ABSTRACT Title of Dissertation: GRAVITY DRIVEN INSTABILITIES OF TRANSIENT DIFFUSIVE BOUNDARY LAYERS IN POROUS MEDIA Don Daniel, Doctor of Philosophy, 2014 Dissertation directed by: Assistant Professor Amir Riaz Department of Mechanical Engineering This study is motivated by the geological storage of carbon dioxide in subsur- face saline aquifers. After injection in an aquifer, CO2 dissolves in brine to form a diffusive solute boundary layer that is gravitationally unstable leading to natural convection within the aquifer. The exploration of the underlying hydrodynamic in- stability is of practical importance because of enhanced CO2 dissolution and storage in aquifers. In comparison to the classical Rayleigh-Be´nard convection in a heated fluid cell, the analysis is not straightforward because the CO2 boundary layer is unsteady and nonlinear. The physics of the convective instability is described by a mathematical operator that is both non-autonomous and non-normal. Conse- quently, it is uncertain whether classical stability results for the onset of convection are valid. In addition, it is unclear how theoretical predictions compare with exper- iments. To explore these issues, the physical mechanisms and perturbation structures of transient, diffusive boundary layers are examined using multiple theoretical and computational tools. Traditional schemes based on linear stability theory, due to unique physical constraints, are unlikely to produce physically relevant perturba- tion structures. Therefore, a novel optimization procedure is formulated such that the optimization is restricted to physically admissible fields. The new method is compared with traditional stability approaches such as quasi-steady eigenvalue and classical optimization procedures along with fully resolved nonlinear direct numeri- cal simulations. After establishing a suitable analytical framework, the role of viscosity and permeability variation is examined on the onset of natural convection. Onset of convection occurs sooner when viscosity decreases with aquifer depth. These effects of viscosity variation are in contrast to observations in classical viscous fingering problems. When the porous medium is horizontally layered, qualitatively different instability characteristics can occur depending on the relative length scales of the boundary layer and the permeability variation. For sufficiently high permeability contrast, small changes in the permeability field can lead to large variations in the onset times for convection. Resonance effects are observed only when the porous medium is vertically layered. The current study provides a framework to explore gravity driven instabilities arising both in pure fluid and porous media applications. The framework can be extended to study more complex systems such as those involving chemically reacting species and random anisotropic permeability fields. GRAVITY DRIVEN INSTABILITIES OF TRANSIENT DIFFUSIVE BOUNDARY LAYERS IN POROUS MEDIA by Don Daniel Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2014 Advisory Committee: Assistant Professor Amir Riaz, Chair/Advisor, Professor Bala Balachandran, Professor James Duncan, Professor Kenneth Kiger, Professor Dianne P. O’Leary, Dean’s Representative. c© Copyright by Don Daniel 2014 Dedication To Issac Samuel and David Daniel! ii Acknowledgments To my Ph.D. advisor Amir Riaz for timely research guidance, care and for providing me with a very strong learning platform. To Nils Tilton for several important discussions in improving my research and writing skills, especially during his post-doc stint at the University of Maryland. To Hamdi Tchelepi for insightful discussions and collaboration opportunities. To Azar Nazeri and the Petroleum Institute, Abu Dhabi for funding during various stages of my academic career. To my dissertation defense and proposal committee members, Bala Balachan- dran, Dianne P. O’Leary, James Duncan, Kenneth Kiger, and Konstantina Trivisa for helpful suggestions and recommendations that aided in the prompt completion of this dissertation. To my colleagues for their valuable discussions, input, and support over the years: Marcos Vanella, Grigoris Panagakos, Khaled Abdelaziz, Clarence Baney, Moon Soo Lee, Mohammad Alizadeh, Zhipeng Qin, Abishek Gopal, Siavash Toosi, Zohreh Ghorbani, Kevin Shipley, and Meiqian Chen. To friends and acquaintances: O.V. Kiran, Ben Jacob, Bonney Luke Thomas, Deepu Prasad, Ashraf T. P., Rohit Jacob, Ajay Mathew Thomas, Vaishag, Jithin, Siddarth Sashidharan, Ebrahim Kheriwala, and others for having helped me in ways that I sometimes do not perceive. Thank you. To the many kind authors who provided free open source information on several topics such as tutorials on Matlab, Latex, etc, all of which helped me in completing iii my academic and research work in an efficient manner. To my family: mom Susen Daniel, dad Geevarghese Daniel Kutty, sister-in-law Sarah Daniel, brother Dan Daniel for understanding and love. To all my spiritual guides! To my dear God in whom I live, I breath, and I have my being! iv Table of Contents List of Tables vii List of Figures viii List of Abbreviations x 1 Introduction 1 1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 1.3 Challenges and Objectives . . . . . . . . . . . . . . . . . . . . . . . . 