so that the fully implicit SUPG formulation with linear adsorption given in (5.2.12) has a

solution Ch. Assume also that the variational formulation with linear adsorption given by

(5.2.1) has an exact solution C ∈ H2(0, T,Hk+1(Ω)). Then there exists a positive constant

K such that

‖C − Ch‖L2(0,T ;H1(Ω)) ≤ K

(hk(1 + δ)‖C‖H1(0,T ;Hk+1(Ω) + ∆t

∥∥∥∥∂2C

∂t2

∥∥∥∥L2(0,T ;L2(Ω))

).

Proof. Subtracting (5.2.12) from (5.2.1) gives

(ω + (1− ω)ρsK2)

(∂C

∂t− dt(Cnh ), v

)+ (u · ∇en, v) + δ (u · ∇en,u · ∇v)

+ (D∇en,∇v)− δ (∇ · (D∇en),u · ∇v) = 0.

119

Page 131: Modeling, Analysis, and Simulation of Adsorption in ...

Consequently,

(ω + ρs(1− ω)K2) (dt(χn), v) + (u · ∇χn, v)

+ δ (u · ∇χn,u · ∇v) + (D∇(χn),∇v)− δ (∇ · (D∇χn),u · ∇v)

= (ω + ρs(1− ω)K2) (rn, v) + (ω + ρs(1− ω)K2) (dt(ηn), v) + (u · ∇ηn, v)

+ δ (u · ∇ηn,u · ∇v) + (D∇(ηn),∇v)− δ (∇ · (D∇ηn),u · ∇v)

(5.2.15)

where rn is the residual in the derivative approximation at time tn+1

rn =∂C

∂t− dt(Cn)

and χn and ηn are defined as

ηn := Cn − Cn, χn := Cnh − Cn

where Cn is the elliptic projection of Cn defined in Lemma 3.4.8. Bounding the terms in

(5.2.15) in a similar manner to the error bound in the case of constant adsorption, we obtain

ω + (1− ω)ρsK2

2∆t

(‖χn‖20 − ‖χn−1‖20

)+λ

2‖∇χn‖20 + δ

(1− 5δβ2

1K2i h−2

)‖u · ∇χn‖20

≤(ω + (1− ω)ρsK2 +

1

2+β2

2β21

)‖χn‖20 +

ω + (1− ω)ρsK2

2‖dt(ηn)‖20

+ω + (1− ω)ρsK2

2‖rn‖20 +

(1

2+

2δ2‖u‖2∞λ

)‖u · ∇ηn‖20

+2δ2‖u‖2∞

λ

(β2

2‖∇ηn‖20 + β21‖∇2ηn‖20

).

Bounding the second and third terms on the right hand side of (5.2.3) in a similar manner

120

Page 132: Modeling, Analysis, and Simulation of Adsorption in ...

to the the case of constant adsorption, we obtain

ω + (1− ω)ρsK2

2∆t

(‖χn‖20 − ‖χn−1‖20

)+λ

2‖∇χn‖20 + δ

(1− 5δβ2

1K2i h−2

)‖u · ∇χn‖20

≤(ω + (1− ω)ρsK2 +

1

2+β2

2β21

)‖χn‖20 +

ω + (1− ω)ρsK2

2

(1

∆t

∫ tntn−1

∥∥∥∥∂η∂t∥∥∥∥2

0

dt

)

+ω + (1− ω)ρsK2

2

(∆t

3

∫ tntn−1

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

)+

(1

2+

2δ2‖u‖2∞λ

)‖u · ∇ηn‖20

+2δ2‖u‖2∞

λ

(β2

2‖∇ηn‖20 + β21‖∇2ηn‖20

),

and multiplying by 2∆tω and summing from n = 1 to n = N gives

‖χN‖20 + ∆tN∑n=1

2

ω + (1− ω)ρsK2

2‖∇χn‖20 + δ

(1− 5δβ2

1K2i h−2

)‖u · ∇χn‖20

)≤ ∆t

N∑n=1

2

ω + (1− ω)ρsK2

(ω + (1− ω)ρsK2 +

1

2+β2

2β21

)‖χn‖20

+ ∆t

(1

∆t

∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt

)+ ∆t

(∆t

3

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

)

+ ∆tN∑n=1

2

ω + (1− ω)ρsK2

(1

2+

2δ2‖u‖2∞λ

)‖u · ∇ηn‖20

+ ∆tN∑n=1

4δ2‖u‖2∞(ω + (1− ω)ρsK2)λ

(β2

2‖∇ηn‖20 + β21‖∇2ηn‖20

)+ ‖χ0‖20

where T = tN .

To ensure positivity of the terms on the left, we have similar constraints on δ = Kh

as in the case of constant adsorption, namely

K <2λ

5β1K2i

.

Then applying Discrete Gronwall’s Inequality (Lemma 3.4.5) and using the fact that ‖χ0‖0 =

121

Page 133: Modeling, Analysis, and Simulation of Adsorption in ...

0, we see

‖χN‖20 + ∆tN∑n=1

2

ω + (1− ω)ρsK2

2‖∇χn‖20 + δ

(1− 5δβ2

1K2i h−2

)‖u · ∇χn‖20

)≤ K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+(∆t)2

3

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

+ ∆tN∑n=1

2

ω + (1− ω)ρsK2

(1

2+

2δ2‖u‖2∞λ

)‖u · ∇ηn‖20

+ ∆tN∑n=1

4δ2‖u‖2∞(ω + (1− ω)ρsK2)λ

(β2

2‖∇ηn‖20 + β21‖∇2ηn‖20

).

As in the case of constant adsorption, we obtain

∆t

N∑n=1

‖∇χn‖20 ≤ K

(∫ T

0

∥∥∥∥∂η∂t∥∥∥∥2

0

dt+ (∆t)2

∫ T

0

∥∥∥∥∂2C

∂t2

∥∥∥∥2

0

dt

+ ∆t(1 + δ2)N∑n=1

(‖u · ∇ηn‖20 + ‖∇ηn‖20

)).

(5.2.16)

Then by the Triangle Inequality and the Projection Inequality, (5.2.16) gives us

∆t

N∑n=1

‖en‖20 ≤ K

(h2k(1 + δ2)‖C‖2H1(0,T ;Hk+1(Ω) + (∆t)2

∥∥∥∥∂2C

∂t2

∥∥∥∥2

L2(0,T ;L2(Ω))

),

and consequently

‖C − Ch‖L2(0,T ;H1(Ω)) ≤ K

(hk(1 + δ)‖C‖H1(0,T ;Hk+1(Ω) + ∆t

∥∥∥∥∂2C

∂t2

∥∥∥∥L2(0,T ;L2(Ω))

)

where K is a positive constant independent of h and ∆t.

122

Page 134: Modeling, Analysis, and Simulation of Adsorption in ...

Chapter 6

Analysis for Nonlinear, Explicit

Adsorption

For the case of the nonlinear isotherm with an explicit representation (as in Lang-

muir’s isotherm), we consider a few different discretization schemes. First, we apply an

idea introduced by Nochetto in [121] and applied in several mixed formulations [6, 171, 147]

to formulate a time-integrated version of the transport equation. Based on this technique,

we develop two finite element formulations: one using mixed finite elements and one using

SUPG. We also consider a formulation developed from the original fully implicit SUPG

discretization scheme in (3.3.7)-(3.3.8). Using the explicit representation of the isotherm,

the two-equation scheme can be reduced to one-equation and analyzed as such. The time-

integrated discretization schemes are developed fully below in sections 6.1.1 and 6.2.1 while

the discretization scheme based on the original discretization (3.3.7)-(3.3.8) is derived in

section 6.3.1.

6.1 Time-Integrated Mixed Method Formulation

As with the original description of the problem in Section 2.1, we let Ω be a bounded

domain in Rd, d = 1, 2, or 3 with a Lipschitz continuous boundary ∂Ω = Γ. Also let Γ+,

123

Page 135: Modeling, Analysis, and Simulation of Adsorption in ...

Γ−, and Γn be disjoint subsets of Γ such that Γ = Γ+ ∪ Γ− ∪ Γn (where Γ+,Γ−,Γn are

defined in detail in section 2.1.6) and let [0, T ] be a finite time interval with T denoting the

final time. We still consider solving (2.2.1)-(2.2.7). For the mixed method discretization

schemes, we assume assume the following:

(C1) ω and ρs are constant in time and space. [50]

(C2) u is nonzero, independent of time [34] and essentially bounded in space [147, 34] with

∇ · u = 0 [34].

(C3) D = [dij ]i,j=1...n is symmetric positive definite [4, 50], independent of time (since

u is assumed independent of time), and essentially bounded in space [4, 147, 34];

specifically, we notate the boundedness of D using ‖D‖∞ ≤ β1.

(C4) C is nonnegative [50, 9, 48] and is at least in L2(0, T ;H1 (Ω)

)[147], and C = 0 on

ΓD. ,

(C5) q = q(C) ∈ C1 is an explicit [147, 50], Lipschitz continuous [147, 134] function of C,

q(0) = 0 [50, 9, 10], q(C) > 0 for C > 0 [9, 10], and q is nondecreasing [147, 9, 10, 48,

132, 137] .

As noted previously, the assumption that C = 0 on ΓD is not always true in practice.

However, the assumption is still reasonable as a transformation can be applied to C in any

case when a nonzero Dirichlet boundary condition arises to give a solution that is zero on

ΓD.

6.1.1 Development of Time-Integrated Mixed Finite Element Formula-

tion

To develop the time-integrated, mixed method finite element discretization, we begin

by letting

Q = uC −D∇C. (6.1.1)

124

Page 136: Modeling, Analysis, and Simulation of Adsorption in ...

Then (2.2.1) can be rewritten as the set of coupled equations

(ωC + (1− ω)ρsq)t +∇ ·Q = f, (6.1.2)

Q− uC + D∇C = 0, (6.1.3)

and we consider solving equations (6.1.2) and (6.1.3) over [0, T ]×Ω with the following initial

and boundary conditions:

C = Cin on Γ+, (6.1.4)

Q · n = 0 on Γ− ∪ Γn, (6.1.5)

C(0,x) = C0 in Ω, (6.1.6)

q(C(0,x)) = q(C0) = q0 in Ω. (6.1.7)

Since D is positive definite, then D−1 exists so we can rewrite (6.1.3) as

D−1Q−D−1uC +∇C = 0. (6.1.8)

Also, since C ∈ L2(0, T ;H1(Ω) and q is Lipschitz continuous with q(0) = 0, then

‖q(C)‖20 = ‖q(C)− q(0)‖20

=

∫Ω|q(C)− q(0)|2dΩ

≤∫

ΩK|C − 0|dΩ

= K‖C‖20

125

Page 137: Modeling, Analysis, and Simulation of Adsorption in ...

⇒ ‖q‖L2(0,T ;L2(Ω)) =

∫ T

0‖q‖20dt

≤ K∫ T

0‖C‖20dt

= K‖C‖L2(0,T ;L2(Ω)

≤ K‖C‖L2(0,T ;H1(Ω) <∞

and consequently we have q ∈ L2(0, T ;L2(Ω)). Therefore we will rewrite (6.1.2) by inte-

grating in time to obtain

ωC + (1− ω)ρsq +∇ ·∫ t

0Q dτ =

∫ t

0f dτ + ωC0. (6.1.9)

We define the spaces Hdiv(Ω) and Hdiv0,N (Ω) as follows:

Hdiv(Ω) := z | z ∈ (L2(Ω))d,∇ · z ∈ L2(Ω),

Hdiv0,N (Ω) := z | z ∈ Hdiv(Ω), 〈z · n, v〉ΓN

= 0.

Multiplying (6.1.9) and (6.1.8) by v ∈ L2(Ω) and z ∈ Hdiv0,N (Ω) respectively and

integrating over Ω, we obtain

(ωC, v) + ((1− ω)ρs, q, v) +

(∇ ·∫ t

0Q dτ, v

)=

(∫ t

0f dτ, v

)+ (ωC0, v) ,

(D−1Q, z

)−(D−1uC, z

)+ (∇C, z) = 0.

By Green’s theorem

(∇C, z) = 〈C, z · n〉 − (C,∇ · z) = − (C,∇ · z)

where the boundary terms disappears since we have z ∈ Hdiv0,N (Ω) and C = 0 on ΓD by

assumption.

Consequently, we have the following mixed method variational formulation: Find

126

Page 138: Modeling, Analysis, and Simulation of Adsorption in ...

(C,Q) ∈ L2(Ω)×Hdiv0,N (Ω) such that

(ωC, v) + ((1− ω)ρs q, v) +

(∇ ·∫ t

0Q dτ, v

)=

(∫ t

0f dτ, v

)+ (ωC0, v) (6.1.10)

(D−1Q, z

)−(D−1uC, z

)− (C,∇ · z) = 0 (6.1.11)

for all v ∈ V = L2(Ω), z ∈ Z = Hdiv0,N (Ω).

As with the original SUPG formulation, we let Th = E be a triangulation of Ω so

that

Ω =⋃E∈Th

E.

We seek approximate solutions in discrete subspaces of V and Z, Vh ⊂ V = L2(Ω) and

Zh ⊂ Z = Hdiv0,N (Ω), where Vh and Zh are formed using standard mixed finite element

spaces such as the Raviart-Thomas spaces of order k, RTk.

In the error analysis, we will make use of the standard L2 projection Ph : L2(Ω)→ Vh

defined by

((Phv − v

), vh

)= 0 ∀vh ∈ Vh (6.1.12)

and also a weighted (L2)d projection Ph :(L2(Ω)

)d → Zh given by

(D−1 (Phz− z) , zh

)= 0 ∀zh ∈ Zh. (6.1.13)

In addition, we define an Hdiv projection operator Πh : Hdiv(Ω)→ Vh as in [6, 16, 147, 171]

so that for z ∈ H1(Ω):

(∇ · (Πhz− z) , vh) = 0 ∀vh ∈ Vh (6.1.14)

Then for v ∈ Hk+1(Ω) and z ∈ (Hk+1(Ω))d, we have the following inequalities pertaining

127

Page 139: Modeling, Analysis, and Simulation of Adsorption in ...

to the projection operators defined in (6.1.12)-(6.1.14):

‖v − Phv‖0 ≤ Khs‖v‖s,

‖z− Phz‖0 ≤ Khs‖z‖s,

‖z−Πhz‖0 ≤ Khs‖z‖s,

for some 0 ≤ s ≤ k+1 where K is a generic positive constant independent of the discretiza-

tion parameters. Note that in the error analysis, we wish to apply (6.1.14) to∫ t

0 Qdτ which

can only be done if this function is sufficiently smooth. Therefore we explicitly make the

necessary regularity assumption as in [6]:

(C6)

∫ t

0Qdτ ∈ (H1(0, T ;H1(Ω)))d.

Then the semi-discrete variational formulation is given by the following: Find

(Ch,Qh) ∈ Vh × Zh such that

(ωCh, vh) + ((1− ω)ρs q(Ch), vh) +

(∇ ·∫ t

0Qh dτ, vh

)=

(∫ t

0f dτ, vh

)+ (ωC0, vh) (6.1.15)

(D−1Qh, zh

)−(D−1uCh, zh

)− (Ch,∇ · zh) = 0 (6.1.16)

for all vh ∈ Vh, zh ∈ Zh.

To formulate the fully discrete variational formulation, we will partition the time

interval [0, T ] as

t0 = 0 < t1 < · · · < tN = T

and let ∆t = tn+1− tn denote the step size for t where tn = n∆t. Also we will use the same

notation for Cn and Cnh and the following notation for Qn and Qnh:

Qn denotes the exact solutions at time tn, i.e. Qn = Q(tn).

Qnh denotes the approximate solutions at time tn, i.e. Qn

h = Qh(tn),

128

Page 140: Modeling, Analysis, and Simulation of Adsorption in ...

Then the fully discrete variational problem can be written as follows: For n = 1, ..., N , find

(Cnh ,Qnh) ∈ Vh × Zh such that

(ωCnh , vh) + ((1− ω)ρs q(Cnh ), vh) +

∇ · n∑j=1

Qjh ∆t, vh

=

n∑j=1

f j ∆t, vh

+ (ωC0, vh) (6.1.17)

(D−1Qn

h, zh)−(D−1uCnh , zh

)− (Cnh ,∇ · zh) = 0 (6.1.18)

for all vh ∈ Vh, zh ∈ Zh.

6.1.2 Error Estimates

We include a priori error estimates for both the semi-continuous and fully discrete

formulations.

6.1.2.1 Semi-continuous Error Estimates

We first state and prove a priori error estimates for the semi-continuous mixed

method formulation given by (6.1.15)-(6.1.16).

Theorem 6.1.1. Suppose that (C1)-(C6) are satisfied and assume the mixed method vari-

ational formulation (6.1.10)-(6.1.11) has an exact solution (C,Q) ∈ L2(0, T ;Hs(Ω)) ×

L2(0, T ;H1(Ω)) for some 0 ≤ s ≤ k+ 1 and (Ch,Qh) solves the semidiscrete mixed method

formulation (6.1.15)-(6.1.16). Then for each T > 0 there exists a constant K > 0 indepen-

dent of h such that

∥∥∥∥∫ T

0(C − Ch) dt

∥∥∥∥2

0

+

∥∥∥∥∫ T

0(Q−Qh) dt

∥∥∥∥2

0

≤ K(h2s‖C‖2s + h2‖Q‖21

).

129

Page 141: Modeling, Analysis, and Simulation of Adsorption in ...

Proof. Subtracting (6.1.15) and (6.1.16) from (6.1.10) and (6.1.11) respectively, we obtain

(ω (C − Ch) , vh) + ((1− ω)ρs (q(C)− q(Ch)) , vh)

+

(∇ ·∫ t

0(Q−Qh), dτ, vh

)= 0,

(6.1.19)

(D−1 (Q−Qh) , zh

)−(D−1u (C − Ch) , zh

)− (C − Ch,∇ · zh) = 0.

(6.1.20)

Then using the projections defined in (6.1.12)-(6.1.14), we can rewrite (6.1.19) and (6.1.20)

as

(ω(C − Ch), vh) + ((1− ω)ρs (q(C)− q(Ch)) , vh) +

(∇ ·Πh

∫ t

0(Q−Qh), dτ, vh

)= 0,

(D−1 (PhQ−Qh) , zh

)−(D−1u (C − Ch) , zh

)−(PhC − Ch,∇ · zh

)= 0.

Choosing vh = PhC − Ch and zh = Πh

∫ t0 (Q−Qh) dτ we obtain

(ω(C − Ch), PhC − Ch

)+(

(1− ω)ρs (q(C)− q(Ch)) , PhC − Ch)

+

(∇ ·Πh

∫ t

0(Q−Qh), dτ, PhC − Ch

)= 0,

(6.1.21)

(D−1 (PhQ−Qh) ,Πh

∫ t

0(Q−Qh) dτ

)−(

D−1u (C − Ch) ,Πh

∫ t

0(Q−Qh) dτ

)−(PhC − Ch,∇ ·Πh

∫ t

0(Q−Qh) dτ

)= 0.

(6.1.22)

Summing (6.1.21) and (6.1.22) gives us

(ω(C − Ch), PhC − Ch

)+(

(1− ω)ρs (q(C)− q(Ch)) , PhC − Ch)

+

(D−1 (PhQ−Qh) ,Πh

∫ t

0(Q−Qh) dτ

)−(

D−1u (C − Ch) ,Πh

∫ t

0(Q−Qh) dτ

)= 0,

130

Page 142: Modeling, Analysis, and Simulation of Adsorption in ...

and adding and subtracting C and ΠhQ where appropriate we obtain

ω‖C − Ch‖20 + ((1− ω)ρs (q(C)− q(Ch)) , C − Ch)

+

(D−1 (ΠhQ−Qh) ,Πh

∫ t

0(Q−Qh) dτ

)−(

D−1u (C − Ch) ,Πh

∫ t

0(Q−Qh) dτ

)=(ω(C − Ch), C − PhC

)+(

(1− ω)ρs (q(C)− q(Ch)) , C − PhC)

+

(D−1 (ΠhQ− PhQh) ,Πh

∫ t

0(Q−Qh) dτ

).

(6.1.23)

Now notice

(D−1 (ΠhQ−Qh) ,Πh

∫ t

0(Q−Qh) dτ

)

=

(D−1/2Πh (Q−Qh) ,D−1/2Πh

∫ t

0(Q−Qh)dτ

)=

1

2

d

dt

(D−1/2Πh

∫ t

0((Q−Qh) dτ,D−1/2Πh

∫ t

0(Q−Qh) dτ

)=

1

2

d

dt

∥∥∥∥D−1/2Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥2

0

,

so integrating in time from 0 to T gives us

∫ T

0

(D−1 (ΠhQ−Qh) ,Πh

∫ t

0(Q−Qh) dτ

)dt

=

∫ T

0

(1

2

d

dt

∥∥∥∥D−1/2Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥2

0

)dt

=1

2

∥∥∥∥D−1/2Πh

∫ T

0(Q−Qh) dτ

∥∥∥∥2

0

.

where D−1/2 is the unique positive definite matrix such that D−1/2D−1/2 = D−1. Therefore,

131

Page 143: Modeling, Analysis, and Simulation of Adsorption in ...

if we integrate (6.1.23) in time from 0 to T we obtain

∫ T

0ω ‖C − Ch‖20 dt+

∫ T

0((1− ω) ρs (q(C)− q(Ch)) , C − Ch) dt

+1

2

∥∥∥∥D−1/2Πh

∫ T

0(Q−Qh) dτ

∥∥∥∥2

0

−∫ T

0

(D−1u (C − Ch) ,Πh

∫ t

0(Q−Qh) dτ

)dt

=

∫ T

0

(ω(C − Ch), C − PhC

)dt+

∫ T

0

((1− ω)ρs (q(C)− q(Ch)) , C − PhC

)dt

+

∫ T

0

(D−1 (ΠhQ− PhQh) ,Πh

∫ t

0(Q−Qh) dτ

)dt.

(6.1.24)

The last two terms on the left are bounded using the assumptions, Cauchy-Schwarz

Inequality, and Young’s Inequality as shown below.

As D is essentially bounded, then D−1 is essentially bounded and consequently,

∫ T

0

(D−1u (C − Ch) ,Πh

∫ t

0(Q−Qh) dτ

)dt

≤∫ T

0

∥∥D−1u (C − Ch)∥∥

0

∥∥∥∥Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥0

dt

≤∫ T

0

(1

2ε1∥∥D−1u (C − Ch)

∥∥2

0+

1

2ε1

∥∥∥∥Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥2

0

)dt

≤∫ T

0

(K

2ε1∥∥D−1

∥∥2

∞ ‖u‖2∞ ‖C − Ch‖

20 +

1

2ε1

∥∥∥∥Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥2

0

)dt

⇒ −∫ T

0

(D−1u (C − Ch) ,Πh

∫ t

0(Q−Qh) dτ

)dt

≥ −∫ T

0

K

2ε1∥∥D−1

∥∥2

∞ ‖u‖2∞ ‖C − Ch‖

20 dt

−∫ T

0

1

2ε1

∥∥∥∥Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥2

0

dt.

Also, D−1/2 is positive definite so that

1

2

∥∥∥∥D−1/2Πh

∫ T

0(Q−Qh) dτ

∥∥∥∥2

0

≥ 1

∥∥∥∥Πh

∫ T

0(Q−Qh) dτ

∥∥∥∥2

0

.

132

Page 144: Modeling, Analysis, and Simulation of Adsorption in ...

The terms on the right are bounded in a similar manner and are shown below.

∫ T

0

(ω(C − Ch), C − PhC

)dt ≤

∫ T

0‖ω(C − Ch)‖0

∥∥∥C − PhC∥∥∥0dt

≤ 1

2ε2ω

∫ T

0‖C − Ch‖20 dt+

ω

2ε2

∫ T

0

∥∥∥C − PhC∥∥∥2

0dt

∫ T

0

((1− ω)ρs (q(C)− q(Ch)) ,C − PhC

)dt

≤∫ T

0(1− ω)ρs ‖(q(C)− q(Ch))‖0

∥∥∥C − PhC∥∥∥0dt

≤(1− ω)ρs2

ε3

∫ T

0‖q(C)− q(Ch)‖20 dt

+(1− ω)ρs

2ε3

∫ T

0

∥∥∥C − PhC∥∥∥2

0dt

∫ T

0

(D−1 (ΠhQ− PhQh) ,Πh

∫ t

0(Q−Qh) dτ

)dt

≤∫ T

0

∥∥D−1 (Πh − Ph) Q∥∥

0

∥∥∥∥Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥0

dt

≤K2

∥∥D−1∥∥2

∫ T

0‖(Πh − Ph) Q‖20 dt

+1

2

∫ T

0

∥∥∥∥Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥2

0

dt

133

Page 145: Modeling, Analysis, and Simulation of Adsorption in ...

Combining (6.1.24) with the bounds above and collecting like terms, we obtain

(ω − 1

2ε2ω −

K

2ε1∥∥D−1

∥∥2

∞ ‖u‖2∞

)∫ T

0‖C − Ch‖20 dt

+

∫ T

0((1− ω) ρs (q(C)− q(Ch)) , C − Ch) dt+

λ

2

∥∥∥∥Πh

∫ T

0(Q−Qh) dτ

∥∥∥∥2

0

≤ 1

2

ε2+

(1− ω)ρsε3

)∫ T

0

∥∥∥C − PhC∥∥∥2

0dt+

K

2

∥∥D−1∥∥2

∫ T

0‖(Πh − Ph) Q‖20 dt

+1

2ε3 (1− ω) ρs

∫ T

0‖q(C)− q(Ch)‖20 dt

+1

2

(1 +

1

ε1

)∫ T

0

∥∥∥∥Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥2

0

dt.

(6.1.25)

Since q is nondecreasing and Lipschitz continuous, then

∫ T

0‖q(C)− q(Ch)‖20 dt =

∫ T

0(q(C)− q(Ch), q(C)− q(Ch)) dt

=

∫ T

0(|q(C)− q(Ch)| , |q(C)− q(Ch)|) dt

≤∫ T

0(|q(C)− q(Ch)| ,K |C − Ch|) dt

= K

∫ T

0(q(C)− q(Ch), C − Ch) dt

Therefore

1

2ε3 (1− ω) ρs

∫ T

0‖q(C)− q(Ch)‖20 dt ≤

K

2ε3 (1− ω) ρs

∫ T

0(q(C)− q(Ch), C − Ch) dt

134

Page 146: Modeling, Analysis, and Simulation of Adsorption in ...

which can then be combined with the similar term on the left of (6.1.25) to obtain

(ω − 1

2ε2ω −

K

2ε1∥∥D−1

∥∥2

∞ ‖u‖2∞

)∫ T

0‖C − Ch‖20 dt

+

(1− K

2ε3

)(1− ω) ρs

∫ T

0((q(C)− q(Ch)) , C − Ch) dt

2

∥∥∥∥Πh

∫ T

0(Q−Qh) dτ

∥∥∥∥2

0

≤ 1

2

(1

ε2+

1

ε3

)∫ T

0

∥∥∥C − PhC∥∥∥2

0dt+

K

2

∥∥D−1∥∥2

∫ T

0‖(Πh − Ph) Q‖20 dt

+1

2

(1 +

1

ε1

)∫ T

0

∥∥∥∥Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥2

0

dt.

(6.1.26)

To simplify the coefficients on the left, we choose εi appropriately; specifically, we

let

ε1 =ω

2K∥∥D−1

∥∥2

∞ ‖u‖2∞, ε2 =

1

2, ε3 =

1

K, and ε4 = 1

then (6.1.26) simplifies to

1

∫ T

0‖C − Ch‖20 dt+

λ

2

∥∥∥∥Πh

∫ T

0(Q−Qh) dτ

∥∥∥∥2

0

+1

2(1− ω) ρs

∫ T

0(q(C)− q(Ch), C − Ch) dt

≤ 2 +K

2

∫ T

0

∥∥∥C − PhC∥∥∥2

0dt+

K

2

∥∥D−1∥∥2

∫ T

0‖(Πh − Ph) Q‖20 dt

+1

2

(2K

∥∥D−1∥∥2

∞ ‖u‖2∞

ω+ 1

)∫ T

0

∥∥∥∥Πh

∫ t

0(Q−Qh) dτ

∥∥∥∥2

0

dt.

(6.1.27)

Then Continuous Gronwall’s Inequality applied to (6.1.27) gives us

∫ T

0‖C − Ch‖20 dt+

∫ T

0(q(C)− q(Ch), C − Ch) dt+

∥∥∥∥Πh

∫ T

0(Q−Qh) dτ

∥∥∥∥2

0

≤ K(∫ T

0

∥∥∥C − PhC∥∥∥2

0dt+

∫ T

0‖(Πh − Ph) Q‖20 dt

). (6.1.28)

Since q is nondecreasing, then∫ T

0 (q(C)− q(Ch), C − Ch) dt > 0 and consequently, (6.1.28)

135

Page 147: Modeling, Analysis, and Simulation of Adsorption in ...

implies

∫ T

0‖C − Ch‖20 dt+

∥∥∥∥Πh

∫ T

0(Q−Qh) dτ

∥∥∥∥2

0

≤ K(∫ T

0

∥∥∥C − PhC∥∥∥2

0dt+

∫ T

0‖(Πh − Ph) Q‖20 dt

)(6.1.29)

We wish to bound∫ T

0 ‖C − Ch‖20 dt below by

∥∥∥∫ T0 Ph (C − Ch) dt∥∥∥2

0. To do so we

consider the fact that∫ T

0 (C − Ch)dt ∈ L2 so Ph∫ T

0 (C − Ch)dt ∈ Vh. Then

(Ph∫ T

0(C − Ch)dt−

∫ T

0(C − Ch)dt, vh

)= 0

for all vh ∈ Vh. We choose vh = Ph∫ T

0 (C − Ch)dt to obtain

(Ph∫ T

0(C − Ch)dt−

∫ T

0(C − Ch)dt, Ph

∫ T

0(C − Ch)dt

)= 0

⇔∥∥∥∥Ph ∫ T

0(C − Ch)dt

∥∥∥∥2

0

=

(∫ T

0(C − Ch)dt, Ph

∫ T

0(C − Ch)dt

).

Bounding the right side by using the Cauchy-Schwarz Inequality, we see

∥∥∥∥Ph ∫ T

0(C − Ch)dt

∥∥∥∥2

0

≤∥∥∥∥∫ T

0(C − Ch)dt

∥∥∥∥0

∥∥∥∥Ph ∫ T

0(C − Ch)dt

∥∥∥∥0

,

and consequently, ∥∥∥∥Ph ∫ T

0(C − Ch)dt

∥∥∥∥0

≤∥∥∥∥∫ T

0(C − Ch)dt

∥∥∥∥0

⇒∥∥∥∥Ph ∫ T

0(C − Ch)dt

∥∥∥∥2

0

≤∥∥∥∥∫ T

0(C − Ch)dt

∥∥∥∥2

0

≤ K∫ T

0‖C − Ch‖2dt (6.1.30)

where K is dependent on T but independent of h.

136

Page 148: Modeling, Analysis, and Simulation of Adsorption in ...

Combining (6.1.30) with (6.1.29) gives

∥∥∥∥Ph ∫ T

0(C − Ch) dτ

∥∥∥∥2

0

+

∥∥∥∥Πh

∫ T

0(Q−Qh) dτ

∥∥∥∥2

0

≤K(∫ T

0

∥∥∥C − PhC∥∥∥2

0dt+

∫ T

0‖(Πh − Ph) Q‖20 dt

),

≤K(∫ T

0

∥∥∥C − PhC∥∥∥2

0dt+

∫ T

0‖Q− PhQ‖20 dt

+

∫ T

0‖Q−ΠhQ‖20 dt

)

By adding and subtracting appropriate terms and applying the triangle inequality

∥∥∥∥∫ T

0(C − Ch)dt

∥∥∥∥2

0

≤ K

(∫ T

0

∥∥∥C − PhC∥∥∥2

0dt+

∥∥∥∥Ph ∫ T

0(C − Ch)dt

∥∥∥∥2

0

)

and

∥∥∥∥∫ T

0(Q−Qh)dt

∥∥∥∥2

0

≤ K

(∫ T

0

∥∥∥Q− PhQ∥∥∥2

0dt+

∥∥∥∥Ph ∫ T

0(Q−Qh)dt

∥∥∥∥2

0

).

Therefore, we have

∥∥∥∥∫ T

0(C − Ch)dt

∥∥∥∥2

0

+

∥∥∥∥∫ T

0(Q−Qh)dt

∥∥∥∥2

0

≤K(∫ T

0

∥∥∥C − PhC∥∥∥2

0dt+

∫ T

0‖Q− PhQ‖20 dt

+

∫ T

0‖Q−ΠhQ‖20 dt

)

and by applying the inequalities associated with the three projections, we obtain

∥∥∥∥∫ T

0(C − Ch) dt

∥∥∥∥2

0

+

∥∥∥∥∫ T

0(Q−Qh) dt

∥∥∥∥2

0

≤ K(h2s‖C‖2s + h2‖Q‖21

)where K is independent of h, but depends on T .

137

Page 149: Modeling, Analysis, and Simulation of Adsorption in ...

6.1.2.2 Fully Discrete Error Estimate

We next state and prove a priori error estimates for the fully discrete mixed method

formulation given by (6.1.17) and (6.1.18). We use the following notation for the mixed

method analysis:

Qn

=1

∆t

∫ tn

tn−1

Qdτ, φn = Qn −Qn

h, φn

=n∑j=1

φj∆t.

Note then that we have

∫ tn

0Qdτ =

n∑j=1

Qj∆t and φ

n=

∫ tn

0Qdτ −

n∑j=1

Qjh∆t.

Theorem 6.1.2. Suppose that (C2)-(C6) are satisfied and assume the mixed method vari-

ational formulation (6.1.10)-(6.1.11) has an exact solution (C,Q) ∈ H1(0, T ;Hs(Ω)) ×

H1(0, T ;H1(Ω)) for some 0 ≤ s ≤ k + 1, and (Cnh ,Qnh) solves the fully discrete mixed

method formulation (6.1.17)-(6.1.18) for each 1 ≤ n ≤ N . Then for each T > 0 there exists

a constant K > 0 independent of h and ∆t such that

∥∥∥∥∥∫ T

0Cdτ −

N∑n=1

Cnh∆t

∥∥∥∥∥2

0

+

∥∥∥∥∥∫ T

0Qdτ −

N∑n=1

Qnh∆t

∥∥∥∥∥2

0

≤ K((

∆th2s + (∆t)2)‖C‖2H1(0,T ;Hs(Ω)) + (∆th2 + (∆t)2) ‖Q‖2H1(0,T ;H1(Ω))

+ (∆t)2 ‖ΠhQt‖2L2(0,T ;L2(Ω))

)

Proof. Subtracting the fully discrete formulation (6.1.17)-(6.1.18) from the continuous for-

138

Page 150: Modeling, Analysis, and Simulation of Adsorption in ...

mulation (6.1.10)-(6.1.11) at time n, we obtain

(ω(Cn − Cnh ), vh) + ((1− ω) ρs (q (Cn)− q (Cnh )) , vh)

+

(∇ ·

∫ tn

0Qdτ −

n∑j=1

Qjh∆t

, vh

)= 0

(D−1 (Qn −Qn

h) , zh)− (D−1u (Cn − Cnh ) , zh)− (Cn − Cnh ,∇ · zh) = 0.

Rewriting these equations using the projections and notation defined previously, we have

(ω(Cn − Cnh ), vh) + ((1− ω) ρs (q (Cn)− q (Cnh )) , vh) +(∇ ·(

Πhφn), vh

)= 0, (6.1.31)(

D−1 (PhQn −Qnh) , zh

)− (D−1u (Cn − Cnh ) , zh)−

(PhCn − Cnh ,∇ · zh

)= 0. (6.1.32)

We choose vh = PhCn − Cnh and zh = Πhφn

so (6.1.31) and (6.1.32) become

(ω(Cn − Cnh ), PhCn − Cnh

)+(

(1− ω) ρs (q (Cn)− q (Cnh )) , PhCn − Cnh)

+(∇ ·(

Πhφn), PhCn − Cnh

)= 0,

(6.1.33)

(D−1 (PhQn −Qn

h) ,Πhφn)−(D−1u (Cn − Cnh ) ,Πhφ

n)

−(PhCn − Cnh ,∇ ·

(Πhφ

n))

= 0(6.1.34)

Summing (6.1.33) and (6.1.34) gives

(ω(Cn − Cnh ), PhCn − Cnh

)+(

(1− ω) ρs (q (Cn)− q (Cnh )) , PhCn − Cnh)

+(D−1 (PhQn −Qn

h) ,Πhφn)−(D−1u (Cn − Cnh ) ,Πhφ

n)

= 0

139

Page 151: Modeling, Analysis, and Simulation of Adsorption in ...

and adding and subtracting Cn and ΠhQn where appropriate, we obtain

ω ‖Cn − Cnh‖20 + ((1− ω) ρs (q (Cn)− q (Cnh )) , Cn − Cnh )

+(D−1 (ΠhQ

n −Qnh) ,Πhφ

n)−(D−1u (Cn − Cnh ) ,Πhφ

n)

=(ω(Cn − Cnh ), Cn − PhCn

)+(D−1 (ΠhQ

n − PhQn) ,Πhφn)

+(

(1− ω) ρs (q (Cn)− q (Cnh )) , Cn − PhCn).

(6.1.35)

Note that Qnh ∈ Zh ⊂ Hdiv(Ω) so we apply the projection Πh to Qn

h and use the fact that

ΠhQnh = Qn

h to obtain

ΠhQn −Qn

h = Πh (Qn −Qnh)

= Πh

(Qn −Qn

h + Qn − Q

n)= Πh

(φn −

(Qn −Qn

)).

Inserting this expression into (6.1.35), we obtain

ω ‖Cn − Cnh‖20 + ((1− ω) ρs (q (Cn)− q (Cnh )) , Cn − Cnh )

+(D−1Πhφ

n,Πhφn)−(D−1u (Cn − Cnh ) ,Πhφ

n)

=(ω(Cn − Cnh ), Cn − PhCn

)+(D−1 (Πh − Ph) Qn,Πhφ

n)

+(D−1

(Πh

(Qn −Qn

)),Πhφ

n)

+(

(1− ω) ρs (q (Cn)− q (Cnh )) , Cn − PhCn).

(6.1.36)

Multiplying (6.1.36) by the time step size ∆t and summing over times steps n = 1 to N

140

Page 152: Modeling, Analysis, and Simulation of Adsorption in ...

gives us

N∑n=1

∆tω ‖Cn − Cnh‖20 +

N∑n=1

∆t ((1− ω) ρs (q (Cn)− q (Cnh )) , Cn − Cnh )

+

N∑n=1

∆t(D−1Πhφ

n,Πhφn)−

N∑n=1

∆t(D−1u (Cn − Cnh ) ,Πhφ

n)

=

N∑n=1

∆t

[(ω(Cn − Cnh ), Cn − PhCn

)+(D−1 (Πh − Ph) Qn,Πhφ

n)

+(D−1

(Πh

(Qn −Qn

)),Πhφ

n)

+(

(1− ω) ρs (q (Cn)− q (Cnh )) , Cn − PhCn)].

(6.1.37)

We first consider the third term on the left-hand side on the equality in (6.1.37):

N∑n=1

∆t(D−1Πhφ

n,Πhφn)

=N∑n=1

(D−1/2Πhφ

n∆t,D−1/2Πhφn)

=N∑n=1

(D−1/2Πh

(φn − φ

n−1),D−1/2Πhφ

n)

=N∑n=1

[(D−1/2Πhφ

n,D−1/2Πhφ

n)−(D−1/2Πhφ

n−1,D−1/2Πhφ

n)]

=N∑n=1

[1

2

(D−1/2Πhφ

n,D−1/2Πhφ

n)− 1

2

(D−1/2Πhφ

n−1,D−1/2Πhφ

n−1)

+1

2

(D−1/2Πhφ

n,D−1/2Πhφ

n)−(D−1/2Πhφ

n−1,D−1/2Πhφ

n)

+1

2

(D−1/2Πhφ

n−1,D−1/2Πhφ

n−1)]

=

N∑n=1

[1

2

∥∥∥D−1/2Πhφn∥∥∥2

0− 1

2

∥∥∥D−1/2Πhφn−1∥∥∥2

0

]

+1

2

N∑n=1

∥∥∥D−1/2Πh

(φn − φ

n−1)∥∥∥2

0

=1

2

∥∥∥D−1/2ΠhφN∥∥∥2

0+

1

2

N∑n=1

∥∥∥D−1/2Πhφn∆t

∥∥∥2

0

≥1

2λ∥∥∥Πhφ

N∥∥∥2

0+

1

N∑n=1

(∆t)2 ‖Πhφn‖20

141

Page 153: Modeling, Analysis, and Simulation of Adsorption in ...

where λ > 0. We proceed by bounding the rest of the terms in a similar manner to the

semi-continuous error analysis as follows:

−(D−1u (Cn − Cnh ) ,Πhφ

n)≥ −K

2ε1∥∥D−1

∥∥2

∞ ‖u‖2∞ ‖Cn − Cnh‖

20 −

1

2ε1

∥∥∥Πhφn∥∥∥2

0

(ω(Cn − Cnh ), Cn − PhCn

)≤ 1

2ε2ω ‖Cn − Cnh‖

20 +

ω

2ε2

∥∥∥Cn − PhCn∥∥∥2

0

((1− ω) ρs (q (Cn)− q (Cnh )) , Cn − PhCn

)≤ 1

2ε3 ‖(1− ω)ρs (q (Cn)− q (Cnh ))‖20 +

(1− ω)ρs2ε3

∥∥∥Cn − PhCn∥∥∥2

0

≤ K

2ε3 (1− ω) ρs ((q (Cn)− q (Cnh )) , Cn − Cnh ) +

(1− ω)ρs2ε3

∥∥∥Cn − PhCn∥∥∥2

0

(D−1 (Πh − Ph) Qn,Πhφ

n)≤ K

2

∥∥D−1∥∥2

∞ ‖(Πh − Ph) Qn‖20 +1

2

∥∥∥Πhφn∥∥∥2

0

(D−1

(Πh

(Qn −Qn

)),Πhφ

n)≤ K

2

∥∥D−1∥∥2

∞∥∥Πh

(Qn −Qn

)∥∥2

0+

1

2

∥∥∥Πhφn∥∥∥2

0

Inserting the above bounds into (6.1.37) and collecting terms, we obtain

N∑n=1

∆t

(ω − 1

2ε2ω −

K

2ε1∥∥D−1

∥∥2

∞ ‖u‖2∞

)‖Cn − Cnh‖

20

+

N∑n=1

∆t (1− ω) ρs

(1− K

2ε3

)(q (Cn)− q (Cnh ) , Cn − Cnh )

+1

N∑n=1

∆t∥∥∥Πhφ

N∥∥∥2

0+

1

N∑n=1

(∆t)2 ‖Πhφn‖20

≤N∑n=1

∆t

[1

2

(1

ε2+

1

ε3

)∥∥∥Cn − PhCn∥∥∥2

0+K

2

∥∥D−1∥∥2

∞∥∥Πh

(Qn −Qn

)∥∥2

0

+K

2

∥∥D−1∥∥2

∞ ‖(Πh − Ph) Qn‖20 +1

2

(1

ε1+ 2

)∥∥∥Πhφn∥∥∥2

0

].

(6.1.38)

142

Page 154: Modeling, Analysis, and Simulation of Adsorption in ...

To simplify the coefficients on the left, we choose

ε1 =ω

2K∥∥D−1

∥∥2

∞ ‖u‖2∞, ε2 =

1

2, and ε3 =

1

K.

Then (6.1.38) simplifies to

1

N∑n=1

∆t ‖Cn − Cnh‖20 +

1

2(1− ω) ρs

N∑n=1

∆t (q (Cn)− q (Cnh ) , Cn − Cnh )

+1

2λ∥∥∥Πhφ

N∥∥∥2

0+

1

N∑n=1

(∆t)2 ‖Πhφn‖20

≤N∑n=1

∆t

[2 +K

2

∥∥∥Cn − PhCn∥∥∥2

0+K

2

∥∥D−1∥∥2

∞∥∥Πh

(Qn −Qn

)∥∥2

0

+K

2

∥∥D−1∥∥2

∞ ‖(Πh − Ph) Qn‖20 +

(K∥∥D−1

∥∥2

∞ ‖u‖2∞

ω+ 1

)∥∥∥Πhφn∥∥∥2

0

]

which implies

N∑n=1

∆t ‖Cn − Cnh‖20 +

N∑n=1

∆t (q (Cn)− q (Cnh ) , Cn − Cnh )

+∥∥∥Πhφ

N∥∥∥2

0+

N∑n=1

(∆t)2 ‖Πhφn‖20

≤ KN∑n=1

∆t

[ ∥∥∥Πhφn∥∥∥2

0+∥∥Πh

(Qn −Qn

)∥∥2

0

+ ‖(Πh − Ph) Qn‖20 +∥∥∥Cn − PhCn∥∥∥2

0

].

(6.1.39)

Then Discrete Gronwall’s Inequality applied to (6.1.39) implies

N∑n=1

∆t ‖Cn − Cnh‖20 +

N∑n=1

∆t (q (Cn)− q (Cnh ) , Cn − Cnh )

+∥∥∥Πhφ

N∥∥∥2

0+

N∑n=1

(∆t)2 ‖Πhφn‖20

≤ KN∑n=1

∆t

[ ∥∥Πh

(Qn −Qn

)∥∥2

0+ ‖(Πh − Ph) Qn‖20 +

∥∥∥Cn − PhCn∥∥∥2

0

].

(6.1.40)

143

Page 155: Modeling, Analysis, and Simulation of Adsorption in ...

Since q is nondecreasing then (q (Cn)− q (Cnh ) , Cn − Cnh ) > 0 so (6.1.40) implies

N∑n=1

∆t ‖Cn − Cnh‖20 +

∥∥∥ΠhφN∥∥∥2

0+

N∑n=1

(∆t)2 ‖Πhφn‖20

≤ KN∑n=1

∆t

[ ∥∥Πh

(Qn −Qn

)∥∥2

0+ ‖(Πh − Ph) Qn‖20 +

∥∥∥Cn − PhCn∥∥∥2

0

]

and rewriting the above using our definitions of φN

and φn we see

N∑n=1

∆t ‖Cn − Cnh‖20 +

∥∥∥∥∥∥Πh

∫ T

0Qdτ −

N∑j=1

Qjh∆t

∥∥∥∥∥∥2

0

+N∑n=1

(∆t)2∥∥ΠhQ

n −Qnh

∥∥2

0

≤ KN∑n=1

∆t

[ ∥∥∥Cn − PhCn∥∥∥2

0+ ‖(Πh − Ph) Qn‖20 +

∥∥Πh

(Qn −Qn

)∥∥2

0

]

which implies

N∑n=1

∆t ‖Cn − Cnh‖20 +

∥∥∥∥∥∥Πh

∫ T

0Qdτ −

N∑j=1

Qjh∆t

∥∥∥∥∥∥2

0

≤ KN∑n=1

∆t

[ ∥∥∥Cn − PhCn∥∥∥2

0+ ‖(Πh − Ph) Qn‖20 +

∥∥Πh

(Qn −Qn

)∥∥2

0

] (6.1.41)

Now we wish to bound∑N

n=1 ∆t ‖C − Ch‖20 dt below by∥∥∥∑N

n=1 ∆t(PhCn − Cnh

)∥∥∥2

0.

To do so we consider the fact that∑N

n=1 ∆t(C − Ch)dt ∈ L2 so

PhN∑n=1

∆t(C − Ch)dt =

N∑n=1

∆tPh(Cn − Cnh ) =

N∑n=1

∆t(PhCn − Cnh ) ∈ Vh.

Then

(N∑n=1

∆t(PhCn − Cnh ), vh

)=

(N∑n=1

∆t(C − Ch)dt, vh

)

144

Page 156: Modeling, Analysis, and Simulation of Adsorption in ...

for all vh ∈ Vh. We choose vh =∑N

n=1 ∆t(PhCn − Cnh ) to obtain

∥∥∥∥∥N∑n=1

∆t(PhCn − Cnh )

∥∥∥∥∥2

0

=

(N∑n=1

∆t(C − Ch),N∑n=1

∆t(PhCn − Cnh )

).

Bounding the right side by using the Cauchy-Schwarz Inequality, we see

∥∥∥∥∥N∑n=1

∆t(PhCn − Cnh )

∥∥∥∥∥2

0

∥∥∥∥∥N∑n=1

∆t(C − Ch)

∥∥∥∥∥0

∥∥∥∥∥N∑n=1

∆t(PhCn − Cnh )

∥∥∥∥∥0

,

and consequently,

∥∥∥∥∥N∑n=1

∆t(PhCn − Cnh )

∥∥∥∥∥0

∥∥∥∥∥N∑n=1

∆t(C − Ch)

∥∥∥∥∥0

∥∥∥∥∥N∑n=1

∆t(PhCn − Cnh )

∥∥∥∥∥2

0

∥∥∥∥∥N∑n=1

∆t(C − Ch)

∥∥∥∥∥2

0

≤ KN∑n=1

∆t ‖(C − Ch)‖20 .