8 1.4 Outline of Dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . 10 2 Natural convection in homogeneous porous media 12 2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 2.2 Geometry and governing equations . . . . . . . . . . . . . . . . . . . 13 2.3 Linear stability methods . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.3.1 Quasi-steady eigenvalue analysis . . . . . . . . . . . . . . . . . 15 2.3.2 Classical Optimization . . . . . . . . . . . . . . . . . . . . . . 16 2.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 2.4.1 Effect of amplification measure . . . . . . . . . . . . . . . . . 20 2.4.2 Optimal perturbation structures . . . . . . . . . . . . . . . . . 24 2.4.3 Sensitivity to initial perturbation time . . . . . . . . . . . . . 26 2.4.4 Influence of final time on initial perturbation profiles . . . . . 34 2.4.5 Comparison with quasi-steady eigenvalue analysis . . . . . . . 36 2.5 Modified Optimization Procedure . . . . . . . . . . . . . . . . . . . . 39 2.5.1 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 2.5.2 Filter Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 42 2.5.3 Comparison with classical optimization scheme . . . . . . . . 44 2.5.4 Comparison with eigenvalue and initial value problems . . . . 50 2.6 Direct Numerical Simulations . . . . . . . . . . . . . . . . . . . . . . 53 2.6.1 Simulations of physical systems . . . . . . . . . . . . . . . . . 53 2.6.2 Extent of linear regime and onset of convection . . . . . . . . 56 2.7 Conclusions and summary . . . . . . . . . . . . . . . . . . . . . . . . 59 v 3 Effect of viscosity contrast on gravity-driven instabilities in porous media 63 3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 3.2 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 3.3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 73 3.3.1 The fixed interface model . . . . . . . . . . . . . . . . . . . . 73 3.3.2 The moving interface model . . . . . . . . . . . . . . . . . . . 80 3.3.3 Effect of non-monotonic density profiles . . . . . . . . . . . . . 82 3.3.4 Effect of uniform flow . . . . . . . . . . . . . . . . . . . . . . . 88 3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 4 Natural convection in horizontally layered porous media 98 4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 4.2 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 4.3 Linear perturbation growth . . . . . . . . . . . . . . . . . . . . . . . 103 4.3.1 Influence of permeability variance and phase . . . . . . . . . . 104 4.3.2 Stability mechanisms . . . . . . . . . . . . . . . . . . . . . . . 106 4.3.3 Effect of correlation length . . . . . . . . . . . . . . . . . . . . 111 4.4 Onset of natural convection . . . . . . . . . . . . . . . . . . . . . . . 114 4.4.1 Critical time of convection onset . . . . . . . . . . . . . . . . . 116 4.4.2 Influence of mode switching on convection onset . . . . . . . . 118 4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 5 Natural convection in vertically layered porous media 123 5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 5.2 Governing equations and methodology . . . . . . . . . . . . . . . . . 124 5.3 Linear growth characteristics . . . . . . . . . . . . . . . . . . . . . . . 128 5.3.1 Quasi-steady 2D eigenmodes . . . . . . . . . . . . . . . . . . . 128 5.3.2 Perturbation growth . . . . . . . . . . . . . . . . . . . . . . . 130 5.3.3 Scaling with Rayleigh number . . . . . . . . . . . . . . . . . . 134 5.4 Onset of natural convection . . . . . . . . . . . . . . . . . . . . . . . 135 5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139 6 Contributions and Future work 140 6.1 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140 6.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141 vi List of Tables 2.1 Negative concentrations of classical optimal profiles . . . . . . . . . . 39 2.2 Gaussian parameters for Direct Numerical Simulation . . . . . . . . . 54 3.1 Vorticity components in the fixed interface model . . . . . . . . . . . 76 3.2 Vorticity components in the displaced interface model . . . . . . . . . 91 4.1 Vorticity components due to permeability heterogeneity . . . . . . . . 110 vii List of Figures 1.1 Sketch of CO2 sequestration . . . . . . . . . . . . . . . . . . . . . . . 2 1.2 Dissolution flux of CO2 . . . . . . . . . . . . . . . . . . . . . . . . . . 6 2.1 Geometry of study . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 2.2 Optimization results when maximizing various amplification measures 21 2.3 Dominant perturbations . . . . . . . . . . . . . . . . . . . . . . . . . 