Using the above inequality in (6.1.41) gives

∥∥∥∥∥N∑n=1

∆t(PhCn − Cnh

)∥∥∥∥∥2

0

+

∥∥∥∥∥Πh

∫ T

0Qdτ −

N∑n=1

Qnh∆t

∥∥∥∥∥2

0

≤ K

(N∑n=1

∆t∥∥∥Cn − PhCn∥∥∥2

0+

N∑n=1

∆t ‖(Πh − Ph) Qn‖20

+

N∑n=1

∆t∥∥Πh

(Qn −Qn

)∥∥2

0

(6.1.42)

145

Page 157: Modeling, Analysis, and Simulation of Adsorption in ...

By adding and subtracting appropriate terms and applying the triangle inequality,

∥∥∥∥∥∫ T

0Cdτ −

N∑n=1

Cnh∆t

∥∥∥∥∥2

0

≤K(∥∥∥∥∥

N∑n=1

∆t(Cn − Cn

)∥∥∥∥∥2

0

+

∥∥∥∥∥N∑n=1

∆t(Cn − PhCn

)∥∥∥∥∥2

0

+

∥∥∥∥∥N∑n=1

∆t(PhCn − Cnh

)∥∥∥∥∥2

0

).

≤K( N∑n=1

∆t∥∥Cn − Cn∥∥2

0+

N∑n=1

∆t∥∥∥Cn − PhCn∥∥∥2

0

+

∥∥∥∥∥N∑n=1

∆t(PhCn − Cnh

)∥∥∥∥∥2

0

)

where Cn is defined similarly to Qn

and again we assume ∆t small for the last inequality.

Similarly,

∥∥∥∥∥∫ T

0Qdτ −

N∑n=1

Qnh∆t

∥∥∥∥∥2

0

≤K(∥∥∥∥∥

N∑n=1

∆t(Qn −ΠhQ

n)∥∥∥∥∥2

0

+

∥∥∥∥∥Πh

∫ T

0Qdτ −

N∑n=1

Qnh∆t

∥∥∥∥∥2

0

)

≤K

(N∑n=1

∆t∥∥Qn −ΠhQ

n∥∥2

0+

∥∥∥∥∥Πh

∫ T

0Qdτ −

N∑n=1

Qnh∆t

∥∥∥∥∥2

0

).

Combining the above bounds with (6.1.42), we obtain

∥∥∥∥∥∫ T

0Cdτ −

N∑n=1

Cnh∆t

∥∥∥∥∥2

0

+

∥∥∥∥∥∫ T

0Qdτ −

N∑n=1

Qnh∆t

∥∥∥∥∥2

0

≤ K( N∑n=1

∆t∥∥Cn − Cn∥∥2

0+

N∑n=1

∆t∥∥∥Cn − PhCn∥∥∥2

0

+N∑n=1

∆t∥∥Qn −ΠhQ

n∥∥2

0+

N∑n=1

∆t ‖(Πh − Ph) Qn‖20

+

N∑n=1

∆t∥∥Πh

(Qn −Qn

)∥∥2

0

).

(6.1.43)

The second and fourth terms on the right can be bounded similarly to the semi-continuous

146

Page 158: Modeling, Analysis, and Simulation of Adsorption in ...

case:

∥∥∥Cn − PhCn∥∥∥2

0≤ Kh2s ‖Cn‖2s ,

‖(Πh − Ph) Qn‖20 ≤ Kh2 ‖Qn‖21 .

Also, the first and fifth terms can be bounded the same as the residual term, rn, in the fully

explicit error analysis completed in section 5.1.2 to obtain the following bounds:

∥∥Cn − Cn∥∥2

0≤ K∆t

∫ tn

tn−1

‖Ct‖20∥∥Πh

(Qn −Qn

)∥∥2

0≤ K∆t

∫ tn

tn−1

‖ΠhQt‖20

For the third term, we add and subtract appropriately to obtain terms that we have bounded

previously as follows:

∥∥Qn −ΠhQn∥∥2

0≤ K

(∥∥Qn −Qn∥∥+ ‖Qn −ΠhQ

n‖+∥∥Πh

(Qn − Q

n)∥∥)≤ K

(∆t

∫ tn

tn−1

‖Qt‖20 + h2‖Qn‖21 + ∆t

∫ tn

tn−1

‖ΠhQt‖20

)

147

Page 159: Modeling, Analysis, and Simulation of Adsorption in ...

Putting the above bounds together with (6.1.43) and combining like terms we obtain

∥∥∥∥∥∫ T

0Cdτ −

N∑n=1

Cnh∆t

∥∥∥∥∥2

0

+

∥∥∥∥∥∫ T

0Qdτ −

N∑n=1

Qnh∆t

∥∥∥∥∥2

0

≤K( N∑n=1

(∆t)2

∫ tn

tn−1

‖Ct‖20 +N∑n=1

∆th2s ‖Cn‖2s

+N∑n=1

(∆t)2

∫ tn

tn−1

‖Qt‖20 +

N∑n=1

∆th2 ‖Qn‖21

+N∑n=1

(∆t)2

∫ tn

tn−1

‖ΠhQt‖20

)≤K

(h2s

∫ T

0‖Cn‖2s + (∆t)2

∫ T

0‖Ct‖20 + h2

∫ T

0‖Q‖21

+ (∆t)2

∫ T

0‖Qt‖

20 + (∆t)2

∫ T

0‖ΠhQt‖

20

)≤K

((h2s + (∆t)2

)‖C‖2H1(0,T ;Hs(Ω))

+(h2 + (∆t)2

)‖Q‖2H1(0,T ;H2(Ω))

+ (∆t)2 ‖ΠhQt‖2L2(0,T ;L2(Ω))

).

6.2 Time-Integrated, SUPG Formulation

For the time-integrated, SUPG discretization scheme, we consider the same problem

and use the same notation as with the time-integrated, mixed method discretization scheme.

The assumptions are similar with one additional assumption on the q required because of

the use of a Taylor expansion in the analysis. The assumptions for the time-integrated

SUPG analysis are below.

(D1) ω and ρs are constant in time and space. [50]

(D2) u is nonzero, independent of time [34] and essentially bounded in space [147, 34] with

∇ · u = 0 [34].

148

Page 160: Modeling, Analysis, and Simulation of Adsorption in ...

(D3) D = [dij ]i,j=1...n is symmetric positive definite [4, 50], independent of time (since u

is assumed independent of time), and it and its derivatives are essentially bounded

in space [4, 147, 34]; specifically, we notate the boundedness of D using ‖D‖∞ ≤ β1,∣∣∣ ∂∂xidij∣∣∣ ≤ β2, for all i, j.

(D4) C is nonnegative [50, 9, 48] and is at least in L2(0, T ;H1 (Ω)

)[147], and C = 0 on

ΓD. ,

(D5) q = q(C) ∈ C1 is an explicit [147, 50], Lipschitz continuous [147, 134] function of C,

q(0) = 0 [50, 9, 10], q(C) > 0 for C > 0 [9, 10], and q is nondecreasing [147, 9, 10, 48,

132, 137] .

6.2.1 Development of Time-Integrated, SUPG Formulation

For the discretization with upwinding, we begin by integrating (2.2.1) in time to

obtain

ωC + (1− ω)ρs q(C) +

∫ t

0u · ∇Cdτ −

∫ t

0∇ · (D∇C)dτ

=

∫ t

0fdτ + ωC0.

(6.2.1)

Multiplying (6.2.1) by v ∈ H10,D(Ω), integrating over Ω, and upwinding as done previously,

we obtain the following continuous time-integrated SUPG variational formulation: Find

C ∈ V such that

(ωC, v) + ((1− ω)ρs q(C), v)−(∫ t

0u · ∇Cdτ, v

)+

(∫ t

0D∇Cdτ,∇v

)+

(∫ t

0u · ∇Cdτ, δu · ∇v

)−(∫ t

0∇ · (D∇C), δu · ∇v

)=

(∫ t

0fdτ, v + δu · ∇v

)+ (ωC0, v) .

(6.2.2)

for all v ∈ V . Note that the natural boundary condition on the outflow in this case ends

up implying a no flux boundary condition.

As with the original SUPG formulation and the mixed method formulation, we let

149

Page 161: Modeling, Analysis, and Simulation of Adsorption in ...

Th = E be a triangulation of Ω so that

Ω =⋃E∈Th

E,

and we define the concentration space to be

Vh := v ∈ V : v|E ∈ Pk(E), ∀E ∈ Th,

Although we allow for higher order polynomials in Vh in the analysis, we normally consider

only linear functions, that is P1(E), to simplify the upwinded terms. As Vh ⊂ L2(Ω), we

again have the standard L2 projection Ph : L2(Ω)→ Vh defined in (6.1.12) which we will use

during the error analysis of the upwinded formulation. Also, we define the term containing

a second order derivative similarly to the original upwinding formulation:

(∫ t

0∇ · (D∇C) ,u · ∇v

)=∑E∈Th

∫E

(∫ t

0∇ · (D∇C) dτ

)(u · ∇v) dE

Therefore, the semi-continuous time-integrated SUPG variational formulation is given by

the following: Find Ch ∈ Vh such that

(ωCh, vh) + ((1− ω)ρs q(Ch), vh)−(∫ t

0u · ∇Chdτ, v

)+

(∫ t

0D∇Chdτ,∇vh

)+

(∫ t

0u · ∇Chdτ, δu · ∇vh

)−(∫ t

0∇ · (D∇Ch)dτ, δu · ∇vh

)=

(∫ t

0fdτ, vh + u · ∇vh

)+ (ωC0, vh) .

(6.2.3)

for all vh ∈ Vh.

To formulate the fully discrete variational formulation, we partition the time interval

[0, T ] as

t0 = 0 < t1 < · · · < tN = T

and let ∆t = tn+1 − tn denote the step size for t where tn = n∆t. Then the fully discrete

150

Page 162: Modeling, Analysis, and Simulation of Adsorption in ...

time-integrated SUPG variational formulation can be written as follows: For n = 1, ..., N ,

find Cnh ∈ Vh such that

(ωCnh , vh) + ((1− ω)ρs q(Cnh ), vh)−

n∑j=1

u · ∇Cjh∆t, vh

n∑j=1

D∇Cjh∆t,∇vh

+

n∑j=1

u · ∇Cjh∆t, δu · ∇vh

n∑j=1

∇ · (D∇Cjh)∆t, δu · ∇vh

=

n∑j=1

f j∆t, vh + δu · ∇vh

+ (ωC0, vh) .

(6.2.4)

for all vh ∈ Vh. Note that in practical computation, a simpler, yet equivalent, form can be

used. Subtracting (6.2.4) at time level n from (6.2.4) at time level n − 1 and dividing by

∆t, we obtain for n ≥ 1

(ωCnh − C

n−1h

∆t, v

)+

((1− ω)ρs

q (Cnh )− q(Cn−1h

)∆t

, v

)+ (u · ∇Cnh , v + δ (u · ∇v)) + (D∇Cnh ,∇v)

− (∇ · (D∇Cnh ), δ (u · ∇v)) = (fn, v + δ (u · ∇v))

(6.2.5)

which contains a simple backward Euler time discretization. This simpler form is also what

is used to show solvability.

6.2.2 Solvability

We show that (6.2.4) is uniquely solvable for Ch at each time step n.

Lemma 6.2.1. Assume that (D1)-(D5) are satisfied. Then there exists a unique solution

Cnh ∈ Xh satisfying (6.2.4) (or equivalently (6.2.5)).

Proof. For notational simplicity, we drop the subscript h from the variable. Choosing

151

Page 163: Modeling, Analysis, and Simulation of Adsorption in ...

v = Cn in (6.2.5) gives

a (Cn, Cn) =ω

∆t

(Cn−1, Cn

)+

(1− ω)ρs∆t

(q(Cn−1

), Cn

)+ (fn, Cn + δu · ∇Cn)

where a(·, ·) is defined as

a (Cn, Cn) :=ω

∆t(Cn, Cn) +

(1− ω)ρs∆t

(q (Cn) , Cn)− (u · ∇Cn, Cn)

+ δ (u · ∇Cn,u · ∇Cn) + (D∇Cn,∇Cn)− δ (∇ · (D∇Cn) ,u · ∇Cn) .

We note a (·, ·) is continuous and strongly monotonic on Vh × Vh, which guarantees

the existence of a unique solution [177]. We provide lower bounds on individual terms in

the operator a(·, ·) below, where we make use of the assumptions on q(C), the velocity

u, and the dispersion tensor D. We also use Young’s Inequality and the Cauchy-Schwarz

Inequality in the analysis shown below.

We have

ω

∆t(Cn, Cn) =

ω

∆t‖Cn‖20 ,

(1− ω)ρs∆t

(q (Cn) , Cn) ≥ (1− ω)ρs∆t

(0, Cn) = 0,

(u · ∇Cn, Cn) =1

2

∫Ω

u · ∇ (Cn)2 dΩ

=1

2

∫Γ(Cn)2 (u · n) ds−

∫Ω

(∇ · u)︸ ︷︷ ︸=0

(Cn)2dΩ

=

1

2

∫ΓD

(Cn)2︸ ︷︷ ︸=0

(u · n)ds+1

2

∫ΓN

(Cn)2 (u · n)︸ ︷︷ ︸≥0

ds

≥ 0,

152

Page 164: Modeling, Analysis, and Simulation of Adsorption in ...

δ (u · ∇Cn,u · ∇Cn) = δ ‖u · ∇Cn‖20 ,

(D∇Cn,∇Cn) =∥∥∥D1/2∇Cn

∥∥∥2

0≥ λ ‖∇Cn‖20 ,

and

(∇ · (D∇Cn) , δu · ∇Cn) ≤δε2‖∇ · (D∇Cn)‖20 +

δ

2ε‖u · ∇Cn‖20

≤δε2

[β2

1K2i h−2 ‖∇Cn‖20 + β2

2K2i h−2 ‖Cn‖20

]+

δ

2ε‖u · ∇Cn‖2

≤δK2i h−2ε

2

[β2

1 ‖∇Cn‖20 + β2

2 ‖Cn‖20

]+

δ

2ε‖u · ∇Cn‖20 .

Combining like terms and choosing ε = 1 for simplicity gives

a (Cn, Cn) ≥(ω

∆t− δβ2

2K2i h−2

2

)‖Cn‖20 +

(λ− β2

1δK2i h−2

2

)‖∇Cn‖20

2‖u · ∇Cn‖20 .

To have strong monotonicity of the operator, we must have

∆t− δβ2

2K2i h−2

2

)> 0 and

(λ− β2

1δK2i h−2

2

)> 0.

The choice of δ = O(h2)

would seem to be necessary for both inequalities. However,

we recall that upwinding is only necessary in the case when β1 ≤ h and δ = 0 when β1 > h

so that these terms do not arise if β1 > h [83]. Consequently, we use the fact that β1 ≤ h

to control the h−2 obtained from the inverse inequality in the second positivity constraint,

and hence δ = O(h) is all that is necessary for the second inequality. Specifically when

β1 ≤ h, then we write δ = Kh so that

λ− β21δK

2i h−2

2≥ λ− β1KK

2i

2

153

Page 165: Modeling, Analysis, and Simulation of Adsorption in ...

and choosing

K <2λ

β1K2i

ensures positivity of the coefficient of ‖u · ∇Cn‖20.

For the first inequality, a more restrictive condition on δ, δ = O(h2), can be used,

or the h−2 can be controlled with a combination of a O(h) condition on both δ and ∆t. It

is important to note that the more restrictive condition on δ (or potentially ∆t) arises only

in the case when D varies spatially. If D is constant, then δ = O(h) is sufficient to obtain

strong monotonicity of the operator.

The continuity of a (·, ·) is obtained in a similar manner, using the upper bounds on

q (C), u, and D and associated derivatives, along with the continuity of q (C).

6.3 Fully Implicit SUPG Formulation

For this discretization, we consider the same problem and notation as described in

the development of the original discretization (Chapter 3) since the fully implicit SUPG

formulation is essentially a simplified version of the original discretization in (3.3.7)-(3.3.8).

The assumptions are similar to ones we have considered previously; in particular, they are

the same as the assumptions in section 6.2 with an additional assumption on the regularity

of q.

We assume the following:

(E1) ω and ρs are constant in time and space. [50]

(E2) u is nonzero, independent of time [34] and essentially bounded in space [147, 34] with

∇ · u = 0 [34].

(E3) D = [dij ]i,j=1...n is symmetric positive definite [4, 50], independent of time (since u

is assumed independent of time), and it and its derivatives are essentially bounded

in space [4, 147, 34]; specifically, we notate the boundedness of D using ‖D‖∞ ≤ β1,∣∣∣ ∂∂xidij∣∣∣ ≤ β2, for all i, j.

154

Page 166: Modeling, Analysis, and Simulation of Adsorption in ...

(E4) C is nonnegative and bounded [50, 9, 48], and is at least in L2(0, T ;H1 (Ω)

)[147],

and C = 0 on ΓD.

(E5) C and Ch are nondecreasing in time for the time interval of interest (0 ≤ t ≤ T ).

(E6) q = q(C) ∈ C1 is an explicit [147, 50], Lipschitz continuous [147, 134] function of C,

q(0) = 0 [50, 9, 10], q(C) > 0 for C > 0 [9, 10], and q is nondecreasing [147, 9, 10, 48,

132, 137] .

(E7) The rate of increase in adsorption is Lipschitz continuous and bounded above so that

dqdC = q′ (C) ≤ κ2 [50].

The nondecreasing conditions on C and Ch in (E5) and the Lipschitz continuous

requirement on q′(C) in (E7) are new assumptions in this section. The nondecreasing

assumption on C in (E5) is realistic for the binding phase of the chromatography process

(that which occurs during the time interval in consideration in this work). We assume the

same condition on the discrete solution Ch as any discrete solution which does not satisfy this

condition is an unrealistic solution and hence would be discarded. The Lipschitz continuous

requirement on q′ has not to our knowledge been used in previous analysis; however, it is

reasonable for realistic data (as seen in [167]) and can be shown to be true in the case of

the Langmuir isotherm model.

6.3.1 Development of Fully Implicit SUPG Formulation

To develop the fully implicit SUPG formulation, we return to the original fully

implicit formulation given by (3.3.7)-(3.3.8). Since we have an explicit representation for q

by assumption (E6), we rewrite the adsorption term as follows:

∂q

∂t=

∂q

∂C

∂C

∂t= q′(C)

∂C

∂t(6.3.1)

155

Page 167: Modeling, Analysis, and Simulation of Adsorption in ...

where for simplicity of notation, we have q′(C) = ∂q∂C . Then the continuous SUPG varia-

tional formulation can be written as the following: Find C ∈ V such that

(ω ∂C∂t , v

)+((1− ω)ρsq

′(C)∂C∂t , v)

+ (u · ∇C, v) + δ (u · ∇C,u · ∇v)

+ (D∇C,∇v)− δ(∇ · (D∇C),u · ∇v) = (f, v) + δ (f,u · ∇v) .(6.3.2)

for all v ∈ V .

For the fully discrete variational formulation, we approximate the temporal deriva-

tive of C with the usual finite difference approximation, dt(C). Then the fully implicit

SUPG variational formulation is written as follows: For n = 1, ..., N , find Cnh ∈ Vh such

that

(ωdt(Cnh ), vh) +

((1− ω)ρsq

′(Cnh )dt(Cnh ), vh

)+ (u · ∇Cnh , vh)

+ δ (u · ∇Cnh ,u · ∇vh) + (D∇Cnh ,∇vh)− δ(∇ · (D∇Cnh ),u · ∇vh)

= (fn, vh) + δ(fn,u · ∇vh).

(6.3.3)

for all vh ∈ Vh.

6.3.2 Solvability and Stability

We begin by showing that (6.3.3) is solvable for Ch at each time step n. We first

recall the Newton-Kantarovich theorem [35] which is useful in showing solvability of the

nonlinear, fully-implicit discrete formulation.

Theorem 6.3.1. (Newton-Kantorovich Theorem) Let there be given two Banach spaces

X and Y , an open subset Z of X, a point x0 ∈ Ω, and a mapping g ∈ C1(Z;Y ) such that

g′ (x0) ∈ L (X;Y ) is a bijection, so that(g′ (x0)

)−1 ∈ L (Y ;X) .

156

Page 168: Modeling, Analysis, and Simulation of Adsorption in ...

Assume that there exist three constants λ, µ, ν such that

0 < λµν ≤ 1

2and B (x0; r) ⊂ Ω, where r :=

1

µν,∥∥∥g′ (x0)−1 g (x0)

∥∥∥X≤ λ∥∥∥g′ (x0)−1

∥∥∥L(Y ;X)

≤ µ∥∥g′ (x)− g′ (x)∥∥L(X;Y )

≤ ν ‖x− x‖X for all x, x ∈ B (x0; r) .

Then g′ (x) ∈ L (X;Y ) is a bijection and thus (g′ (x))−1 ∈ L (Y ;X) at each x ∈ B (x0; r),

and the sequence (xk)∞k=0 defined by

xk+1 = xk −(g′ (xk)

)−1g (xk) , k ≥ 0

is contained in the ball B (x0; r−), where

r− :=1−√

1− 2λµν

µν≤ r

and converges to a zero a ∈ B (x0; r−) of g. Besides, for each k ≥ 0,

‖xk − a‖X ≤r

2k

(r−r

)2k

if λµν ≤ 1

2, or ‖xk − a‖X ≤

r

2kif λµν =

1

2.

In order to show the Lipschitz continuity of g′ in the Newton-Kantorovich Theorem

and therefore prove solvability of (6.2.4), we need additional regularity on q. We thus have

one more necessary assumption.

(E8) The second derivative of the adsorption, q′′(C), is Lipschitz continuous and bounded

as a function of C.

Using the listed assumptions and the Newton-Kantorovich Theorem, we are now

equipped to prove solvability of the fully implicit SUPG formulation.

157

Page 169: Modeling, Analysis, and Simulation of Adsorption in ...

Lemma 6.3.2. Assume that (E1)-(E8) are satisfied. Then there exists a solution Cnh ∈ Vh

satisfying (6.2.4) for f ∈ L2(Ω).

Proof. For notational simplicity, we drop the subscript h from the variable. We define

X = Z = Vh, Y = R, and the operator g(x)(v) = a(x; v) based on (6.2.4) as

a(x; v) = ω(x− Cn−1, v

)+ (1− ω)ρs

(q′(x)

(x− Cn−1

), v)

+ ∆t (u · ∇x, v)

+ ∆t (u · ∇x, δu · ∇v) + ∆t (D∇x,∇v) + ∆t (∇ · (D∇x) , δu · ∇v)−∆t (f, v) .

for x, v ∈ V . Then we obtain g′(x) = b(x;w, v) by considering the Frechet derivative of g

as follows:

b(x;w, v) = ω (w, v) + (1− ω)ρs(q′(x)w, v

)+ (1− ω)ρs

(q′′(x)

(x− Cn−1

)w, v

)+ ∆t (u · ∇w, v) + ∆t (u · ∇w, δu · ∇v) + ∆t (D∇w,∇v) + ∆t (∇ · (D∇w), δu · ∇v) .

Note that we choose x0 = Cn−1 for each n.

The Newton-Kantorovich Theorem guarantees a zero of g, and thus a solution of

(6.2.4), exists provided g and g′ satisfy the assumptions. To show g and g′ are suitable and

apply Theorem 6.3.1, we structure the proof as follows.

Step 1: We obtain µ by proving the existence and boundedness of the inverse operator

(g′(x0))−1 = (g′(Cn−1))−1 operator by first showing the boundedness of g′(x0) =

g′(Cn−1).

Step 2: We find λ by proving the boundedness of the operator g(x0) = g(Cn−1) to then

show the boundedness of (g′(x0))−1g(x0) = (g′(Cn−1))−1g(Cn−1).

Step 3: We obtain ν by showing the Lipschitz continuity of g′(x).

Step 4: We show the product λµν is nonnegative and sufficiently small with an appro-

priate choice of the time step size ∆t.

Step 5: We apply the Newton Kantorovich Theorem.

Step 1: Proof of boundedness of (g′(x0))−1. We begin by showing the boundedness of

158

Page 170: Modeling, Analysis, and Simulation of Adsorption in ...

g′(x0) = g′(Cn−1). Notice

g′(Cn−1) = b(Cn−1;w, v) = ω (w, v) + (1− ω)ρs(q′(Cn−1)w, v

)+ ∆t (u · ∇w, v)

+ ∆t (u · ∇w, δu · ∇v) + ∆t (D∇w,∇v) + ∆t (∇ · (D∇w), δu · ∇v) .

and consequently g′(x0) is a linear operator. Also, for b(Cn−1; s, s) we have

ω(s, s) = ω‖s‖20 ≤ ω‖s‖21,

(1− ω)ρs(q′(Cn−1)s, s) ≤ (1− ω)ρsκ2‖s‖20 ≤ (1− ω)ρsκ2‖s‖21,

∆t (u · ∇s, s) ≤ ∆t‖u‖∞‖∇s‖0‖s‖0 ≤ ∆t‖u‖∞‖s‖21,

∆t(u · ∇s, δu · ∇v) ≤ ∆tδ‖u‖2∞‖∇s‖20 ≤ ∆tδ‖u‖2∞‖s‖21,

∆t(D∇s,∇s) ≤ ∆t‖D‖∞‖∇s‖20 ≤ ∆t‖D‖2∞‖s‖21,

∆t(∇ · (D∇w), δu · ∇s) ≤ ∆tδ‖u‖∞[β1

∥∥∇2s∥∥

0+ β2‖∇s‖0

]‖∇s‖0

≤ ∆tδ‖u‖∞[β1Kih

−1‖∇s‖0 + β2‖∇s‖0]‖∇s‖0

≤ ∆tδ‖u‖∞ (Ki6 + β2) ‖s‖1,

where in the last inequality we assume β1 ≤ h to eliminate the h−1; recall that δ = 0 when

β1 > h and consequently the term above does not appear in that case.

Consequently, we have

‖g′(x0)‖L(X;Y ) ≤ K1,

159

Page 171: Modeling, Analysis, and Simulation of Adsorption in ...

and g′(x0) is a bounded linear operator. Then the Bounded Inverse Theorem [97] guarantees

the existence of a bounded inverse which we write explicitly as

∥∥∥g′ (x0)−1∥∥∥L(Y ;X)

≤ µ.

Step 2: Proof of boundedness of (g′(x0))−1g(x0). We begin by showing the bounded-

ness of g(x0) = g(Cn−1). Notice

g(Cn−1) = a(Cn−1; v) = ∆t(u · ∇Cn−1, v

)+ ∆t

(u · ∇Cn−1, δu · ∇v

)+ ∆t

(D∇Cn−1,∇v

)+ ∆t

(∇ ·(D∇Cn−1

), δu · ∇v

)−∆t (f, v) .

(6.3.4)

We bound each term in a(Cn−1; v

)as follows.

∆t(u · ∇Cn−1, v

)≤ ∆t‖u‖∞

∥∥∇Cn−1∥∥

0‖v‖0 ≤ ∆t‖u‖∞

∥∥∇Cn−1∥∥

1‖v‖1,

∆t(u · ∇Cn−1, δu · ∇v

)≤ ∆t‖u‖2∞

∥∥∇Cn−1∥∥

0‖∇v‖0 ≤ ∆t‖u‖2∞

∥∥∇Cn−1∥∥

1‖∇v‖1,

∆t(D∇Cn−1,∇v

)≤ ∆t‖D‖∞

∥∥∇Cn−1∥∥

0‖∇v‖0 ≤ ∆t‖D‖∞

∥∥∇Cn−1∥∥

1‖∇v‖1,

∆t(∇ ·(D∇Cn−1

), δu · ∇v

)≤ ∆t

[β1

∥∥∇2Cn−1∥∥

0+ β2

∥∥∇Cn−1∥∥

0

]‖v‖0

≤ ∆t[β1h

−1∥∥∇Cn−1

∥∥0

+ β2

∥∥∇Cn−1∥∥

0

]‖v‖0

≤ ∆t (1 + β2)∥∥Cn−1

∥∥1‖v‖1,

since again we assume β1 ≤ h. Last

−∆t(f, v) ≤ ∆t‖f‖0‖v‖0 ≤ ∆t‖f‖0‖v‖1.

160

Page 172: Modeling, Analysis, and Simulation of Adsorption in ...

Consequently, we have

a(Cn−1; v

)≤ ∆tK2‖v‖.

since Cn−1 ∈ H1(Ω) implies∥∥Cn−1

∥∥1<∞ and f ∈ L2(Ω) implies ‖f‖0 <∞. As a result,

g(x0) is a bounded linear operator which we denote with

‖g(x0)‖C1(Ω;Y ) ≤ ∆tK2.

As g(x0) and (g′(x0))−1 are both bounded operators, we have

∥∥∥g′ (x0)−1 g (x0)∥∥∥X≤ ‖g(x0)‖C1(Ω;Y )‖g′(x0)‖L(X;Y ) ≤ λ,

where

λ = K1K2∆t.

Step 3: Proof of Lipschitz continuity of g′(x). We consider ‖b(x;w, v) − b(y;w, v)‖.

The linear terms will cancel in the subtraction, leaving

‖b(x;w, v)− b(y;w, v)‖ =∥∥∥(1− ω)ρs

((q′(x)− q′(y)

)w, v

)+ (1− ω)ρs

[(q′′(x)

(x− Cn−1

)w, v

)−(q′′(y)

(y − Cn−1

)w, v

)] ∥∥∥≤(1− ω)ρs

∥∥((q′(x)− q′(y))w, v

)∥∥+ (1− ω)ρs

∥∥(q′′(x)(x− Cn−1

)w, v

)−(q′′(y)

(y − Cn−1

)w, v

)∥∥ .For the first term above

∥∥((q′(x)− q′(y))w, v

)∥∥ ≤ ‖q′(x)− q′(y)‖0‖wv‖0

≤ ‖q′(x)− q′(y)‖0‖w‖L4‖v‖L4

≤ KK3‖x− y‖0‖w‖1‖v‖1

161

Page 173: Modeling, Analysis, and Simulation of Adsorption in ...

where the last inequality uses the assumption of the Lipschitz continuity of q′ and the

Sobolov Embedding Theorem to bound the L4 norm by the H1 norm.

For the second term

∥∥∥ (q′′(x)(x− Cn−1

)w, v

)−(q′′(y)

(y − Cn−1

)w, v

) ∥∥∥=∥∥∥ ((q′′(x)

(x− Cn−1

)− (q′′(y)

(y − Cn−1

))w, v

) ∥∥∥≤K

∥∥q′′(x)(x− Cn−1

)− (q′′(y)

(y − Cn−1

)∥∥0‖w‖1‖v‖1

=K∥∥q′′(x)x− q′′(y)y + (q′′(x)− q′′(y))Cn−1

∥∥0‖w‖1‖v‖1

≤K∥∥q′′(x)x− q′′(y)y

∥∥0‖w‖1‖v‖1

+K∥∥(q′′(x)− q′′(y))Cn−1

∥∥0‖w‖1‖v‖1

=K∥∥(q′′(x)− q′′(y))x+ q′′(y)(x− y)

∥∥0‖w‖1‖v‖1

+K∥∥(q′′(x)− q′′(y))Cn−1

∥∥0‖w‖1‖v‖1

≤K∥∥(q′′(x)− q′′(y))x

∥∥0‖w‖1‖v‖1 +K

∥∥q′′(y)(x− y)∥∥

0‖w‖1‖v‖1

+K∥∥(q′′(x)− q′′(y))Cn−1

∥∥0‖w‖1‖v‖1

≤K(K4‖x‖0 + ‖q′′(y)‖0 +K4‖Cn−1‖0

)‖w‖1‖v‖1‖x− y‖0.

Here the last inequality uses the Lipschitz continuity of q′′ (Assumption (E8)). Also ‖q′′(y)‖0

is bounded by Assumption (E8), ‖x‖0 and ‖Cn−1‖0 are bounded since C is bounded by

Assumption (E4), and ‖w‖1, and ‖v‖1 are finite as w, v ∈ V ⊂ H1(Ω). Consequently, we

have

‖b(x;w, v)− b(y;w, v)‖ ≤ ν‖x− y‖0 ≤ ν‖x− y‖1

where ν = (K3 +K4K5 +K6 +K4K5) ‖w‖1‖v‖1.

Step 4: Proof that 0 < λµν ≤ 12 . From step 1, we have there exists a µ such that

∥∥∥g′ (x0)−1∥∥∥L(Y ;X)

≤ µ.

162

Page 174: Modeling, Analysis, and Simulation of Adsorption in ...

Also, from steps 1 and 2 we have

λ = K1K2∆t,

where K1 and K2 satisfies

‖g′(x0)‖L(X;Y ) ≤ K1,

‖g(x0)‖C1(Ω;Y ) ≤ ∆tK2.

Last, from step 3 we

ν =(K3 +K4‖x‖0 + ‖q′′(y)‖0 +K4‖Cn−1‖0

)‖w‖1‖v‖1

where K3 and K4 are the Lipschitz continuity constants of q′ and q′′ respectively. Clearly

λ, µ, and ν are each positive and consequently λµν > 0. Also since λ is explicitly a function

of ∆t, then

λµν = K1K2∆tµν ≤ 1

2

if we choose ∆t such that

∆t ≤ 1

2K1K2µν.

Step 5: Application of the Newton-Kantorovich Theorem: As all the assumptions

of Theorem 6.3.1 are satisfied for g and g′ as defined above, then there exists a zero of g

and consequently a solution of (6.2.4).

We next state and prove an a priori stability bound for the fully discrete upwinded

formulation given by (6.3.3).

Theorem 6.3.3. Suppose the assumptions of Lemma 6.3.2 are satisfied so that the fully dis-

crete formulation given by (6.3.3) has a solution Cnh ∈ L2(0, T ;H1(Ω)) with f ∈ L2(0, T ;L2(Ω)).

163

Page 175: Modeling, Analysis, and Simulation of Adsorption in ...

Then there exists a constant K > 0 independent of h and ∆t such that for all N > 0

‖CNh ‖20 + ∆tN∑n=1

(‖u · ∇Cnh‖

20 + ‖∇Cnh‖

20

)≤ K

(‖C0

h‖20 + ∆t(1 + δ2)N∑n=1

‖fn‖20

).

Proof. We take vh = Cnh in (6.3.3) to obtain

((ω + (1− ω) ρsq

′ (Cnh ))dt (Cnh ) , Cnh

)+ (u · ∇Cnh , Cnh ) + δ ‖u · ∇Cnh‖

20

+ (D∇Cnh ,∇Cnh )− (∇ · (D∇Cnh ) , δu · ∇Cnh ) = (fn, Cnh ) + (fn, δu · ∇Cnh ) .(6.3.5)

Notice since q is nondecreasing, we have q′(C) ≥ κ1 ≥ 0 for all C. Consequently,

ω + (1− ω) ρsq′ (Cnh ) ≥ ω + (1− ω)ρsκ1

which, along with the nondecreasing and positive assumptions on Ch, gives the following

lower bound for the first term in (6.3.5):

((ω + (1− ω)) ρsq

′ (Cnh ) dt(Cnh ), Cnh

)≥ ω + (1− ω)ρsκ1

∆t

(Cnh − C

n−1h , Cnh

)≥ ω + (1− ω)ρsκ1

2∆t

(‖Cnh‖

20 −

∥∥Cn−1h

∥∥2

0

).

We obtain lower bounds for the rest of the terms on the left by using the assumptions,

164

Page 176: Modeling, Analysis, and Simulation of Adsorption in ...

the Cauchy-Schwartz Inequality, and Young’s Inequality as follows:

(u · ∇Cnh , Cnh ) =

∫Ω

u · ∇Cnh CnhdΩ

=1

2

∫Ω

u · ∇ (Cnh )2 dΩ

=1

2

∫Γ

(Cnh )2 (u · n) ds−∫

Ω(∇ · u)︸ ︷︷ ︸

=0

(Cnh )2 dΩ

=

1

2

∫ΓD

(Cnh )2︸ ︷︷ ︸=0

(u · n) ds+

∫ΓN

(Cnh )2 (u · n)︸ ︷︷ ︸≥0

ds

≥ 0,

(D∇Cnh ,∇Cnh ) ≥ λ ‖∇Cnh‖20 ,

δ (∇ · (D∇Cnh ),u · ∇Cnh ) ≤(∇ ·D : ∇Cnh + D : ∇2Cnh , δu · ∇Cnh

)≤β2δ ‖∇Cnh‖0 ‖u · ∇C

nh‖0 + β1δ

∥∥∇2Cnh∥∥

0‖u · ∇Cnh‖0

≤β2δKih−1 ‖Cnh‖0 ‖u · ∇C

nh‖0

+ β1δKih−1 ‖∇Cnh‖0 ‖u · ∇C

nh‖0

≤ β22

2ε1‖Cnh‖

20 +

δ2K2i h−2

2ε1‖u · ∇Cnh‖

20

+1

2ε2‖∇Cnh‖

20 +

δ2β21K

2i h−2

2ε2‖u · ∇Cnh‖

20

≤β22λ

2β21

‖Cnh‖20 +

δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

4‖∇Cnh‖

20 +

δ2β21K

2i h−2

λ‖u · ∇Cnh‖

20

⇒ −δ (∇ · (D∇Cnh ),u · ∇Cnh ) ≥− β22λ

2β21

‖Cnh‖20 −

δ2β21K

2i h−2

2λ‖u · ∇Cnh‖

20

− λ

4‖∇Cnh‖

20 −

δ2β21K

2i h−2

λ‖u · ∇Cnh‖

20.

165

Page 177: Modeling, Analysis, and Simulation of Adsorption in ...

The terms on the left are bounded using similar techniques as shown below.

(fn, Cnh ) + (fn, δu · ∇Cnh ) ≤ 1

2‖fn‖20 +

1

2‖Cnh‖

20 +

δ2

2ε‖fn‖20 +

1

2ε ‖u · ∇Cnh‖

20

≤ 1

2‖fn‖20 +

1

2‖Cnh‖

20 +

δ2

2ε‖fn‖20 +

‖u‖2∞2

ε ‖∇Cnh‖20

≤(

1

2+ δ2 ‖u‖2∞

λ

)‖fn‖20 +

1

2‖Cnh‖

20 +

λ

4‖∇Cnh‖

20 .

Using the above bounds in (6.3.5) and combining like terms we obtain

ω + (1− ω)ρsκ1

2∆t

(‖Cnh‖

20 −

∥∥Cn−1h

∥∥2

0

)+λ

2‖∇Cnh‖

20

+ δ

(1− 3δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

≤ 1

2

(β2

2β21

+ 1

)‖Cnh‖

20 +

(1

2+ δ2 ‖u‖2∞

λ

)‖fn‖20

(6.3.6)

Multiplying (6.3.6) through by 2∆tω+(1−ω)ρsκ1

and summing from n = 1 to N gives

∥∥ChN∥∥2

0+

2∆t

ω + (1− ω)ρsκ1

N∑n=1

λ

2‖∇Cnh‖

20

+2∆t

ω + (1− ω)ρsκ1

N∑n=1

δ

(1− 3δβ2

1K2i h−2

)‖u · ∇Cnh‖

20

≤ ∆t

ω + (1− ω)ρsκ1

N∑n=1

(β2

2β21

+ 1

)‖Cnh‖

20

+2∆t

ω + (1− ω)ρsκ1

N∑n=1

(1

2δ2 ‖u‖2∞

λ

)‖fn‖20 +

∥∥C0h

∥∥2

0

(6.3.7)

To ensure positivity of the terms on the left, we choose δ appropriately. The choice of

δ = O(h2) seems necessary to control the last term on the left. However recall that δ is

defined to be 0 when β1 > h so that the terms contributing to the second component do

not arise [50]. Consequently, we assume β1 ≤ h and use this assumption to help control the

h2. Specifically, we choose δ = Kh such that

K <2λ

3β1K2i

,

166

Page 178: Modeling, Analysis, and Simulation of Adsorption in ...

then

3δβ21K

2i h−2

2λ≤ 3Kβ2

1K2i h−1

2λ≤ 3β1K

2i

2λK < 1.

Then since the coefficients on the left are positive and the coefficients on the right

are fintie, applying Discrete Gronwall’s Inequality (Lemma 3.4.5) to (6.3.7) gives

∥∥CNh ∥∥2

0+ ∆t

N∑n=1

(‖u · ∇Cnh‖

20 + ‖∇Cnh‖

20

)≤ K

(‖C0

h‖20 + ∆t

N∑n=1

(1 + δ2) ‖fn‖20

).

167

Page 179: Modeling, Analysis, and Simulation of Adsorption in ...

Chapter 7

Nonlinear Analysis with Implicit

Adsorption

In order to consider adsorption given by an implicit formulation as in the Nfor model

in equation (2.1.11), we consider the concentration in the solid phase q as an unknown and

couple the constitutive adsorption model equation to the transport equation to obtain a

system of two unknowns as described below.

7.1 Preliminaries

In this section, we derive the variational formulations considered in this chapter and

state the assumptions necessary for the analysis.

7.1.1 Derivation of the Variational Formulation

To develop the variational formulation, we begin by using the isotherm equation

(2.2.2) to rewrite the transport equation (2.2.1). Rearranging and differentiating (2.2.2) in

time, we have

∂q

∂t− ∂g(q, C)

∂t= 0

168

Page 180: Modeling, Analysis, and Simulation of Adsorption in ...

⇒ ∂q

∂t− ∂g

∂q

∂q

∂t− ∂g

∂C

∂C

∂t= 0

⇒(

1− ∂g

∂q

)∂q

∂t=∂g

∂C

∂C

∂t

⇒ ∂q

∂t=

∂g/∂C

1− ∂g/∂q∂C

∂t= g(C, q)

∂C

∂t(7.1.1)

where we write

g(q, c) =∂g/∂C

1− ∂g/∂q(7.1.2)

for brevity. Substituting (7.1.1) in (2.2.1), we have the modified transport equation:

ω∂C

∂t+ (1− ω)ρsg(C, q)

∂C

∂t+ u · ∇C −∇ · (D∇C) = f.

Consequently, we use the following modeling equations:

ω∂C

∂t+ (1− ω)ρsg(C, q)

∂C

∂t+ u · ∇C −∇ · (D∇C) = f, (7.1.3)

q = g(C, q). (7.1.4)

Applying normal techniques (including upwinding) to (7.1.3)-(7.1.4) and using the

concentration spaces V and W as defined in Chapter 3, we obtain the following continuous

SUPG variational formulation: Find (C,w) ∈ (V,W ) such that

(ω∂C

∂t, v

)+

((1− ω)ρsg(C, q)

∂C

∂t, v

)+ (u · ∇C, v) + (u · ∇C, δu · ∇v)

+ (D∇C,∇v)− (∇ · (D∇C) , δu · ∇v) = (f, v + δu · ∇v)

(7.1.5)

(q, w) = (g(C, q), w) (7.1.6)

for all v ∈ V,w ∈W .

To formulate the fully discrete formulation, we define the finite element spaces and

partition the interval as described in Chapter 3. To resolve the linearity and implicit

169

Page 181: Modeling, Analysis, and Simulation of Adsorption in ...

definition of (7.1.6), we evaluate the right side at the previous time step giving the fully

discrete variational problem as: For each n = 1, 2, ..., N , find (Cnh , qnh) ∈ (Vh,Wh) such that

(ωdt(Cnh ), vh) +

((1− ω)ρsg(qnh , C

n−1h )dt(C

nh ), vh

)+ (u · ∇Cnh , vh)

+ (u · ∇Cnh , δu · ∇vh) + (D∇Cnh ,∇vh)− (∇ · (D∇Cnh ) , δu · ∇vh)

= (f, v + δu · ∇vh)

(7.1.7)

(qnh , wh) =(g(qn−1

h , Cn−1h ), wh

)(7.1.8)

for all vh ∈ Vh, qh ∈Wh

7.1.2 Assumptions

To analyze this formulation, we assume the following:

(F1) ω and ρs are constant in time and space. [50]

(F2) u is nonzero, independent of time [34] and essentially bounded in space [147, 34] with

∇ · u = 0 [34].

(F3) D = [dij ]i,j=1...n is symmetric positive definite [4, 50], independent of time (since u

is assumed independent of time), and it and its derivatives are essentially bounded

in space [4, 147, 34]; specifically, we notate the boundedness of D using ‖D‖∞ ≤ β1,∣∣∣ ∂∂xidij∣∣∣ ≤ β2, for all i, j.

(F4) C is nonnegative [50, 9, 48] and bounded; specifically

0 ≤ C(x, y) ≤ Cmax ∀(x, y) ∈ Ω.

(F5) C and Ch are nondecreasing in time for the time interval of interest (0 ≤ t ≤ T ).

(F6) q is nonnegative [9, 10] and bounded by the maximum adsorption constant, qmax;

specifically

0 ≤ q(x, y) ≤ qmax ∀(x, y) ∈ Ω.

170

Page 182: Modeling, Analysis, and Simulation of Adsorption in ...

(F7) C and q both have time derivatives that are continuous and essentially bounded in

time; in other words,

∂C

∂t∈ C(0, T ) ∩ L∞(0, T ) and

∂q

∂t∈ C(0, T ) ∩ L∞(0, T ).

(F8) The function g in (2.2.2) is defined such that g(0, 0) = 0.

(F9) The function g(C, q) defined by (7.1.2) is nonnegative, Lipschitz continuous, and

bounded; specifically,

|g(q1, C1)− g(q2, C2)| ≤ κ1|q1 − q2|+ κ2|C1 − C2|.

for any (q1, C2) and (q2, C2) and

|g(q, C)| ≤ κ3 ∀ (q, C).

Notice that assumption (F9) can be shown for g defined as in the Nfor adsorption

model by using the fact that C and q are both nonnegative and bounded above. As stated

with the assumptions in section 6.3, the nondecreasing assumption on C in (F5) is real-

istic for the phase of the chromatography process that occurs during the time interval in

consideration in this work, and we extend the assumption Ch to ensure a realistic discrete

solution.

7.2 Solvability and Stability

We begin by showing that (7.1.7)-(7.1.8) is uniquely solvable for Ch at each time

step n.

Lemma 7.2.1. (Solvability) Assume (F1)-(F9) are satisfied. The there exists a unique

solution Cnh ∈ Vh satisfying (7.1.7) - (7.1.8).

171

Page 183: Modeling, Analysis, and Simulation of Adsorption in ...

Proof. Since the right hand side of (7.1.8) is evaluated at t = tn−1, then qn is obtained from

known data and is solvable at each time step tn. This implies that (7.1.7) is linear in Cnh

for each n since g is evaluated at (qnh , Cn−1h ). To show Cnh exists, we need only show the

positivity of the bilinear form in (7.1.7).

Choosing vh = Cnh in (7.1.7) and rearranging, we have

a(Cnh , Cnh ) = (fn, Cnh + δu · ∇Cnh ) +

ω

∆t

(Cn−1h , Cnh

)+

(1− ω)ρs∆t

(g(qnh , C

n−1h )Cn−1

h , Cnh)

where

a(Cnh , Cnh ) =

ω

∆t‖Cnh‖

20 +

(1− ω)ρs∆t

(g(qnh , C

n−1h )Cnh , C

nh

)+ (u · ∇Cnh , Cnh )

+ δ ‖u · ∇Cnh‖20 + (D∇Cnh ,∇Cnh )− (∇ · (D∇Cnh ) , δu · ∇Cnh )

(7.2.1)

We bound each term in a(Cnh , Cnh ) as follows:

(1− ω)ρs∆t

(g(qnh , C

n−1h )Cnh , C

nh

)≥ (1− ω)ρsκ0

∆t‖Cnh‖

20

(u · ∇Cnh , Cnh ) =

∫Ω

(u · ∇Cnh )CnhdΩ

=1

2

∫Ω

u · ∇(Cnh )2dΩ

=1

2

∫Γ(Cnh )2(u · n)ds−

∫Ω

(∇ · u)︸ ︷︷ ︸=0

(Cnh )2dΩ

=

1

2

∫ΓD

(Cnh )2︸ ︷︷ ︸=0

(u · n)ds+1

2

∫ΓN

(Cnh )2 (u · n)︸ ︷︷ ︸≥0

ds

≥ 0

(D∇Cnh , Cnh ) ≥ λ ‖∇Cnh‖20

172

Page 184: Modeling, Analysis, and Simulation of Adsorption in ...

(∇ · (D∇Cnh ), δu · ∇Cnh ) = ((∇ ·D) · (∇Cnh ), δu · ∇Cnh ) +(D : ∇2Cnh , δu · ∇Cnh

)≤ β2 ‖∇Cnh‖0 ‖δu · ∇C

nh‖0 + β1

∥∥∇2Cnh∥∥

0‖δu · ∇Cnh‖0

≤ β2δ‖u‖∞‖∇Cnh‖20 + β1Kih

−1‖∇Cnh‖0‖δu · ∇Cnh‖0

≤ β2δ‖u‖∞‖∇Cnh‖20 +

1

2ε2‖∇Cnh‖

20

+β2

1K2i h−2δ2

2ε2‖u · ∇Cnh‖

20

⇒ − (∇ · (D∇Cnh ), δu · ∇Cnh ) ≥ −β2δ‖u‖∞‖∇Cnh‖20 −

1

2ε1‖∇Cnh‖

20

− β21K

2i h−2δ2

2ε1‖u · ∇Cnh‖

20

Inserting the above bounds into (7.2.1) and combining like terms, we obtain

a(Cnh , Cnh ) ≥ ω + (1− ω)ρsκ0

∆t‖Cnh‖

20 + δ

(1− β2

1K2i h−2δ

2ε1

)‖u · ∇Cnh‖

20

+

(λ− β2δ‖u‖∞ −

1

2ε1

)‖∇Cnh‖

20 .

(7.2.2)

To simplify the equation above, we choose ε1 = 1λ then (7.2.2) becomes

a(Cnh , Cnh ) ≥ ω + (1− ω)ρsκ0

∆t‖Cnh‖

20 + δ

(1− β2

1K2i h−2δ

)‖u · ∇Cnh‖

20

+

2− β2δ‖u‖∞

)‖∇Cnh‖

20 .

(7.2.3)

To ensure positivity of the components in (7.2.3), we choose δ appropriately. For

the third term on the right, we assume h ≤ 1 and choose δ = Kh such that

K <λ

2β2‖u‖∞.

Consequently

β2δ‖u‖∞ = β2Kh‖u‖∞ <λ

2h ≤ λ

2.

For the positivity of the second term in (7.2.3), it seems that we must choose δ ∼

O(h2). However, we will use the assumption that β1 ≤ h to help control the h2 in the term.

173

Page 185: Modeling, Analysis, and Simulation of Adsorption in ...

Additionally, we choose δ with

K <2

β1Ki

so that

β21Kih

−1δ

2≤ β1Kih

−1δ

2≤ β1KiK

2< 1.

Consequently, we choose δ = Kh such that

K ≤ min

λ

2β2‖u‖∞,

2

β1Ki

to ensure positivity of a(·, ·).

Thus Ker(a) = 0 and since (7.1.7) - (7.1.8) represents a square system of linear

equations, a unique solution exists.

Theorem 7.2.2. (Stability Bound) Suppose the assumptions of Lemma 7.2.1 are satisfied

so that the fully discrete SUPG variational formulation (7.1.7) - (7.1.8) has a solution

(Cnh , qnh) ∈ Vh ×Wh for f ∈ L2(Ω). Assume also that ?? holds and that ∆t is chosen to be

sufficiently small; specifically,

∆t ≤ ω + (1− ω)ρsκ0

(κ1κ4 + κ2)2.

Then there exists a constant K > 0 such that

‖ChN‖20 + ∆tN∑n=1

(‖∇Cnh‖

20 + ‖u · ∇Cnh‖

20

)+ ∆t

N∑n=1

‖qnh‖20

≤ K

(∆t(1 + δ2)

N∑n=1

‖fn‖20 + ‖C0h‖

20

).

174

Page 186: Modeling, Analysis, and Simulation of Adsorption in ...

Proof. Choose vh = Cnh in (7.1.7) to obtain

(ωdt(Cnh ), Cnh ) + ((1− ω)ρsg(qnh , C

n−1h )dt(C

nh ), Cnh ) + (u · ∇Cnh , Cnh )+

δ‖u · ∇Cnh‖20 + (D∇Cnh ,∇Cnh )− (∇ · (D∇Cnh ), δu · ∇Cnh )

= (fn, Cnh + δu · ∇Cnh )

(7.2.4)

The terms on the left are bounded below as follows.

(ωdt(Cnh ), Cnh ) =

ω

2∆t

(‖Cnh‖

20 + ‖Cnh − C

n−1h ‖20 − ‖Cn−1

h ‖20)

((1− ω)ρsg(qnh , C

n−1h )dt(C

nh ), Cnh

)≥ ((1− ω)ρsκ0dt(C

nh ), Cnh )

≥ (1− ω)ρsκ0

2∆t

(‖Cnh‖

20 − ‖Cn−1

h ‖20)

(u · ∇Cnh , Cnh ) =

∫Ω

(u · ∇Cnh )CnhdΩ

=1

2

∫Ω

u · ∇(Cnh )2dΩ

=1

2

∫Γ(Cnh )2(u · n)ds−

∫Ω

(∇ · u)︸ ︷︷ ︸=0

(Cnh )2dΩ

=

1

2

∫ΓD

(Cnh )2︸ ︷︷ ︸=0

(u · n)ds+1

2

∫ΓN

(Cnh )2 (u · n)︸ ︷︷ ︸≥0

ds

≥ 0

(D∇Cnh ,∇Cnh ) ≥ λ‖∇Cnh‖20

175

Page 187: Modeling, Analysis, and Simulation of Adsorption in ...

(∇ · (D∇Cnh ), δu · ∇Cnh ) = ((∇ ·D) : ∇Cnh , δu · ∇Cnh ) + (D : ∇2Cnh , δu · ∇Cnh )

≤ β2‖∇Cnh‖0‖δu · ∇Cnh‖0 + β1‖∇2Cnh‖0‖δu · ∇Cnh‖0

≤ β2Kih−1‖Cnh‖0‖δu · ∇Cnh‖0 + β1Kih

−1‖∇Cnh‖0‖δu · ∇Cnh‖0

≤ β22

2ε1‖Cnh‖

20 +

K2i h−2δ2

2ε1‖u · ∇Cnh‖

20

+1

2ε2‖∇Cnh‖

20 +

β21K

2i h−2δ2

2ε2‖u · ∇Cnh‖

20

⇒ − (∇ · (D∇Cnh ), δu · ∇Cnh ) ≥ − β22

2ε1‖Cnh‖

20 −

K2i h−2δ2

2ε1‖u · ∇Cnh‖

20

− 1

2ε2‖∇Cnh‖

20 −

β21K

2i h−2δ2

2ε2‖u · ∇Cnh‖

20

The one term on the right is bounded above using standard techniques.

(fn, Cnh + δu · ∇Cnh ) ≤ ‖fn‖0‖Cnh‖+δ‖|u‖∞‖fn‖0‖∇Cnh‖0

≤ 1

2‖fn‖20 +

1

2‖Cnh‖

20 +

δ2‖|u‖2∞2ε3

‖fn‖20 +1

2ε3‖∇Cnh‖

20

Inserting the above bounds into (7.2.4) and combining like terms, we see

ω + (1− ω)ρsκ0

2∆t

(‖Cnh‖

20 + ‖Cnh − C

n−1h ‖20 − ‖Cn−1

h ‖20)

2‖∇Cnh‖

20

+ δ

(1− 2β2

1K2i h−2

λδ

)‖u · ∇Cnh‖

20

≤(

1

2+β2

4β21

)‖Cnh‖

20 +

(1

2+‖u‖∞λ

4δ2

)‖fn‖20

(7.2.5)

We now turn our attention to (7.1.8), the equation in terms of q. Choosing wh = qnh ,

we see

‖qnh‖20 =

(g(qn−1

h , Cn−1h ), qnh

). (7.2.6)

176

Page 188: Modeling, Analysis, and Simulation of Adsorption in ...

Bounding the term on the right as

(g(qn−1h , Cn−1

h ), qnh) = (|g(qn−1h , Cn−1

h )− g(0, 0)|, |qnh |)

≤ (κ1|qn−1h − 0|+ κ2|Cn−1

h − 0|, |qnh |)

≤ (κ1κ4|Cn−1h − 0|+ κ2|Cn−1

h − 0|, |qnh |)

≤ ((κ1κ4 + κ2)|Cn−1h |, |qnh |)

≤ (κ1κ4 + κ2)‖Cn−1h ‖0‖qnh‖0

≤ (κ1κ4 + κ2)2

2‖Cn−1

h ‖20 +1

2‖qnh‖

20

≤ (κ1κ4 + κ2)2

2‖Cnh − C

n−1h ‖20 +

(κ1κ4 + κ2)2

2‖Cnh‖

20 +

1

2‖qnh‖

20

and combining this bound with (7.2.6), we have

‖qnh‖20 ≤ (κ1κ4 + κ2)2‖Cnh − C

n−1h ‖20 + (κ1κ4 + κ2)2‖Cnh‖

20. (7.2.7)

Combining (7.2.7) with (7.2.5) gives

ω + (1− ω)ρsκ0

∆t

(‖Cnh‖

20 +

(1− ∆t(κ1κ4 + κ2)2

ω + (1− ω)ρsκ0

)‖Cnh − C

n−1h ‖20 − ‖Cn−1

h ‖20)

2‖∇Cnh‖

20 + δ

(1− β2

1K21h−2

λδ

)‖u · ∇Cnh‖

20 + ‖qnh‖

20

≤(

1

2+β2

4β21

+ (κ1κ4 + κ2)2

)‖Cnh‖

20 +

(1

2+‖u‖2∞λ

4δ2

)‖fn‖20

177

Page 189: Modeling, Analysis, and Simulation of Adsorption in ...

Multiplying by ∆tω+(1−ω)ρsκ0

and summing over time step n = 1 to N , we obtain

‖ChN‖20 +∆t

ω + (1− ω)ρsκ0

N∑n=1

(1− ∆t(κ1κ4 + κ2)2

ω + (1− ω)ρsκ0

)‖Cnh − C

n−1h ‖20

+∆t

ω + (1− ω)ρsκ0

N∑n=1

λ

2‖∇Cnh‖

20

+∆t

ω + (1− ω)ρsκ0

N∑n=1

δ

(1− β2

1K2i h−2

λδ

)‖u · ∇Cnh‖

20

+∆t

ω + (1− ω)ρsκ0

N∑n=1

‖qnh‖20

≤ ∆t

ω + (1− ω)ρsκ0

N∑n=1

(1

2+β2

4β21

+ (κ1κ4 + κ2)2

)‖Cnh‖

20

+∆t

ω + (1− ω)ρsκ0

N∑n=1

(1

2+‖u‖2∞λ

4δ2

)‖fn‖20 + ‖Ch0‖20.

To ensure positivity of the first sum on the left, we have a requirement on ∆t,

namely

∆t ≤ ω + (1− ω)ρsκ0

(κ1κ4 + κ2)2.

The choice of δ = O(h2)

would seem to be necessary for the third summation on the left.

However, as noted in Johnson [83], upwinding is not needed for β1 > h; in this case, δ = 0

and this term does not appear. If β1 ≤ h, then it serves to control the h−2 obtained from

the inverse inequality, and hence δ = O(h) is all that is necessary. Specifically when β1 ≤ h,

then we write δ = Kh so that

1− 2β21K

2i h−2δ

λ≥ 1− 2β1K

2i K

λ.

Consequently choosing

K <λ

2β1K2i

ensures positivity of the coefficient of ‖u · ∇Cnh‖20.

With positivity of all the terms on the left, we drop the ‖Cnh − Cn−1h ‖20 term and

178

Page 190: Modeling, Analysis, and Simulation of Adsorption in ...

manipulate the constants to obtain

‖ChN‖20 + ∆tN∑n=1

‖∇Cnh‖20 + ∆t

N∑n=1

‖u · ∇Cnh‖20 + ∆t

N∑n=1

‖qnh‖20

≤ K

(∆t

N∑n=1

‖Cnh‖20 + ∆t(1 + δ2)

N∑n=1

‖fn‖20 + ‖C0h‖

20

).

Then applying the Discrete Gronwall’s Inequality gives

‖ChN‖20 + ∆tN∑n=1

(‖∇Cnh‖

20 + ‖u · ∇Cnh‖

20

)+ ∆t

N∑n=1

‖qnh‖20

≤ K

(∆t(1 + δ2)

N∑n=1

‖fn‖20 + ‖C0h‖

20

).

179

Page 191: Modeling, Analysis, and Simulation of Adsorption in ...

Chapter 8

Numerical Results

In this section, we present four numerical experiments. The first experiment con-

firms theoretical convergence rates for the numerical approximation to the steady-state

advection-diffusion equation. The second and third experiments confirm theoretical con-

vergence rates for the fully explicit and fully implicit approximations respectively to the

time-dependent advection-diffusion-reaction equation. The fourth experiment compares re-

sults from a numerical simulation to experimental data in an effort to determine better

ways to accurately predict breakthrough.

The results for the convergence rates were all obtained using FreeFEM codes while

the numerical simulation results used a combination of FreeFEM and deal.II codes. Specifi-

cally, the simulation results for involving instantaneous adsorption with Langmuir’s isotherm

model were obtained from deal.II codes while the instantaneous adsorption results involv-

ing Nfor’s isotherm were obtained using FreeFEM. Deal.II was used instead of FreeFEM for

the Langmuir results in order to obtain 3D results. The solution algorithm for implicitly-

defined isotherms was written first using FreeFEM for ease and has yet to be implemented

in deal.II. Consequently the results obtained for Nfor’s isotherm use a FreeFEM code. All

of the results involving non-instantaneous adsorption were obtained using FreeFEM. The

two- and three-dimensions results comparing different velocity profiles were all obtained

using deal.II codes.

180

Page 192: Modeling, Analysis, and Simulation of Adsorption in ...

For simplicity of notation in this chapter, we use abbreviated expressions for the

norms as follows:

‖ · ‖k = ‖ · ‖Hk(Ω),

‖ · ‖j,k = ‖ · ‖Hj(0,T ;Hk(Ω)).

In particular, we will often use

‖ · ‖0,1 = ‖ · ‖L2(0,T ;H1(Ω)).

8.1 Steady-State Convergence Rates

First we present convergence rates for the steady-state advection diffusion problem.

For the steady-state case, we take u = 〈1, 1〉, D = I, Ω = [0, 1] × [0, 1], Vh the space of

continuous piecewise linear functions, and f and the boundary conditions determined by

the true solution

C(x, y) = (x− x2) sin(πy).

Note that for simplicity, this function has homogeneous Dirichlet boundary conditions.

Recall from chapter 4 that for the steady state problem with Dirichlet boundary conditions,

the predicted convergence rates using linear approximating functions are

‖C − Ch‖1 ≈ O(h+ δ).

Consequently, we would expect O(h) convergence rates if no upwinding is used; if upwinding

is used, then we expect that taking δ ≈ O(h) would ensure the same rate of convergence.

The numerical results presented in Table 8.1 are consistent with this estimate.

181

Page 193: Modeling, Analysis, and Simulation of Adsorption in ...

δ ↓ h → 18

116

132

164

1128

0

‖C − Ch‖0 5.46E-03 1.39E-03 3.48E-04 8.70E-05 2.18E-05

Rate – 1.97 1.99 2.00 2.00

‖∇C −∇Ch‖0 1.14E-01 5.75E-02 2.88E-02 1.44E-02 7.20E-03

Rate – 0.99 1.00 1.00 1.00

‖C − Ch‖1 1.14E-01 5.75E-02 2.88E-02 1.44E-02 7.20E-03

Rate – 0.99 1.00 1.00 1.00

h

‖C − Ch‖0 1.64E-02 8.90E-03 4.62E-03 2.35E-03 1.19E-03

Rate – 0.92 0.96 0.98 0.99

‖∇C −∇Ch‖0 1.59E-01 8.48E-02 4.36E-02 2.21E-02 1.11E-02

Rate – 0.94 0.97 0.99 0.99

‖C − Ch‖1 1.60E-01 8.52E-02 4.39E-02 2.22E-02 1.12E-02

Rate – 0.94 0.97 0.99 0.99

h2

‖C − Ch‖0 5.99E-03 1.55E-03 3.90E-04 9.78E-05 2.45E-05

Rate – 1.93 1.98 2.00 2.00

‖∇C −∇Ch‖0 1.15E-01 5.76E-02 2.88E-02 1.44E-02 7.20E-03

Rate – 1.00 1.00 1.00 1.00

‖C − Ch‖1 1.15E-01 5.76E-02 2.88E-02 1.44E-02 7.20E-03

Rate – 1.00 1.00 1.00 1.00

Table 8.1: Approximation errors and experimental convergence rates for the approximationto the steady-state problem. As ∆t and h are cut in half, the H1 error is reduced the sameamount which is consistent with the theoretical convergence rates.

8.2 Linear Convergence Rates

The numerical validation of the fully explicit and fully implicit a-priori estimates

were obtained using a linear isotherm model with K1 = K2 = 1 so that

q(C) = 1 + C.

For both cases, we take u =< 1, 1 >, D = I, Ω = [0, 1] × [0, 1], and Vh to be the space

of continuous, piecewise linear functions. The right-hand side f , along with the boundary

182

Page 194: Modeling, Analysis, and Simulation of Adsorption in ...

conditions and initial conditions, are determined by the true solution

C(x, y, t) = t(x− x2

)sin(πy).

8.2.1 Fully Explicit Discretization

Recall from sections 5.1.2 and 5.2.2 that the predicted convergence rate for the fully

explicit formulation using linear approximating functions is

‖C − Ch‖0,1 ≈ O(h+ ∆t)

with the time step restriction

∆t ≤ K(u,D, δ)h2.

In this case, the upwinding parameter δ does not affect the convergence rate although it

does affect the constraint on the time step. The numerical results presented in Table 8.2

are consistent with the error estimates shown previously. The numerical results presented

in Table 8.1 are consistent with this estimate.

8.2.2 Fully Implicit Discretization

Recall from sections 5.1.3 and 5.2.3 that the predicted convergence rates for the

fully implicit formulations using linear approximating functions are

‖C − Ch‖0,1 ≈ O(h(1 + δ) + ∆t).

In this case, the convergence depends on δ and just as in the steady-state case if we take

δ ≈ O(h), then we preserve the convergence rate in h. The numerical results presented in

Table 8.3 are consistent with this estimate.

183

Page 195: Modeling, Analysis, and Simulation of Adsorption in ...

(δ,∆t) ↓ h → 12

14

18

116

132

(0, h

2

16

)|||C − Ch|||∞,0 6.44E-02 1.99E-02 5.26E-03 1.33E-03 3.34E-04

Rate – 1.69 1.92 1.98 2.00

‖C − Ch‖0,0 3.70E-02 1.13E-02 2.98E-03 7.54E-04 1.89E-04

Rate – 1.71 1.93 1.98 2.00

‖∇C −∇Ch‖0,0 2.29E-01 1.28E-01 6.59E-02 3.32E-02 1.66E-02

Rate – 0.84 0.96 0.99 1.00

‖C − Ch‖0,1 2.32E-01 1.29E-01 6.60E-02 3.32E-02 1.66E-02

Rate – 0.85 0.96 0.99 1.00

(h, h

2

256

)|||C − Ch|||∞,0 6.00E-02 2.90E-02 1.69E-02 9.23E-03 4.75E-03

Rate – 1.05 0.78 0.87 0.96

‖C − Ch‖0,0 3.45E-02 1.68E-02 9.91E-03 5.43E-03 2.72E-03

Rate – 1.03 0.76 0.87 1.00

‖∇C −∇Ch‖0,0 2.34E-01 1.64E-01 9.50E-02 5.07E-02 2.56E-02

Rate – 0.51 0.79 0.91 0.99

‖C − Ch‖0,1 2.36E-01 1.65E-01 9.55E-02 5.10E-02 2.57E-02

Rate – 0.52 0.79 0.91 0.99

Table 8.2: Approximation errors and experimental convergence rates for the fully-explicitapproximation with a linear adsorption model. As ∆t and h are cut in half, the H1 erroris reduced the same amount which is consistent with the theoretical convergence rates.

8.3 Nonlinear Convergence Rates

The numerical validation of the nonlinear a-priori error estimates were obtained

using a standard nonlinear isotherm model. For all three cases, we take u =< 1, 1 >,

D = I, Ω = [0, 1] × [0, 1], and Vh (and Wh if appropriate) to be the space of continuous,

piecewise linear functions. The right-hand side f , along with the boundary conditions and

initial conditions, are determined by the true solution

C(x, y, t) = t(x− x2

) (y − y2

).

184

Page 196: Modeling, Analysis, and Simulation of Adsorption in ...

δ ↓ (∆t, h) →(

132 ,

18

) (164 ,

116

) (1

128 ,132

) (1

256 ,164

) (1

512 ,1

128

)

0

|||C − Ch|||∞,0 5.36E-03 1.36E-03 3.41E-04 8.53E-05 2.13E-05

Rate – 1.98 2.00 2.00 2.00

‖C − Ch‖0,0 3.14E-03 7.86E-04 1.96E-04 4.89E-05 1.22E-05

Rate – 2.00 2.00 2.00 2.00

‖∇C −∇Ch‖0,0 6.75E-02 3.36E-02 1.67E-02 8.34E-03 4.17E-03

Rate – 1.01 1.01 1.00 1.00

‖C − Ch‖0,1 6.76E-02 3.36E-02 1.67E-02 8.34E-03 4.17E-03

Rate – 1.01 1.01 1.00 1.00

h

|||C − Ch|||∞,0 1.66E-02 9.07E-03 4.71E-03 2.40E-03 1.21E-03

Rate – 0.88 0.94 0.97 0.99

‖C − Ch‖0,0 9.90E-03 5.34E-03 2.76E-03 1.40E-03 7.06E-04

Rate – 0.89 0.95 0.98 0.99

‖∇C −∇Ch‖0,0 9.56E-02 5.04E-02 2.58E-02 1.30E-02 6.55E-03

Rate – 0.92 0.97 0.98 0.99

‖C − Ch‖0,1 9.61E-02 5.07E-02 2.59E-02 1.31E-02 6.59E-03

Rate – 0.92 0.97 0.98 0.99

h2

|||C − Ch|||∞,0 5.91E-03 1.53E-03 3.85E-04 9.65E-05 2.41E-05

Rate – 1.95 1.99 2.00 2.00

‖C − Ch‖0,0 3.47E-03 8.86E-04 2.22E-04 5.55E-05 1.39E-05

Rate – 1.97 2.00 2.00 2.00

‖∇C −∇Ch‖0,0 6.79E-02 3.36E-02 1.67E-02 8.34E-03 4.17E-03

Rate – 1.01 1.01 1.00 1.00

‖C − Ch‖0,1 6.80E-02 3.37E-02 1.67E-02 8.34E-03 4.17E-03

Rate – 1.02 1.01 1.00 1.00

Table 8.3: Approximation errors and experimental convergence rates for the fully-implicitapproximation with a linear adsorption model. As ∆t and h are cut in half, the H2 erroris reduced the same amount which is consistent with the theoretical convergence rates.

8.3.1 Time-Integrated Discretization

For the nonlinear time-integrated a-priori error rates, we use Langmuir’s isotherm

with qmax = Keq = 1 where

q (C) =qmaxKeqC

1 +KeqC=

C

1 + C.

185

Page 197: Modeling, Analysis, and Simulation of Adsorption in ...

We simplify the problem formulation to a single (nonlinear) transport equation in

one unknown Cn using

∂q

∂t=

dq

dC· ∂C∂t

=1

(1 + C)2 ·∂C

∂t.

We compute solutions by lagging the nonlinearity q′(C) as [125]

1

(1 + Cn)2 ·Cn − Cn−1

∆t≈ 1

(1 + Cn−1)2 ·Cn − Cn−1

∆t.

Although the error analysis of the time-integrated discretization has yet to be com-

pleted, we expect from other time-integrated analysis that the convergence rates using linear

elements to be ∥∥∥∥∥∫ T

0C dt−

N∑n=0

Cnh∆t

∥∥∥∥∥0

≈ O ((1 + δ) (h+ ∆t)) .

We present numerical results in Table 8.4 that are consistent with this expectation.

δ ↓ (∆t, h) →(

132 ,

18

) (164 ,

116

) (1

128 ,132

) (1

256 ,164

) (1

512 ,1

128

)0

∥∥∥∥∥∫ T

0Cdt−

N∑n=0

Cnh∆t

∥∥∥∥∥0

∣∣∣∣∣00

2.573e-03 1.123e-03 5.206e-04 2.499e-04 1.223e-04

Rate – 1.20 1.11 1.06 1.03

h

∥∥∥∥∥∫ T

0Cdt−

N∑n=0

Cnh∆t

∥∥∥∥∥0

∣∣∣∣∣00

8.767e-03 4.799E-03 2.506e-03 1.281e-03 6.46860e-04

Rate – 0.87 0.94 0.97 0.98

Table 8.4: Time-integrated approximation errors and convergence rates for nonlinear ad-sorption with Langmuir adsorption model. As ∆t and h are cut in half, the time-integratederror is reduced the same amount which is consistent with the theoretical convergence rates.

8.3.2 Backward Euler Discretization

For the nonlinear a-priori convergence rates of the backward Euler discretization, we

again use Langmuir’s isotherm with qmax = Keq = 1 and we compute solutions by lagging

the nonlinearity q′(C) as [125].

Although the error analysis of this discretization has yet to be completed, we expect

186

Page 198: Modeling, Analysis, and Simulation of Adsorption in ...

the convergence rates using linear elements to be

‖C − Ch‖0,1 ≈ O ((1 + δ)(h+ ∆t)) .

We present numerical results in Table 8.5 that are consistent with this expectation.

δ ↓ (∆t, h) →(

132 ,

18

) (164 ,

116

) (1

128 ,132

) (1

256 ,164

) (1

512 ,1

128

)

0

|||C − Ch|||∞,0 1.40E-03 3.56E-04 8.92E-05 2.23E-05 5.58E-06

Rate – 1.98 2.00 2.00 2.00

‖C − Ch‖0,0 8.23E-04 2.06E-04 5.13E-05 1.28E-05 3.19E-06

Rate – 2.00 2.00 2.00 2.00

‖∇C −∇Ch‖0,0 1.78E-02 8.87E-03 4.42E-03 2.20E-03 1.10E-03

Rate – 1.01 1.01 1.00 1.00

‖C − Ch‖0,1 1.78E-02 8.87E-03 4.42E-03 2.20E-03 1.10E-03

Rate – 1.01 1.01 1.00 1.00

h

|||C − Ch|||∞,0 2.89E-03 1.48E-03 7.53E-04 3.81E-04 1.92E-04

Rate – 0.97 0.97 0.98 0.99

‖C − Ch‖0,0 1.73E-03 8.78E-04 4.46E-04 2.25E-04 1.13E-04

Rate – 0.98 0.98 0.99 0.99

‖∇C −∇Ch‖0,0 2.09E-02 1.08E-02 5.45E-03 2.74E-03 1.37E-03

Rate – 0.96 0.98 0.99 1.00

‖C − Ch‖0,1 2.10E-02 1.08E-02 5.47E-03 2.75E-03 1.38E-03

Rate – 0.96 0.98 0.99 1.00

h2

|||C − Ch|||∞,0 1.47E-03 3.76E-04 9.47E-05 2.37E-05 5.93E-06

Rate – 1.96 1.99 2.00 2.00

‖C − Ch‖0,0 8.62E-04 2.18E-04 5.46E-05 1.36E-05 3.41E-06

Rate – 1.98 2.00 2.00 2.00

‖∇C −∇Ch‖0,0 1.79E-02 8.87E-03 4.42E-03 2.20E-03 1.10E-03

Rate – 1.01 1.01 1.00 1.00

‖C − Ch‖0,1 1.79E-02 8.88E-03 4.42E-03 2.20E-03 1.10E-03

Rate – 1.01 1.01 1.00 1.00

Table 8.5: Approximation errors and convergence rates for nonlinear adsorption with Lang-muir adsorption model. As ∆t and h are cut in half, the H1 error is reduced the sameamount which is consistent with the theoretical convergence rates.

187

Page 199: Modeling, Analysis, and Simulation of Adsorption in ...

8.3.3 Two Unknowns Discretization

To verify the nonlinear a-priori convergence rates for the two unknowns discretiza-

tion, we use Nor’s isotherm with

Keq = Λ = z = cs = cm = qmax = 1, ν = n = 0.5, and Ks = Kp = −1

so that

q = (1− q) exp−1− C

and consequently

q =C exp−1− C

1 + C exp−1− C.

Although the error analysis of the discretization with two unknowns has yet to be

completed, we expect the convergence rates using linear elements to be

‖C − Ch‖0,1 ≈ O ((1 + δ)(h+ ∆t))

and

‖q − qh‖0,0 ≈ O ((1 + δ)(h+ ∆t)) .

We present numerical results in Table 8.6 that are consistent with this expectation.

8.4 Numerical Simulations

For the comparison with data generated in the laboratory, we choose parameter

values to mimic the associated laboratory conditions. The data was obtained by Juan

Wang in the Department of Chemical and Biomolecular Engineering at Clemson University

as part of Scott Husson’s Bioseparations and Advanced Separation Materials research group.

The multimodal membrane (MMM) is modified from commercially available regener-

ated cellulose membrane. The regenerated cellulose membranes purchased from Whatman,

188

Page 200: Modeling, Analysis, and Simulation of Adsorption in ...

δ ↓ (∆t, h) →(

132 ,

18

) (164 ,

116

) (1

128 ,132

) (1

256 ,164

) (1

512 ,1

128

)0

|||C − Ch|||∞,0 1.40E-03 3.56E-04 8.92E-05 2.23E-05 5.58E-06

Rate – 1.98 2.00 2.00 2.00

‖C − Ch‖0,0 8.23E-04 2.06E-04 5.13E-05 1.28E-05 3.19E-06

Rate – 2.00 2.00 2.00 2.00

‖∇C −∇Ch‖0,0 1.78E-02 8.87E-03 4.42E-03 2.20E-03 1.10E-03

Rate – 1.01 1.01 1.00 1.00

‖C − Ch‖0,1 1.78E-02 8.87E-03 4.42E-03 2.20E-03 1.10E-03

Rate – 1.01 1.01 1.00 1.00

‖q − qh‖∞,0 7.84E-04 2.79E-04 1.11E-04 4.85E-05 2.39E-05

Rate – 1.49 1.33 1.20 1.02

‖q − qh‖0,0 5.91E-04 2.35E-04 1.03E-04 4.81E-05 2.32E-05

Rate – 1.33 1.19 1.10 1.05

h

|||C − Ch|||∞,0 2.89E-03 1.48E-03 7.53E-04 3.81E-04 1.92E-04

Rate – 0.97 0.97 0.98 0.99

‖C − Ch‖0,0 1.73E-03 8.77E-04 4.46E-04 2.25E-04 1.13E-04

Rate – 0.98 0.98 0.99 0.99

‖∇C −∇Ch‖0,0 2.09E-02 1.08E-02 5.45E-03 2.74E-03 1.37E-03

Rate – 0.96 0.98 0.99 1.00

‖C − Ch‖0,1 2.10E-02 1.08E-02 5.47E-03 2.75E-03 1.38E-03

Rate – 0.96 0.98 0.99 1.00

‖q − qh‖∞,0 1.17E-03 5.73E-04 2.86E-04 1.43E-04 7.14E-05

Rate – 1.03 1.00 1.00 1.00

‖q − qh‖0,0 7.97E-04 3.89E-04 1.93E-04 9.61E-05 4.80E-05

Rate – 1.03 1.01 1.01 1.00

h2

|||C − Ch|||∞,0 1.47E-03 3.77E-04 9.47E-05 2.37E-05 5.93E-06

Rate – 1.96 1.99 2.00 2.00

‖C − Ch‖0,0 8.62E-04 2.18E-04 5.46E-05 1.36E-05 3.41E-06

Rate – 1.98 2.00 2.00 2.00

‖∇C −∇Ch‖0,0 1.79E-02 8.87E-03 4.42E-03 2.20E-03 1.10E-03

Rate – 1.01 1.01 1.00 1.00

‖C − Ch‖0,1 1.79E-02 8.88E-03 4.42E-03 2.20E-03 1.10E-03

Rate – 1.01 1.01 1.00 1.00

‖q − qh‖∞,0 8.03E-04 2.85E-04 1.13E-04 4.89E-05 2.39E-05

Rate – 1.49 1.34 1.20 1.03

‖q − qh‖0,0 6.01E-04 2.39E-04 1.04E-04 4.83E-05 2.32E-05

Rate – 1.33 1.20 1.11 1.05

Table 8.6: Approximation errors and convergence rates for nonlinear adsorption with Nforadsorption model. As ∆t and h are cut in half, the H1 error for C and the L2 error for qis reduced the same amount which is consistent with the theoretical convergence rates.

189

Page 201: Modeling, Analysis, and Simulation of Adsorption in ...

Inc. have an average effective pore size of 1 µm, 0.7 mm thickness and 47 mm diameter. A

stack of 10 MMMs is placed in a Mustang Coin® module (Pall Corporation, Port Washing-

ton, NY) with one piece of 25 µm nominal pore diameter filter paper (Whatman 5) placed

on each side of the stack. The resulting height of the module is approximately 0.7 cm.

The effective membrane diameter within the module is 1.4 cm; however, a diameter of 1.6

cm is used for calculation of the membrane bed volume and consequently in computational

simulations because radial distribution of the adsorbing species within the membrane stack

is likely to happen. The module is installed in an AKTA purifier. The membrane bed

porosity is directly measured to be 0.84.

A protein-rich serum is pushed vertically through the membrane using a constant

pressure differential providing a constant flow rate of 0.1 mL/min. The vertical velocity is

thus chosen to have a parabolic velocity profile with zero velocity at the sides of the module

and maximum velocity chosen to ensure the correct flow rate. We assume a 0 cm/min

horizontal velocity.

Values for the diffusion of protein molecules in porous membranes range from 6 ×

10−4 cm2/min (for small proteins such as glycine) to 2.28×10−5 cm2/min (for large proteins

such as immunoglobulin G) [154]. We choose a diffusion value of 2.28 × 10−5 cm2/min as

immunoglobulin G was used in the experiment.

Since the experimental velocity is larger than this diffusion, we use a more general

model for the dispersion tensor, D, than simply diffusion, d0. We choose a model for low

Peclet numbers [50] to account for the spreading effect due to both diffusion and velocity:

DL = ωd0 + αL |u| , DT = ωd0 + αT |u|

where ω is the porosity, d0 is the coefficient of diffusion, |u| is the mean microscopic ve-

locity, and αL and αT are the longitudinal and transverse dispersiveness respectively. The

190

Page 202: Modeling, Analysis, and Simulation of Adsorption in ...

dispersion tensor is then expressed in its principal direction of anisotropy giving

D =

DL 0

0 DT

with DL the longitudinal dispersion coefficient (in the direction of the flow) and DT the

transverse dispersion coefficient (in the direction at a right angle to the velocity). In our

case the velocity is purely vertical so that the longitudinal direction is vertical and the

transverse direction is horizontal.

The values for the porosity and the diffusion coefficient have been discussed above.

The value for the mean microscopic velocity |u| is approximated by means of the constant

flow rate. As the experimental flow rate is 0.1 mL/min, then

|u| ≈ flowrate

πr2= 0.0497 cm/min.

The values for the longitudinal and transverse dispersivities, αL and αT respectively,

are known to vary depending on the physical properties of the membrane, but they cannot

be obtained experimentally. However, much is known about their behavior from previous

experimentation [50, 2, 172]. Experimental values for αL on a small scale range from

around a few millimeters to nearly a hundred centimeters depending on the porosity [50, 2],

and αL decreases as the porosity increases [172]. Consequently for our purposes with a

porosity value closer to 1, αL should be on the lower end of this range on the order of a few

millimeters. The specific value of αL is chosen to match up the experimental and numerical

points of breakthrough.

Additionally, αT is generally found to be between 15 and 1

100 of αL [50]. Through

comparison of the experimental and numerical breakthrough curves, larger values of αL give

more accurate numerical results. We therefore take αT = 15αL.

We consider two isotherm models for the numerical simulation results shown below.

191

Page 203: Modeling, Analysis, and Simulation of Adsorption in ...

First we consider Nfor’s model

q = KeqΛν+n

(1

zcs

)ν ( 1

cm

)n(1− q

qmax

)ν+n

C exp Kscs +KpC

to capture the multimodal adsorption properties of the MMMs. The parameter values were

obtained by fitting the model to known data and are as follows:

Keq = 502, Λ = 0.34009 M, ν = 0.1, n = 1.0, z = 1, cs = 0.37 M,

cm = 55.5 M, qmax = 216 mg/mL, Ks = −0.1 M−1, and Kp = −0.1 mL/mg.

where M denotes a Molar unit (=moles/L). Note that an absence of units implies the

quantity is unitless. For the simulations involving Nfor’s isotherm, we take h and ∆t to

be small enough to ensure the error is not significantly affecting the results; specifically we

choose h = 0.0125 and ∆t = 0.0078125.

As a comparison, we also consider Langmuir’s model

q (C) =qmaxKeqC

1 +KeqC

where qmax and Keq were obtained by fitting the model to known data; specifically,

qmax = 150 mg/mL and Keq = 2.06 mL/mg.

For the simulations involving Langmuir’s isotherm, we again choose h and ∆t to be small

enough to ensure the error is not significantly affecting the results; specifically we choose

h = 0.0125 and ∆t = 0.03125.

The metric of productivity is important in evaluating bind-and-elute chromatog-

raphy processes. Previously [33], Husson and coworkers defined productivity as the mass

of protein that can bind per volume of membrane per time, and provided a convenient

192

Page 204: Modeling, Analysis, and Simulation of Adsorption in ...

expression to calculate it:

Productivity =Bdynamic

tbreak

where Bdynamic is the dynamic binding capacity and tbreak is the time it takes to reach the

point of 10% breakthrough, i.e., the time at which 10% of the value of the feed concentration

appears in the effluent from the column. From a productivity standpoint, differences in

dynamic binding capacity do not impact productivity nearly as significantly as differences

in residence times [33]. Therefore, high productivities can be achieved when the time to

breakthrough is short; that is, the membrane quickly reaches its capacity to bind proteins.

The information on protein breakthrough is visualized using a breakthrough curve,

which is a plot of the concentration at the effluent as a function of accumulated effluent

volume. We choose to assess the validity of our approach by comparing our simulated break-

through curve with the experimental curve generated from laboratory data. Breakthrough

curves are greatly affected by diffusion and dispersion values. Consequently, having accu-

rate diffusion and dispersion values is important when analyzing breakthrough curves. Since

upwinding adds some artificial diffusion along the velocity streamlines, we take the upwind-

ing parameter δ to be 0 to ensure there is no artificial diffusion affecting the breakthrough

results.

Figure 8.1 shows experimental results along with results generated using the disper-

sion model described above. The initial baseline absorbance reading indicates that protein

is binding to the column. As the membrane bed capacity is reached under flow conditions,

the unbound protein begins to break through the column, illustrated by the initial increase

in concentration. Both experimental and model breakthrough curves exhibit the typical

S-shape characteristic of favorable binding within a packed bed. Experimentally, break-

through occurs after approximately 1.366 mL of fluid has passed through the column which

is matched very closely by the model.

The model shows complete bed saturation, i.e., the effluent concentration reaches the

feed value, after approximately 3 mL of fluid has passed through the column; whereas, the

193

Page 205: Modeling, Analysis, and Simulation of Adsorption in ...

experimental breakthrough curve does not reach saturation during the measurement period.

Such behavior for the experimental data is typical for the membranes used in this study.

At high degrees of loading, the adsorption rate decelerates due to surface exclusion from

previously adsorbed molecules [155]. Since the model does not account for the kinetic rate

of adsorption, it is not surprising that it does not capture this long-term tailing behavior.

Figure 8.1: Comparison of numerical and experimental breakthrough curves using instan-taneous adsorption. Note the typical S-shape characteristic and the close match of break-through for both curves.

Although it is standard practice to stop flow at 10% breakthrough, Nadarajah and

Mehta [117] proposed an overload and elute purification strategy for purification of pro-

tein antibodies to be able to utilize bed capacity more fully. Also, recently Husson and

coworkers found that the kinetic rate of IgG binding with multimodal ligands becomes the

rate limiting step at high process flow rates for our first-generation multimodal membranes

[167]. Therefore, extension of the model to incorporate transient kinetic adsorption rates is

of interest.

Given the accuracy of the numerical results obtained with Langmuir’s model in

194

Page 206: Modeling, Analysis, and Simulation of Adsorption in ...

this case, one might question why Nfor’s model should be considered given that Langmuir’s

model is simpler. Recall from the discussion of Nfor’s model in Chapter 1 that the constants

in the Nfor model are independent of salt concentration, which is not the case for the

Langmuir model. Thus, we could predict breakthrough at different salt concentrations

with the same model parameters for Nfor; whereas we would need to reevaluate model

parameters for Langmuir anytime salt concentration is varied. Put simply, Nfor is predictive

and Langmuir is not.

8.4.1 Non-Instantaneous Adsorption

To incorporate non-instantaneous, or non-equilibrium, adsorption, we consider a

system of three equations:

ω∂C

∂t+ (1− ω)ρs

∂q

∂t+∇ · (uC)−∇ · (D∇C) = f (8.4.1)

∂q

∂t= km(q∗ − q) (8.4.2)

q = g(q, C) (8.4.3)

where km is the adsorption reaction rate and q∗ is an equilibrium concentration, i.e. the

concentration at the interface between the concentration in the mobile phase and the con-

centration in the adsorbed phase. The same initial and boundary conditions are applied to

C and q with an additional initial condition on q∗, namely q∗ = 0.

Applying the finite element method and linearizing (8.4.3) similar to the system

with two unknowns by evaluating g at the previous time step gives the following variational

195

Page 207: Modeling, Analysis, and Simulation of Adsorption in ...

formulation: For n = 1, 2, ..., N , find (C, q∗, q) ∈ V ×W ×W such that

(ωCn − Cn−1

∆t, v

)+

((1− ω)ρs

qn − qn−1

∆t, v

)+ (∇ · (uCn), v)− (∇ · (D∇Cn), v) = (f, v) ,(

qn − qn−1

∆t, w

)= (km((q∗)n − qn), w) ,

(qn, z) =(g(qn−1, Cn−1), z

).

for all v ∈ V , w ∈W , z ∈W .

Figure 8.2 shows results using non-instantaneous adsorption with h and ∆t chosen

to be 0.0125 and 0.0078125 respectively; as with the simulations involving instantaneous

adsorption, the upwinding parameter was chosen to be δ = 0. Breakthrough curves resulting

from various constant reaction rates are compared. Although km = 0 corresponds to no

adsorption, the results for that case are provided as a point of comparison. Recall that

as km → ∞, the behavior of the system (8.4.1)-(8.4.3) approaches that of instantaneous

adsorption. The results in Figure 8.2 support this as there is no distinguishable difference

between the breakthrough curves for km ≥ 10. In fact, most of the variation in the results

occurs between km = 0 and km = 1. For lower values of km, the shape of the breakthrough

curves changes in addition to the 10% breakthrough point changing quite drastically. For

0 ≤ km ≤ 1, the 10% breakthrough point ranges from 0.915 to 1.299. The effects of a

finite reaction rate, as shown in Figure 8.2, support the supposition that incorporating

non-instantaneous adsorption would provide more accurate numerical results.

Additionally, the influence of reaction rate becomes important at higher flow rates,

i.e. shorter residence times. This can be related to the first Damkohler number, which is

a dimensionless quantity representing the ratio of the reaction rate to the rate of convec-

tive mass transport through the membrane bed. Husson and coworkers discuss the first

Damkohler number and the influence of reaction rate with varying flow rates in [32].

It would stand to reason that the rate of adsorption would not stay constant through

196

Page 208: Modeling, Analysis, and Simulation of Adsorption in ...

Figure 8.2: Effect of varying the mass transfer coefficient km on breakthrough curve. No-tice as km increases, the breakthrough curves asymptotically approach the results withinstantaneous adsorption as shown in Figure 8.1.

the entire process. In fact, the reaction rate most likely decreases with time as the mem-

branes become saturated with protein. It may therefore be beneficial to consider a varying

adsorption rate. Having a reaction rate dependent on the amount of concentration in the

adsorbed phase would be most accurate, but for simplicity in the algorithm, considering

a reaction rate that is solely a function of time way be sufficient. Figure 8.3 shows the

breakthrough curve resulting from one such function as compared to the results shown in

Figure 8.2.

The reaction rate function for the results in Figure 8.3 was chosen to have a large

value for small t that decreases with time in order to give results closer to life; specifically,

we have

km(t) =K

(t− 1)p

where K and p are chosen to match up the 10% breakthrough point and the shape of the

breakthrough curve. In this case, with all other parameters chosen as in the Numerical

197

Page 209: Modeling, Analysis, and Simulation of Adsorption in ...

Figure 8.3: Comparison of transient adsorption rate with constant adsorption rates.

Simulations section above, we pick

K = 10(13.66)p and p = 10.

The values in the numerator, 10 and 13.66, were chosen to match up 10% breakthrough

point, while the power of 10 was chosen to decrease the adsorption faster to better match

up the breakthrough curve. A comparison of the numerical and experimental results, as

shown in Figure 8.4, confirms that the results with a transient adsorption rate are in fact

closer to the experimental results.

8.4.2 Varying Velocity Profiles

In order to investigate the effects of varying velocity profiles on the breakthrough

curves, we ran two- and three-dimensional simulations involving five different velocity pro-

files. With the exception of the simulation involving spatially-constant velocity, the velocity

profiles were parabolic in nature with between 1 and 5 vertices of maximum magnitude.

198

Page 210: Modeling, Analysis, and Simulation of Adsorption in ...

Figure 8.4: Comparison of numerical and experimental breakthrough curves using instan-taneous adsorption and non-instantaneous adsorption. Note the numerical result obtainedusing a transient adsorption rate are significantly closer to the experimental results thanthose obtained using instantaneous adsorption.

More detailed descriptions of the five cases are below.

In 2D, there were three velocity profiles considered:

1. Spatially-constant velocity with a value chosen to ensure a flow rate of 0.1 mL/min.

2. Parabolic velocity with the vertex at the center of the inflow boundary, zero values

on side boundaries, and the value at the vertex chosen to ensure a flow rate of 0.1

mL/min.

3. Velocity profile with two parabolas having zero value at the center of the inflow bound-

ary, vertices at ±√

22 r, and values at the vertices chosen to ensure a flow rate of 0.1

mL/min.

In 3D, there were two velocity profiles considered:

1. Paraboloid velocity profile with the vertex at the center of the inflow boundary, zero

199

Page 211: Modeling, Analysis, and Simulation of Adsorption in ...

value on the entire circular side boundary, and a vertex value chosen to ensure a flow

rate of 0.1 mL/min.

2. Velocity profile with five paraboloids having the vertices placed in an X design (1 at

the center of the inflow boundary and the other 4 all 23r distance away from the center

at 45, 135, 225, and 315), all zero values on and outside circles of radius 13r from

the vertices, and values at all five vertices chosen as the same value to ensure a flow

rate of 0.1 mL/min.

All the results in this section were obtained using deal.II with temporal discretization

value being ∆t = 0.03125 and a final simulation time of T = 80. The spatial discretization

value varied somewhat depending on the velocity profile and consequently is described in

each subsection below. The upwinding parameter δ was chosen to be 0 in each case to

ensure there is no artificial diffusion affecting the breakthrough results.

The results for all five cases assumed instantaneous adsorption using a Langmuir

isotherm model with qmax = 150 mg/mL and Keq = 2.06 mL/mg. Also, a low Peclet

dispersion model was used with d0 = 2.28 × 10−5 cm2/min, |u| = 0.0497 cm/min, αL =

0.001, and αT = 1/10 ∗ αL = 0.0001. The porosity and dimensions of the membrane were

taken to be the same as simulation values described in previous sections: width= 0.7 cm,

height= 1.6 cm, ω = 0.84. The inflow concentration was also the same: Cin = 3 mg/mL.

8.4.2.1 2D Constant Velocity Results

For the case of a two-dimensional, spatially-constant velocity profile, six initial spa-

tial refinements in deal.II were used resulting in a spatial discretization parameter of ap-

proximately 0.02, i.e. h ≈ 0.02. This value is much smaller than what is necessary to

get accurate breakthrough curve results considering the values chosen for the dispersion;

however, h was chosen to be this small to ensure that the concentration front as shown in

Figure 8.5 were fully resolved.

The evolution of the concentration front in the membrane is shown in Figure 8.5.

200

Page 212: Modeling, Analysis, and Simulation of Adsorption in ...

The shape of the concentration front follows the shape of the velocity profile (spatially

constant over the width of the membrane), and the front progressively moves through the

height of the membrane as time elapses. The membrane is very nearly saturated by t = 20

as shown in the last image in Figure 8.5.

8.4.2.2 2D Single Parabola Velocity Results

For the case of a two-dimensional, parabolic velocity profile, six initial spatial re-

finements were used again to give h ≈ 0.02. The evolution of the concentration front in

the membrane is shown in Figure 8.6. The shape of the concentration front follows the

shape of the velocity profile (parabolic over the width of the membrane), and the front

progressively moves through the height of the membrane as time elapses. The saturation

of the membrane occurs at a later time when compared to the case of a constant velocity

profile; in this case, saturation occurs at approximately t = 70 which is later than the last

image shown in Figure 8.6.

8.4.2.3 2D Double Parabola Velocity Results

For the case of a two-dimensional velocity profile involving two parabolas, we again

chose h ≈ 0.02. The evolution of the concentration front in the membrane is shown in

Figure 8.7. The shape of the concentration front follows the shape of the velocity profile

(two parabolas over the width of the membrane), and the front progressively moves through

the height of the membrane as time elapses. The saturation of the membrane occurs later

than in the case of constant velocity but earlier than the case of a single parabolic velocity

profile. In this case, saturation occurs at approximately t = 65 which is later than the last

image shown in Figure 8.7.

8.4.2.4 3D Single Parabola Velocity Results

For the case of a three-dimensional, single parabolic velocity profile, four initial

spatial refinements in deal.II were used resulting in a spatial discretization parameter of

201

Page 213: Modeling, Analysis, and Simulation of Adsorption in ...

approximately 0.08, i.e. h ≈ 0.08. Although this value is larger than that chosen for the

two-dimensional cases, it is still small enough to give accurate breakthrough curves results

and resolve the concentration profiles well enough for visual inspection. A smaller value was

not used because of time and computational constraints (i.e. a value of h ≈ 0.08 resulted in

a run time of about 9 days on a Dell Precision M4500 with an Intel Core i7 X940 @2.13GHz

and 8 GB of RAM running in Ubuntu 14.04 LTS).

The evolution of the concentration front in the membrane is shown in Figure 8.8. As

with the two-dimensional cases, the shape of the concentration front follows the shape of the

velocity profile, and the front progressively moves through the height of the membrane as

time elapses. The saturation of the membrane occurs slightly later than the two-dimensional

case with a single parabolic velocity profile; specifically, saturation occurs at approximately

t = 72.5 which is later than the last image shown in Figure 8.8.

8.4.2.5 3D Quintuple Parabola Velocity Results

For the case of a three-dimensional velocity profile involving five parabolas, three

initial spatial refinements in deal.II were used resulting in a spatial discretization parameter

of approximately 0.17, i.e. h ≈ 0.17. Although this value is larger than that chosen for the

other three-dimensional cases, it is still small enough to give accurate breakthrough curves

results. The concentration profiles could be resolved somewhat better especially on the

outflow boundary, but they are resolved enough to see the general behavior of the profile.

The larger value of h was chosen to reduce the computational time of the simulation; in this

case the run time was about 25 hours on the same machine (Dell Precision M4500 with an

Intel Core i7 X940 @2.13GHz and 8 GB of RAM running in Ubuntu 14.04 LTS).

The evolution of the concentration front in the membrane is shown in Figure 8.9.

As with all previous cases, the shape of the concentration front follows the shape of the

velocity profile, and the front progressively moves through the height of the membrane as

time elapses. The saturation of the membrane occurs later than in any of the other cases.

In this case, saturation occurs at approximately t = 80 which is at the end of the simulation

202

Page 214: Modeling, Analysis, and Simulation of Adsorption in ...

and occurs later than the last image shown in Figure 8.9.

8.4.2.6 Comparison of Breakthrough Curves

As the purpose of running these five simulations was to see the effects of varying

velocity profiles on breakthrough, a comparison of the breakthrough curves was generated

and is shown in Figure 8.10. The cases of a single parabolic velocity profile and a double

parabolic velocity profile resulted in very similar breakthrough curves while the other three

cases are distinctly different. Even with two of the two-dimensional cases being similar, there

is enough variation in the breakthrough curve results to indicate that the velocity profile

can drastically affect the breakthrough results, and consequently a further investigation into

velocity profiles, perhaps incorporating variable velocity, should be conducted.

We note here the difference between the 2D and 3D single parabolic velocity profile

results. This difference can most likely be contributed to how the outflow concentration is

computed in the two-dimensional case. For both the 2D and 3D cases, the concentration is

averaged over the outflow; this results in the average being taken over width of the outflow

in the 2D case instead of the area of the outflow in the 3D case. Another reason for the

differences in the 2D and 3D single parabolic cases could be the fact that the 2D simulations

are taken as a slice through the center of the membrane which may not accurately take into

account the cylindrical nature of the membrane in the chromatography column.

203

Page 215: Modeling, Analysis, and Simulation of Adsorption in ...

(a) t = 0.03125 (b) t = 6.5

(c) t = 13 (d) t = 20

Figure 8.5: The 2D evolution of the concentration in the membrane assuming a spatially-constant velocity profile. Notice that the membrane is essentially saturated with protein byt = 20 as shown in the last image.

204

Page 216: Modeling, Analysis, and Simulation of Adsorption in ...

(a) t = 0.03125 (b) t = 7.5

(c) t = 15 (d) t = 23

Figure 8.6: The 2D evolution of the concentration in the membrane assuming a singleparabola velocity profile. Saturation of the membrane does not occur in this case until afterthe final image shown; specifically, it occurs at approximately t = 70.

205

Page 217: Modeling, Analysis, and Simulation of Adsorption in ...

(a) t = 0.03125 (b) t = 7.5

(c) t = 15 (d) t = 23

Figure 8.7: The 2D evolution of the concentration in the membrane assuming a doubleparabola velocity profile. Saturation of the membrane does not occur in this case until afterthe final image shown; specifically, it occurs at approximately t = 65.

206

Page 218: Modeling, Analysis, and Simulation of Adsorption in ...

(a) t = 0.03125 (b) t = 7.5

(c) t = 15 (d) t = 23

Figure 8.8: The 3D evolution of the concentration in the membrane assuming a singleparabola velocity profile. Saturation of the membrane does not occur in this case until afterthe final image shown; specifically, it occurs at approximately t = 72.5.

207

Page 219: Modeling, Analysis, and Simulation of Adsorption in ...

(a) t = 0.125 (b) t = 1.875

(c) t = 3.75 (d) t = 5.75

Figure 8.9: The 3D evolution of the concentration in the membrane assuming a velocityprofile involving five parabolas. Saturation of the membrane does not occur in this caseuntil after the final image shown; specifically, it occurs at approximately t = 80.

208

Page 220: Modeling, Analysis, and Simulation of Adsorption in ...

Figure 8.10: Comparison of breakthrough curves for five different velocity profiles. Althoughsome of the velocity profiles resulted in remarkably similar breakthrough curves (e.g. singleand double parabola cases in 2D), there is enough variation in the curves to indicate furtherinvestigation into the velocity if warranted.

209

Page 221: Modeling, Analysis, and Simulation of Adsorption in ...

Chapter 9

Summary and Future Work

We conclude with a brief summary of work completed and possible future extensions.

9.1 Summary

In this work, we considered the time-dependent, multi-dimensional, advection-

diffusion-reaction equation used to model reactive transport in porous media. Specifically,

we focused on modeling the mechanism of protein chromatography as part of a protein sep-

arations process with the goal of developing software tools capable of simulating the protein

chromatography process under the effect of complex, implicit adsorption relationships in

the presence of highly advective flows.

With this goal in mind, we developed and analyzed various streamline upwind

Petrov-Galerkin (SUPG) finite element discretizations of the reactive transport problem.

Beginning with the steady-state equation of the SUPG formulation, we showed the bilinear

form was coercive and bounded in the H1-norm; in addition, we derived H1 error bounds

for the steady-state equation. The error bound was derived with an explicit dependence on

the upwinding parameter δ.

Continuing with the transient problem, we analyzed the cases of constant and lin-

ear adsorption. We showed solvability of the fully discrete (both Forward and Backward

210

Page 222: Modeling, Analysis, and Simulation of Adsorption in ...

Euler) SUPG formulations and derived L∞(L2) stability and L2(H1) error bounds for the

semi-discrete and fully-discrete problems. All the stability and error bounds for the linear

formulations were shown with an explicit dependence on the upwinding parameter δ.

As the implicit adsorption relationship was nonlinear, we also developed and an-

alyzed finite element formulations involving a nonlinear adsorption term. In the case of

an explicit adsorption relationship, three different formulations were analyzed: a time-

integrated mixed methods formulation, a time-integrated SUPG formulation, and a fully

implicit SUPG formulation.

For the time-integrated mixed methods formulation, error estimates for the semi-

discrete and fully discrete cases were proven in terms of a “time-integrated” error:

∥∥∥∥∫ T

0(C − Ch) dt

∥∥∥∥2

0

+

∥∥∥∥∫ T

0(Q−Qh) dt

∥∥∥∥2

0

.

Solvability of the time-integrated SUPG formulation was also shown. For the fully implicit

SUPG formulation, the solvability was proven with the use of the Newton-Kantarovich

theorem to deal with the nonlinearity arising from the fully implicit discretization; the

stability was proven in terms of the L2 spatial norm evaluated at the final time. The stability

bounds for both SUPG formulations were written in terms of the upwinding parameter δ.

For the implicit adsorption relationship, a simple formulation was proposed which

not only dealt with the implicit definition of the isotherm but also dealt with the nonlinear-

ity: the right hand side of the isotherm relationship was evaluated at the previous time step.

As expected, the solvability and stability bound for this relationship were shown to have a

requirement on the time step size based on the Lipschitz continuity of the right hand side

of the isotherm. While this time step requirement is unfortunate, it is not an unreasonable

restriction in real world simulations. As with the previous nonlinear stability bounds, the

stability bound for the case of implicit adsorption was explicitly written in terms of the

upwinding parameter δ.

We also showed results from various numerical experiments. The first three ex-

211

Page 223: Modeling, Analysis, and Simulation of Adsorption in ...

perimental results supported the steady-state and fully-discrete linear convergence rates

while the next three experiments showed the expected, although yet unproven, nonlinear

convergence rates including one using the implicit adsorption model proposed by Nfor.

Results from simulations were also provided. The comparison of numerical results

using instantaneous adsorption with experimental results (Figure 8.1) showed fairly good

breakthrough curve predictions from both Langmuir’s adsorption model and Nfor’s adsorp-

tion model; the good predictions were most likely due to the addition of the low Peclet

dispersion model which allowed some freedom in the choice of dispersivities. However,

while the 10% breakthrough point was matched by appropriate choices of the dispersivities,

the shape of the breakthrough curve did not match up as well later on.

Consequently, we also investigated the affects of non-instantaneous adsorption in the

hopes of obtaining better matching results. Incorporating non-instantaneous adsorption,

particularly that involving a transient adsorption rate, greatly improved the shape of the

numerical breakthrough curve in the case of Nfor’s adsorption model significantly.

Last, we investigated the effects of varying velocity profiles on the breakthrough

curves by running two-dimensional and three-dimensional simulations involving five different

velocity profiles. Images showing the evolution of the concentration fronts for each case

were included. A comparison of the breakthrough curves for the five simulations showed

significant differences caused by the different velocity profiles.

9.2 Future Work

The analysis presented in this work considered many different finite element dis-

cretizations of the reactive transport equation; however, the error analysis of most of the

nonlinear cases has yet to be completed and therefore should be the first step in further anal-

ysis. Later analysis should include the three-equation system involving non-instantaneous

adsorption. Considering a constant adsorption rate, or similarly one dependent on time,

would result in an additional linear equation in the system being analyzed. Since the ad-

212

Page 224: Modeling, Analysis, and Simulation of Adsorption in ...

sorption rate in real world simulations most likely depends on the adsorbed concentration,

considering an adsorption rate function dependent on q in the analysis would be another

extension of the current analysis.

Numerically, more work involving non-instantaneous adsorption needs to be com-

pleted. While the numerical results incorporating a transient adsorption rate are indeed

much closer to the experimental results, improvements can still be made by varying other

parameters such as the dispersivities, αL and αT . In addition, other functions can be used

for the transient adsorption rate, e.g., a logarithmic function of time, or the adsorption rate

can be made to be a function of the adsorbed concentration, q.

Due to the initial results comparing different velocity profiles, further numerical

study should be conducted into the effects of velocity on the breakthrough curves. As part

of this study, a variable velocity should be incorporated into the model to determine if the

assumption of temporally-constant velocity is accurate.

Since there are now a number of parameters which can be varied to increase the ac-

curacy of the numerical results, finding the best set of parameters could be greatly hastened

by creating an optimization algorithm. The algorithm, posed over the set of parameters

involved in the dispersion model and the transient reaction rate function, would minimize

the difference between the experimental and numerical breakthrough curves.

The minimization problem could also be adapted to verify the measured values of

the parameters used in the governing equations. It can be posed over any subset of the

parameters, giving additional insight into their specific effects on the separation process.

The effect of individual parameters became apparent recently when a change was made to

the porosity value used in simulations. Initially the manufacturer’s porosity value was being

used, but a change was made after a different value was obtained experimentally by Juan

Wang in the Department of Chemical and Biomolecular Engineering at Clemson University.

The experimentally determine porosity value resulted in a numerical breakthrough curve

which was much closer to the experimental breakthrough curve than the results using the

manufacturer’s porosity.

213

Page 225: Modeling, Analysis, and Simulation of Adsorption in ...

Last, a numerical analysis of the different algorithms developed for the reactive

transport problem would be very informative. All the different algorithms we developed,

including a couple of discretizations not described in this work, worked successfully for

the parameter values used in the comparison experiments. However, if we wish to apply

algorithms described in this work to other applications, it would be beneficial to know which

parameter sets are feasible for each algorithm. Consequently, future work should include

an in-depth comparison of the different algorithms, including a sensitivity analysis of the

algorithms to parameter values.

214

Page 226: Modeling, Analysis, and Simulation of Adsorption in ...

Bibliography

[1] AchemAsia. New technologies beef up pharmaceutical production. http:

//www.achemasia.de/print_content.php?theHref=press/tb2_e.html&spkz=

E&displayStatus=inline, 2013.

[2] N. Agarwal, M.J. Semmens, P.J. Novak, and R.M. Hozalski. Zone of influence of agas permeable membrane system for delivery of gases to groundwater. Water Resour.Res., 41(5), 2005.

[3] D.G. Anderson. Iterative procedures for nonlinear integral equations. J. Assoc. Com-put. Mach., 12(4):547–560, 1965.

[4] T. Arbogast and M.F. Wheeler. A characteristics-mixed finite element method foradvection-dominated transport problems. SIAM J. Numer. Anal., 32(2):404–424,1995.

[5] T. Arbogast, M.F. Wheeler, and N.-Y. Zhang. A nonlinear mixed finite elementmethod for a degenerate parabolic equation arising in flow in porous media. SIAM J.Numer. Anal., 33:1669–1687, 1996.

[6] T. Arbogast, M.F. Wheeler, and N.-Y. Zhang. A nonlinear mixed finite elementmethod for a degenerate parabolic equation arising in flow in porous media. SIAM J.Numer. Anal., 33:1669–1687, 1996.

[7] D.N. Arnold and F. Brezzi. Mixed and nonconforming finite element methods: Imple-mentation, postprocessing and error estimates. RAIRO-Math. Model. Num., 19(1):7–32, 1985.

[8] R.E. Bank, J.F. Burgler, W. Fichtner, and R.K. Smith. Some upwinding techniquesfor finite element approximations of convection-diffusion equations. Numer. Math.,58:185–202, 1990.

[9] J.W. Barrett and P. Knabner. Finite element approximations of the transport ofreactive solutes in porous media. Part I: Error estimates for nonequilibrium adsorptionprocesses. SIAM J. Numer. Anal., 34(1):201–227, 1997.

[10] J.W. Barrett and P. Knabner. Finite element approximations of the transport ofreactive solutes in porous media. Part II: Error estimates for equilibrium adsorptionprocesses. SIAM J. Numer. Anal., 34(2):455–479, 1997.

215

Page 227: Modeling, Analysis, and Simulation of Adsorption in ...

[11] J. Bear. Dynamics of Fluids in Porous Media. Dover, 1988.

[12] S. Bhattacharjee, J. Dong, Y. Ma, S. Hovde, J.H. Geiger, G.L. Baker, and M.L.Bruening. Formation of high-capacity protein-adsorbing membranes through simpleadsorption of poly (acrylic acid)-containing films at low pH. Langmuir, 28(17):6885–6892, 2012.

[13] B.V. Bhut and S.M. Husson. Dramatic performance improvement of weak anion-exchange membranes for chromatographic bioseparations. J. Membrane Sci., 337:215–223, 2009.

[14] B.V. Bhut, S.R. Wickramasinghe, and S.M. Husson. Preparation of high-capacity,weak anion-exchange membranes for protein separations using surface-initiated atomtransfer radical polymerization. J. Membrane Sci., 325:176–283, 2008.

[15] P.B. Bochev, M.D. Gunzburger, and J.N. Shadid. Stability of the SUPG finite ele-ment method for transient advection-diffusion problems. Comput. Method. Appl. M.,193:2301–2323, 2004.

[16] D. Boffi, F. Brezzi, and M. Fortin. Mixed Finite Element Methods and Applications.Springer-Verlag, 2010.

[17] E.T. Bouloutas and M.A. Celia. An improved cubic Petrov-Galerkin method forsimulation of transient advection-diffusion processes in rectangularly decomposabledomains. Comput. Method. Appl. M., 42:289–308, 1991.

[18] T.H. Boyer, C.T. Miller, and P.C. Singer. Modeling the removal of dissolved organiccarbon by ion exchange in a completely mixed flow reactor. Water Res., 42:1897–1906,2008.

[19] F. Brezzi, J. Douglas, Jr., R. Duran, and M. Fortin. Mixed finite elements for secondorder elliptic problems in three variables. Numer. Math., 51:237–250, 1987.

[20] A.N. Brooks and T.J.R. Hughes. Streamline Upwind/Petrov-Galerkin formulationsfor convective dominated flows with particular emphasis on the incompressible Navier-Stokes equations. Comput. Method. Appl. M., 32:199–259, 1982.

[21] F. Brunner, F. Radu, and P. Knabner. Analysis of an upwinded-mixed hybrid finiteelement method for transport problems. SIAM J. Numer. Anal., 52(1):83–102, 2014.

[22] F. Brunner, F.A. Radu, M. Bause, and P. Knabner. Optimal order convergence of amodified BDM1 mixed finite element scheme for reactive transport in porous media.Adv. Water Resour., 35:163–171, 2012.

[23] E. Burman. Consistent SUPG-method for transient transport problems: Stability andconvergence. Comput. Method. Appl. M., 96(2):239–253, 2010.

[24] M.A. Celia. Fundamental concepts for numerical simulation of contaminant transportand biodegradation. In Fundamentals and Applications of Bioremediation: Principles,Vol. I, pages 59–88. Technomic Publishing Co. Inc., 1997.

216

Page 228: Modeling, Analysis, and Simulation of Adsorption in ...

[25] M.A. Celia and P. Binning. A mass conservative numerical solution for two-phaseflow in porous media with application to unsaturated flow. Water Resour. Res.,28(10):2819–2828, 1992.

[26] M.A. Celia and P. Binning. Multiphase models of unsaturated flow: Approaches tothe governing equations and numerical methods. In Computational Methods in WaterResources IX, Vol. I: Numerical Methods in Water Resources, pages 257–272. ElsevierApplied Science, 1992.

[27] M.A. Celia, E.T. Bouloutas, and R.L. Zarba. A general mass-conservative numericalsolution for the unsaturated flow equation. Water Resour. Res., 26(7):1483–1496,1990.

[28] M.A. Celia, T.F. Herrea, and R.E. Ewing. An Eulerian Lagrangian localized adjointmethod for the advection-diffusion equation. Adv. Water Resour., 13:187–206, 1990.

[29] M.A. Celia, J.S. Kindred, and I. Herrera. Contaminant transport and biodegradationI. A numerical model for reactive transport in porous media. Water Resour. Res.,25(6):1141–1148, 1989.

[30] M.A. Celia and S. Zisman. An Eulerian-Lagrangian localized adjoint method for re-active transport in groundwater. In G. Gambolati, A. Rinaldo, C.A. Brebbio, W.G.Gray, and G.F. Pinder, editors, Computational Methods in Subsurface Hydrology: Pro-ceedings of the Eighth International Conference on Computational Methods in WaterResources, pages 383–392, Venice, Italy, 1990.

[31] G. Chavent and J. Jaffre. Mathematical Models and Finite Elements for ReservoirSimulation. North Holland, 1986.

[32] H.C.S Chenette and S.M. Husson. Membrane adsorbers comprising grafted glycopoly-mers for targeted lectin binding. J. Appl. Polym. Sci., 132:41437(1)–41437(7), 2015.

[33] H.C.S. Chenette, J.R. Robinson, E. Hobley, and S.M. Husson. Development of high-productivity, strong cation-exchange adsorbers for protein capture by graft polymer-ization from membranes with different pore sizes. J. Membrane Sci., 423–424:43–52,2012.

[34] J.C. Chrispell, V.J. Ervin, and E.W. Jenkins. A fractional step θ-method forconvection-diffusion problems. Technical Report TR2006 11 CEJ, Clemson Univer-sity, 2006.

[35] P.G. Ciarlet. Linear and Nonlinear Functional Analysis with Applications, volume130 of Applied Mathematics. SIAM, 2013.

[36] B. Cockburn, B. Dong, and J. Guzman. A superconvergent LDG-hybridizableGalerkin method for second-order elliptic problems. Math. Comput., 77:1887–1916,2008.

217

Page 229: Modeling, Analysis, and Simulation of Adsorption in ...

[37] B. Cockburn, B. Dong, J. Guzman, M. Restelli, and R. Sacco. A hybridizable dis-continuous Galerkin method for steady-state convection-diffusion reaction problems.SIAM J. Sci. Comput., 31(5):3827–3846, 2009.

[38] B. Cockburn, J. Gopalakrishnan, and R. Lazarov. Unified hybridization of discon-tinuous Galerkin, mixed, and continuous Galerkin methods for second order ellipticproblems. SIAM J. Numer. Anal., 47:1319–1365, 2009.

[39] B. Cockburn, J. Guzman, and H. Wang. Superconvergent discontinuous Galerkinmethods for second-order elliptic problems. Math. Comput., 78:1–24, 2009.

[40] B. Cockburn, G.E. Karniadakis, and C.-W Shu. Discontinuous Galerkin Methods:Theory, Computation, and Applications, volume 11 of Lecture Notes in ComputationalScience and Engineering. Springer, Berlin, 2011.

[41] B. Cockburn and C.-W. Shu. The local discontinuous Galerkin method for time-dependent convection-diffusion systems. SIAM J. Numer. Anal., 35(6):2449–2463,1998.

[42] R. Codina. Comparison of some finite element methods for solving the diffusion-convection-reaction equation. Comput. Method. Appl. M., 156:185–210, 1998.

[43] T. Codina, E. Oate, and M. Cervera. The intrinsic time for the streamlineupwind/Petrov-Galerkin formulation using quadratic elements. Comput. Method.Appl. M., 94(2):239–262, 1992.

[44] V. Coker. Biotherapeutics outpace conventional therapies. BioPharm Int. Supple-ments, 25(3):20–23, 2012.

[45] P.V. Danckwerts. Continuous flow systems. Distribution of residence times. Chem.Eng. Sci., 2:1–13, 1953.

[46] C. Dawson. Godunov-mixed methods for advective flow problems in one space di-mension. SIAM J. Numer. Anal., 28(5):1282–1309, 1991.

[47] C. Dawson. Godunov-mixed methods for advection-diffusion equations in multidi-mensions. SIAM J. Numer. Anal., 30(5):1315–1332, 1993.

[48] C. Dawson. Analysis of an upwind-mixed finite element method for nonlinear con-taminant transport equations. SIAM J. Numer. Anal., 35(5):1709–1724, 1998.

[49] C. Dawson and V. Aizinger. Upwind-mixed methods for transport equations. Com-putat. Geosci., 3:93–110, 1999.

[50] G. de Marsily. Quantitative Hydrogeology: Groundwater Hydrology for Engineers.Academic Press, 1986.

[51] E.G. Dutra do Carmo and G.B. Alvarez. A new stabilized finite element for-mulation for scalar convection-diffusion problems: the streamline and approximateupwind/Petrov-Galerkin method. Comput. Method. Appl. M., 192:3379–3396, 2003.

218

Page 230: Modeling, Analysis, and Simulation of Adsorption in ...

[52] H. Egger and J. Schoberl. A hybrid mixed discontinuous Galerkin finite-elementmethod for convection-diffusion problems. IMA J. Numer. Anal., 30:1206–1234, 2010.

[53] H.N. Endres, J.A.C. Johnson, C.A. Ross, J.K. Welp, and M.E. Etzel. Evaluation ofan ion-exchange membrane for the purification of plasmid DNA. Biotechnol. Appl.Bioc., 37(3):259–266, 2003.

[54] V.J. Ervin and W.W. Miles. Approximation of time-dependent visco-elastic fluid flow:SUPG approximation. SIAM J. Numer. Anal., 41(2):457–486, 2003.

[55] R.E. Ewing and M.A. Celia. Numerical methods for reactive transport and biodegra-dation. In Computational Methods in Water Resources IX, Vol. I: Numerical Methodsin Water Resources, pages 51–58. Elsevier Applied Science, 1992.

[56] R.R. Ewing, H. Wang, R.C. Sharpley, and M.A. Celia. A three-dimensional finiteelement simulation for transport of nuclear waste. Comput. Meth. Adv. Geomech.,9:2673–2679, 1995.

[57] M. Fahs, A. Younes, and F. Lehmann. An easy and efficient combination of the mixedfinite element method and the method of lines for the resolution of Richards’ equation.Environ. Modell. Softw., 24:1122–1126, 2009.

[58] H. Fang and Y. Saad. Two classes of multisecant methods for nonlinear acceleration.Numer. Linear Algebr., 16(3):197–221, 2009.

[59] M.W. Farthing, C.E. Kees, T.S. Coffey, C.T. Kelley, and C.T. Miller. Efficient steady-state solution techniques for variably saturated groundwater flow. Adv. Water Re-sour., 26:833–849, 2003.

[60] M.W. Farthing, C.E. Kees, T.F. Russell, and C.T. Miller. An ELLAM approxima-tion for advective-dispersive transport with nonlinear sorption. Adv. Water Resour.,29:657–675, 2006.

[61] C. Forkel and M.A. Celia. Numerical simulation of unsaturated flow and contaminanttransport with density and viscosity dependence. In Computational Methods in WaterResources IX, Vol. II: Mathematical Modeling in Water Resources, pages 351–358.Elsevier Applied Science, 1992.

[62] L.P. Franca, S.L. Frey, and T.J.R. Hughes. Stabilized finite element methods: I.Application to the advective-diffusive model. Comput. Method. Appl. M., 95(2):253–276, 1992.

[63] J. Frank, W. Hundsdorfer, and J.G. Verwer. On the stability of implicit-explicit linearmultistep methods. Appl. Numer. Math., 25:193–2005, 1997.

[64] A.C. Galeao, R.C. Almeida, S.M.C. Malta, and A.F.D. Loula. Finite element analysisof convection dominated reaction-diffusion problems. Appl. Numer. Math., 48:205–222, 2004.

219

Page 231: Modeling, Analysis, and Simulation of Adsorption in ...

[65] D. Gao, D.-Q. Lin, and S.-J. Yao. Measurement and correlation of protein adsorptionwith mixed-mode adsorbents taking into account the influences of salt concentrationand pH. J. Chem. Eng. Data, 51(4):1205–1211, 2006.

[66] K. Gerdes, J.M. Melenk, C. Schwab, and D. Schotzau. The hp version of the streamlinediffusion finite element method in two space dimensions. Math. Mod. Meth. Appl. S.,11(2):301–337, 2001.

[67] S. Ghose, B. Hubbard, and S.M. Cramer. Protein interactions in hydrophobic chargeinduction chromatography. Biotechnol. Progr., 21(2):498–508, 2005.

[68] R. Glowinski and J. Periaux. Numerical methods for nonlinear problems in fluiddynamics. In A. Lichneqsky and C. Saguez, editors, Supercomputing, pages 381–479.Elsevier Science Publishers B.V., 1987.

[69] J.F. Guarnaccia, P.T. Imhoff, B.C. Missildine, M. Oostrom, M.A. Celia, J.H. Dane,P.R. Jaffe, and G.F. Pinder. Multiphase chemical transport in porous media. Environ-mental Research Brief EPA/600/S-92/002, United States Environmental ProtectionAgency, Robert S. Kerr Environmental Research Laboratory, Ada, OK 74820, March1992.

[70] W. Heinrichs. Defect correction for convection-dominated flow. SIAM J. Sci. Comp.,17(5):1082–1091, 1996.

[71] J.S. Hesthaven and T. Warburton. Nodal Discontinuous Galerkin Methods: Algo-rithms, Analysis, and Application. Springer Science and Business Media, New York,2008.

[72] A. Hiller. Fast growth foreseen for protein therapeutics. Genet. Eng. Biotechn.,29:153–155, 2009. http://www.genengnews.com/gen-articles/fast-growth-foreseen-for-protein-therapeutics/2722/.

[73] Russell Hooper, Matt Hopkins, Roger Pawlowski, Brian Carnes, and Harry K. Moffat.Final report on LDRD project: Coupling strategies for multi-physics applications.Technical report, Sandia, 2007.

[74] P. Houston, C. Schwab, and E. Suli. Stabilized hp-finite element methods for first-order hyperbolic problems. SIAM J. Numer. Anal., 37(5):1618–1643, 2000.

[75] T.J.R. Hughes and A.N. Brooks. A multi-dimensional upwind scheme with no cross-wind diffusion. In T.J.R. Hughes, editor, Finite Element Methods for ConvectionDominated Flows, pages 19–35. American Society of Mechanical Engineers, New York,1979.

[76] T.J.R. Hughes and A.N. Brooks. A theoretical framework for Petrov-Galerkin meth-ods, with discontinuous weighting functions: Applications to the streamline upwindprocedure. In R.H. Gallagher, D.M. Norrie, J.T. Oden, and O.C. Zienkiewicz, edi-tors, Finite Elements in Fluids, volume IV, pages 46–65. John Wiley & Sons, London,1982.

220

Page 232: Modeling, Analysis, and Simulation of Adsorption in ...

[77] T.J.R. Hughes, L.P. Franca, and G.M. Hulbert. A new finite element formula-tion for computational fluid dynamics: VIII. The Galerkin/least-squares method foradvective-diffusive equations. Comput. Method. Appl. M., 73(2):173–189, 1989.

[78] T.J.R. Hughes and M. Mallet. A new finite element formulation for computa-tional fluid dynamics: III. The generalized streamline operator for multidimensionaladvective-diffusive systems. Comput. Method. Appl. M., 58(3):305–328, 1986.

[79] P.S. Huyakorn, P.F. Anderson, J.W. Mercer, and H.O. White. Saltwater intrusion inaquifers: Development and testing of a three-dimensional finite element model. WaterResour. Res., 23(2):293–312, 1987.

[80] V. John and J. Novo. Error analysis of the SUPG finite element discretization ofevolutionary convection-diffusion-reaction equations. SINUM, 49:1149–1176, 2011.

[81] C. Johnson. Streamline diffusion methods for problems in fluid mechanics. In FiniteElements in Fluids VI. Wiley, New York, 1986.

[82] C. Johnson. Numerical Solution of Partial Differential Equations by the Finite Ele-ment Method. Dover, 1987.

[83] C. Johnson. Numerical Solution of Partial Differential Equations by the Finite Ele-ment Method. Dover, Mineola, NY, 2009.

[84] C. Johnson and U. Navert. An analysis of some finite element methods for advection-diffusion problems. In S. Axelsson, L.S. Frank, and A. van der Sluis, editors, Analyticaland Numerical Approaches to Asymptotic Problems in Analysis, pages 99–116. North-Holland Publishing Company, 1981.

[85] C. Johnson, U. Navert, and J. Pitkaranta. Finite element methods for linear hyper-bolic problems. Comput. Method. Appl. M., 45(1–3):285–312, 1984.

[86] J. Douglas, Jr. and J.E. Roberts. Mixed finite element methods for second orderelliptic problems. Comput. Appl. Math., 1:91–103, 1982.

[87] J. Douglas, Jr. and J.E. Roberts. Global estimates for mixed methods for secondorder elliptic equations. Math. Comput., 44(169):39–52, 1985.

[88] S.-H. Ju and K.-J.S. Kung. Mass types, element orders and solution schemes for theRichards’ equation. Comput. Geosci., 23(2):175–187, 1997.

[89] K. Kallberg, H.O. Johansson, and L. Bulow. Multimodal chromatography: An ef-ficient tool in downstream processing of proteins. Biotechnol. J., 7(12):1485–1495,2012.

[90] J. Kacur, B. Malengier, and M. Remesıkova. Solution of contaminant transport andequilibrium and non-equilibrium adsorption. Comput. Method. Appl. M., 194:479–489,2005.

221

Page 233: Modeling, Analysis, and Simulation of Adsorption in ...

[91] J. Kacur, B. Malengier, and M. Remesıkova. Convergence of an operator splittingmethod on a bounded domain for a convection-diffusion-reaction system. J. Math.Anal. Appl., 348:894–914, 2008.

[92] D.W. Kelly, S. Nakazawa, O.C. Zienkiewicz, and J.C. Heinrich. A note on upwindingand anisotropic balancing dissipation in finite element approximations to convectivediffusion problems. Int. J. Numer. Meth. Eng., 15(11):1705–1711, 1980.

[93] D.E. Keyes, L.C. McInnes, C. Woodward, et al. Multiphysics simulations: Challengesand opportunities. Int. J. High Perform. C., 27:4–83, 2011.

[94] L.A. Khan and P.L.-F. Liu. An operator splitting algorithm for coupled one-dimensional advection-diffusion-reaction equations. Comput. Method. Appl. M.,127:181–201, 1995.

[95] Liaqat Ali Khan and Philip L.-F. Liu. Numerical analyses of operator-splitting algo-rithms for the two-dimensional advection-diffusion equation. Comput. Method. Appl.M., 152:337–359, 1998.

[96] J.S. Kindred and M.A. Celia. Contaminant transport and biodegradation 2. Concep-tual model and test simulations. Water Resour. Res., 25(6):1149–1159, 1989.

[97] E. Kreyszig. Introductory Functional Analysis with Applications. Wiley, 1989.

[98] D. Kuzmin, R. Lohner, and S. Turek. Flux-Corrected Transport: Principle, algorithms,and applications. Springer, 2012.

[99] E.S. Langer. Focus on efficiency: Single-use, analytical methods and downstreamprocessing at the forefront. Pharm. Manuf., March:3–11, 2013.

[100] D. Lanser and J.G. Verwer. Analysis of operator splitting for advection-diffusion-reaction problems from air pollution modelling. J. Comput. Appl. Math., 111:201–216,1999.

[101] B. Leader, Q.J. Baca, and D.E. Golan. Protein therapeutics: a summary and phar-macological classification. Nat. Rev. Drug Discov., 7(1):21–39, 2008.

[102] P.A. Lott, H.F. Walker, C.S. Woodward, and U.M. Yang. An accelerated Picardmethod for nonlinear systems related to variably saturated flow. Adv. Water Resour.,38:92–101, 2012.

[103] G.I. Marchuck. Methods of Numerical Mathematics. Springer-Verlag, New York,English edition, 1975.

[104] G.I. Marchuk. Splitting and alternating direction methods. In P.G. Ciarlet and J.L.Lions, editors, Handbook of Numerical Analysis, volume 1, pages 197–652. ElsevierScience Publishers B.V., 1990.

[105] S. Mehl. Use of Picard and Newton iteration for solving nonlinear ground water flowequations. Ground Water, 44(4):583–594, 2006.

222

Page 234: Modeling, Analysis, and Simulation of Adsorption in ...

[106] J.M. Melenk and C. Schwab. The hp streamline diffusion finite element method forconvection-dominated problems in one space dimension. Technical Report 98-10, SwissFederal Institute of Technology Zurich, 1999.

[107] T.J. Menkhaus, H. Varadaraju, L. Zhang, S. Schneiderman, S. Bjustrom, L. Liu, andH. Fong. Electospun nano fiber membranes surface functionalized with 3-dimensionalnanolayers as an innovative adsorption medium with ultra-high capacity and through-put. Chem. Commun., 46(21):3720–3722, 2010.

[108] S. Micheletti, R. Sacco, and F. Saleri. On some mixed finite element methods withnumerical integration. SIAM J. Sci. Comput., 23(1):245–270, 2001.

[109] C.T. Miller, C.N. Dawson, M.W. Farthing, T.Y. Hou, J. Huang, C.E. Kees, C.T.Kelley, and H.P. Langtangen. Numerical simulation of water resources problems:Models, methods, and trends. Adv. Water Resour., 51:405–437, 2013.

[110] C.T. Miller, G.A. Williams, C.T. Kelley, and M.D. Tocci. Robust solution of Richards’equation. Water Resour. Res., 34(10):2599–2610, 1998.

[111] A. Mizukami and T.J.R. Hughes. A Petrov-Galerkin finite element method forconvection-dominated flows: An accurate upwinding technique for satisfying the max-imum principle. Comput. Method. Appl. M., 50(2):181–193, 1985.

[112] J.M. Mollerup. Applied thermodynamics: A new frontier for biotechnology. FluidPhase Equilibr., 241(12):205–215, 2006.

[113] J.M. Mollerup. The thermodynamic principles of ligand binding in chromatographyand biology. J. Biotechnol., 132(2):187–195, 2007.

[114] J.M. Mollerup. A review of the thermodynamics of protein association to ligands,protein adsorption and adsorption isotherm. Chem. Eng. Technol., 31(6):864–874,2008.

[115] J.M. Mollerup, T.B. Hansen, S. Kidal, L. Sejergaard, and A. Staby. Development,modelling, optimisation and scale-up of chromatographic purification of a therapeuticprotein. Fluid Phase Equilibr., 261(12):133–139, 2007.

[116] J.M. Mollerup, T.B. Hansen, S. Kidal, and A. Staby. Quality by design—thermodynamic modelling of chromatographic separation of proteins. J. Chromatogr.A, 1177(2):200–206, 2008.

[117] D. Nadarajah and A. Mehta. Overload and elute chromatography. Patent applicationPCT/US2012/063242, 2012.

[118] U. Navert. The streamline diffusion method for time dependent convection: diffusionproblems with small diffusion. Technical report, Chalmers University of Technology,Goeteborg (Sweden), 1981.

[119] U. Navert. A Finite Element Method for Convection-Diffusion Problems. PhD thesis,Chalmers University of Technology, Goteborg, 1982.

223

Page 235: Modeling, Analysis, and Simulation of Adsorption in ...

[120] B.K. Nfor, M. Noverraz, S. Chilamkurthi, P.D.E.M. Verhaert, L.A.M. van der Wielen,and M. Ottens. High-throughput isotherm determination and thermodynamic mod-eling of protein adsorption on mixed mode adsorbents. J. Chromatogr. A, 1217:6829–6850, 2010.

[121] R.H. Nochetto and C. Verdi. Approximation of degenerate parabolic problems usingnumerical integration. SIAM J. Numer. Anal., 25:784–814, 1988.

[122] C. Paniconi and M. Putti. A comparison of Picard and Newton iteration in thenumerical solution of multidimension variably saturated flow problems. Water Resour.Res., 30(12):3357–3374, 1994.

[123] D.W. Peaceman and H.H. Rachford. The numerical solution of parabolic and ellipticdifferential equations. SIAM, 3(1), 1955.

[124] G.F. Pinder and M.A. Celia. Subsurface Hydrology. Wiley Interscience, 2006.

[125] R.H. Pletcher, J.C. Tannehil, and D. Anderson. Computational Fluid Mechanics andHeat Transfer. Series in Computational and Physical Processes in Mechanics andThermal Sciences. Taylor & Francis, third edition, 2012.

[126] L. Portero and J.C. Jorge. A generalization of Peaceman-Rachford fractional stepmethod. J. Comput. Appl. Math., 189:676–688, 2006.

[127] M. Putti and C. Paniconi. Picard and Newton linearization for the coupled model ofsaltwater intrusion in aquifers. Adv. Water Resour., 18(3):159–170, 1995.

[128] A. Quarteroni and A. Valli. Numerical Approximation of Partial Differential Equa-tions, volume 23 of Springer Series in Computational Mathematics. Springer-Verlag,Berlin Heidelberg, 1994.

[129] F. Radu, I.S. Pop, and P. Knabner. Error estimates for an Euler implicit, mixedfinite element discretization of Richards’ equation: Equivalence between mixed andconformal approaches. RANA 02-06, Eindhoven University of Technology, 2002.

[130] F. Radu, I.S. Pop, and P. Knabner. Order of convergence estimates for an Eulerimplicit, mixed finite element discretization of Richards’ equation. SIAM J. Numer.Anal., 42(4):1452–1478, 2004.

[131] F.A. Radu, M. Bause, A. Prechtel, and S. Attinger. A mixed hybrid finite elementdiscretization scheme for reactive transport in porous media. In K. Kunisch, G. Of,and O. Steinbach, editors, Numerical Mathematics and Advanced Applications, pages513–520. Springer, 2008.

[132] F.A. Radu and I.S. Pop. Newton method for reactive solute transport with equilibriumsorption in porous media. J. Comput. Appl. Math., 234:2118–2127, 2010.

[133] F.A. Radu and I.S. Pop. Mixed finite element discretization and Newton iteration fora reactive contaminant transport model with nonequilibrium sorption: Convergenceanalysis and error estimates. Computat. Geosci., 15:431–450, 2011.

224

Page 236: Modeling, Analysis, and Simulation of Adsorption in ...

[134] F.A. Radu, I.S. Pop, and S. Attinger. Analysis of an Euler implicit-mixed finiteelement scheme for reactive solute transport in porous media. Numer. Meth. Par. DE, 26(2):320–344, 2010.

[135] F.A. Radu, I.S. Pop, S. Attinger, and P. Knabner. Error estimates for an Eulerimplicit-mixed finite element scheme for reactive transport in saturated/unsaturatedsoil. Proc. Appl. Math. Mech., 7:1024705–1024706, 2007.

[136] F.A. Radu, I.S. Pop, and P. Knabner. Error estimates for a mixed finite elementdiscretization of some degenerate parabolic equations. Numer. Math., 109:285–311,2008.

[137] F.A. Radu, N. Suciu, J. Hoffman, A. Vogel, O. Kolditz, C.-H. Park, and S. Attinger.Accuracy of numerical simulations of contaminant transport in heterogeneous aquifers:A comparitive study. Adv. Water Resour., 34:47–61, 2011.

[138] F.A. Radu and W. Wang. Convergence analysis for a mixed finite element scheme forflow in strictly unsaturated porous media. Nonlinear Anal.-Real., 15:266–275, 2014.

[139] A.S. Rathore. Quality by design for biopharmaceuticals. Nat. Biotechnol., 27:26–34,2009.

[140] P.A. Raviart and J.M. Thomas. A mixed finite element method for 2-nd order ellipticproblems. In I. Galligani and E. Magenes, editors, Mathematical Aspects of FiniteElement Methods: Proceedings of the Conference Held in Rome, December 10-12,1975, pages 292–315. Springer, 1977.

[141] M. Remesıkova. Solution of convection-diffusion problems with nonequilibrium ad-sorption. J. Comput. Appl. Math., 169:101–116, 2004.

[142] M. Remesıkova. Numerical solution of two-dimensional convection-diffusion-adsorption problems using an operator splitting scheme. Appl. Math. Comput.,184:116–130, 2007.

[143] B. Riviere. Discontinuous Galerkin Methods for Solving Elliptic and Parabolic Equa-tions. SIAM, 2008.

[144] B. Riviere and M.F. Wheeler. Discontinuous Galerkin methods for flow and transportproblems in porous media. Int. J. Numer. Meth. Biomed. Eng., 18(1):63–68, 2002.

[145] D.L. Ropp, J.N. Shadid, and C.C. Ober. Studies of the accuracy of time integrationmethods for reaction-diffusion equations. J. Comput. Phys., 194:544–574, 2004.

[146] R. Sacco and F. Saleri. Stabilization of mixed finite elements for convection-diffusionproblems. CWI Quart., 10:301–315, 1997.

[147] E. Schneid, P. Knabner, and F. Radu. A priori error estimates for a mixed finiteelement discretization of the Richards’ equation. Numer. Math., 98:353–370, 2004.

[148] N. Singh, S.M. Husson, B. Zdyrko, and I. Luzinov. Surface modification of microp-orous PVDF membranes by ATRP. J. Membrane Sci., 262:81–90, 2005.

225

Page 237: Modeling, Analysis, and Simulation of Adsorption in ...

[149] N. Singh, J. Wang, M. Ulbrict, S.R. Wickramasinghe, and S.M. Husson. Surface-initiated atom transfer radical polymerization: A new method for the preparation ofpolymeric membrane adsorbers. J. Membrane Sci., 309:64–72, 2008.

[150] M. Slodicka. Error estimates of an efficient linearization scheme for a nonlinear ellipticproblem with a nonlocal boundary condition. RAIRO-Math. Model. Num., 35:691–711, 2001.

[151] M. Slodicka. A robust and efficient linearization scheme for doubly nonlinear anddegenerate parabolic problems arising in flow in porous media. SIAM J. Sci. Comput.,23(5):1593–1614, 2002.

[152] M. Slodicka. A robust linearisation scheme for a nonlinear elliptic boundary valueproblem: Error estimates. ANZIAM J., 46:449–470, 2005.

[153] M. Stynes. Steady-state convection-diffusion problems. Acta Numerica, 14:445–508,2005.

[154] S.-Y. Suen and M.R. Etzel. A mathematical analysis of affinity membrane biosepara-tions. Chem. Eng. Sci., 47(6):1355–1364, 1992.

[155] J. Talbot, G. Tarjus, P.R. van Tassel, and P. Viot. From car parking to proteinadsorption: An overview of sequential adsorption processes. Colloid. Surface. A,165:287–324, 2000.

[156] A. Tarafder. Modeling and multi-objective optimization of a chromatographic sys-tem. In A. Bonilla-Petriciolet G.P. Rangaiah, editor, Multi-Objective Optimization inChemical Engineering: Developments and Applications. John Wiley & Sons, 2013.

[157] T.E. Tezduyar. Finite element formulations for hyperbolic systems with particularemphasis on the compressible Euler equations. PhD thesis, California Institute ofTechnology, 1982.

[158] J. Thommes and M.R. Etzel. Alternatives to chromatographic separations. Biotech-nol. Progr., 23(1):42–45, 2007.

[159] M.D. Tocci, C.T. Kelley, and C.T. Miller. Accurate and economical solution of thepressure-head form of Richards’ equation by the method of lines. Adv. Water Resour.,20(1):1–14, 1997.

[160] M.D. Tocci, C.T. Kelley, C.T. Miller, and C.E. Kees. Inexact Newton methods andthe method of lines for solving Richards equation in two space dimensions. Computat.Geosci., 2:291–309, 1998.

[161] S. Turek. Efficient Solvers for Incompressible Flow Problems, volume 6 of LectureNotes in Computational Science and Engineering. Springer-Verlag, Berlin, 1999.

[162] M. Vohralık. A posteriori error estimates for lowest-order mixed finite elementdiscretizations of convection-diffusion-reaction equations. SIAM J. Numer. Anal.,45(4):1570–1599, 2007.

226

Page 238: Modeling, Analysis, and Simulation of Adsorption in ...

[163] J. von Neumann and R.D. Richtmeyer. A method for the numerical calculation ofhydrodynamical shocks. J. Appl. Phys., 21(3):232–237, 1950.

[164] H.F. Walker and P. Ni. Anderson acceleration for fixed-point iterations. SIAM J.Numer. Anal., 49(4):1715–1735, 2011.

[165] H.F. Walker, C.S. Woodward, and U.M. Yang. An accelerated fixed-point iterationfor solution of variably saturated flow. In J. Carrera, X. Sanchez Villa, D. FernandezGarcia, et al., editors, Proceedings of the XVIII International Conference on Com-putational Methods in Water Resources (CMWR 2010), pages 216–223, Center forNumerical Methods in Engineering, Barcelona, Spain, 2010. CIMNE.

[166] J. Wang, R.T. Sproul, L.S. Anderson, and S.M. Husson. Development of multimodalmembrane adsorbers for antibody purification using atom transfer radical polymer-ization. Polymer, 55(6):1404–1411, 2014.

[167] J. Wang, A. Wilson, J.R. Robinson, E.W. Jenkins, and S.M. Husson. A new mul-timodal membrane adsorber for monoclonal antibody purifications. J. Membr. Sci.,492:137–146, 2015.

[168] M.F. Wheeler, W.A. Kinton, and C.N. Dawson. Time-splitting for advection-dominated parabolic problems in one space variable. Communications in AppliedNumerical Methods, 4:413–423, 1988.

[169] J. P. Whiteley, K. Gillow, and S. J. Tavener. Error bounds on block Gauss-Seidelsolutions of coupled multiphysics problems. Int. J. Numer. Meth. Eng., 88:1219–1237,2011.

[170] L.S. Wolfe, C.P. Barringer, S.S. Mostafa, and A.A. Shukla. Multimodal chromatogra-phy: Characterization of protein binding and selectivity enhancement through mobilephase modulators. J. Chromatogr. A, 1340:151–156, 2014.

[171] C. Woodward and C. Dawson. Analysis of expanded mixed finite element methods fora nonlinear parabolic equation modeling flow into variably saturated porous media.SIAM J. Numer. Anal., 37(3):701–724, 2000.

[172] M. Xu and Y. Eckstein. Statistical analysis of the relationships between dispersivityand other physical properties of porous media. Hydrogeol. J., 5(4):4–20, 1997.

[173] Y. Xu and C.-W. Shu. Error estimates of the semi-discrete local discontinuousGalerkin method for nonlinear convection-diffusion and KdV equations. Comput.Method. Appl. M., 196:3805–3822, 2007.

[174] N.N. Yanenko. The Method of Fractional Steps. Springer-Verlag, New York, Englishedition, 1971.

[175] H. Yang, M. Bitzer, and M.R. Etzel. Analysis of protein purification using ion-exchange membranes. Ind. Eng. Chem. Res., 38:4044–4050, 1999.

227

Page 239: Modeling, Analysis, and Simulation of Adsorption in ...

[176] A.H.M. Yusof and M. Ulbricht. Polypropylene-based membrane adsorbers via photo-initiated graft copolymerization: optimizing separation performance by preparationconditions. J. Membrane Sci., 311:294–305, 2008.

[177] E. Zeidler. Applied Functional Analysis: Applications to Mathematical Physics, vol-ume 108 of Applied Mathematical Sciences. Springer-Verlag, New York, 1995.

228


Recommended