25 2.4 Effect of initial perturbation time . . . . . . . . . . . . . . . . . . . . 28 2.5 Effect of initial time on perturbation profiles . . . . . . . . . . . . . . 29 2.6 Optimal point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.7 Convergence of the optimal initial profiles . . . . . . . . . . . . . . . 34 2.8 IVP results using random initial profiles . . . . . . . . . . . . . . . . 35 2.9 Comparing optimal perturbations with least stable QSSA mode . . . 37 2.10 Optimization using filter functions, Ψ1 and Ψ2 . . . . . . . . . . . . . 42 2.11 Optimization using filter function Ψ3 . . . . . . . . . . . . . . . . . . 45 2.12 Comparison of the COP and MOP schemes . . . . . . . . . . . . . . 46 2.13 Optimal MOP point . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 2.14 Comparison of MOP with QSSAξ and IVP . . . . . . . . . . . . . . . 51 2.15 Validation of MOP with randomly perturbed DNS . . . . . . . . . . . 55 2.16 Onset of nonlinear convection – MOP and COP . . . . . . . . . . . . 57 3.1 Sketch of CO2 sequestration . . . . . . . . . . . . . . . . . . . . . . . 66 3.2 Nonmonotonic density and viscosity profiles . . . . . . . . . . . . . . 67 3.3 Various types of transient boundary layers . . . . . . . . . . . . . . . 70 3.4 Marginal stability contours of FI model . . . . . . . . . . . . . . . . . 72 3.5 Most dominant growthrates and wavenumbers of FI model . . . . . . 75 3.6 Eigenmodes of FI model . . . . . . . . . . . . . . . . . . . . . . . . . 76 3.7 Comparison between MI and FI models . . . . . . . . . . . . . . . . . 78 3.8 Agreement between FI and MI models . . . . . . . . . . . . . . . . . 79 3.9 Eigenmodes of MI model . . . . . . . . . . . . . . . . . . . . . . . . . 83 3.10 Effect of varying nonmonotonocity of density variation . . . . . . . . 84 3.11 Effect of density profiles on velocity eigenmode . . . . . . . . . . . . . 86 3.12 Critical viscosity ratio of the displaced interface model. . . . . . . . . 89 3.13 Effect of displacement velocity on the velocity eigenmode . . . . . . . 93 viii 3.14 Time invariance of critical viscosity ratio . . . . . . . . . . . . . . . . 94 4.1 Marginal stability contours of zero growth rate . . . . . . . . . . . . . 104 4.2 Perturbation and permeability profiles . . . . . . . . . . . . . . . . . 107 4.3 Spatial variation of vorticity . . . . . . . . . . . . . . . . . . . . . . . 109 4.4 Marginal stability contours for permeability phase of zero . . . . . . . 111 4.5 Spatial variation of perturbation and vorticity . . . . . . . . . . . . . 112 4.6 Critical time for the onset of instability . . . . . . . . . . . . . . . . . 114 4.7 Onset of convection . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 4.8 Role of perturbation amplitude . . . . . . . . . . . . . . . . . . . . . 120 5.1 Quasi-steady 2D eigenmodes . . . . . . . . . . . . . . . . . . . . . . . 127 5.2 Fourier components of 2D eigenmode . . . . . . . . . . . . . . . . . . 129 5.3 Temporal evolution of growth rate . . . . . . . . . . . . . . . . . . . . 131 5.4 Effect of permeability amplitude . . . . . . . . . . . . . . . . . . . . . 132 5.5 Scaling with Rayleigh number . . . . . . . . . . . . . . . . . . . . . . 134 5.6 Critical onset of natural convection . . . . . . . . . . . . . . . . . . . 138 ix List of Abbreviations NOAA National Oceanic and Atmospheric Administration ESRL Earth System Research Laboratory EVP Eigenvalue Problem QSSA Quasi-Steady EVP QSSAξ QSSA in transformed space COP Classical Optimization Procedure MOP Modified Optimization procedure DNS Direct Numerical Simulations IVP Initial Value Problem FI Fixed Interface model MI Moving Interface model DMA Dominant Mode Analysis x Chapter 1: Introduction 1.1 Motivation The past decades have seen a rapid increase in the man-made production of greenhouse gases such as carbon dioxide. Recent estimates of the concentration of atmospheric CO2 are at an all time high of about 397 ppm (NOAA-ESRL, 2014). The primary sources of CO2 emissions are attributed to the combustion of fossil fuels in power plants. To reduce emissions, it has been proposed to store CO2 in subsurface porous rock formations (Orr, 2009). The process of capturing and storing CO2 is referred to in past literature as CO2 sequestration. According to the sequestration procedure, CO2 would be captured from power plants and then injected into brine-saturated aquifers with large storage capacity. CO2 is stored within rocks through a combination of several mechanisms such as: • Structural trapping This trapping occurs when a layer of impermeable caprock prevents CO2 from leaking back into the atmosphere. • Residual trapping This occurs due to capillary trapping of CO2 within small pore spaces. 1 Figure 1.1: Sketch of CO2 sequestration process • Mineral trapping Given sufficient time, CO2 reacts with surrounding brine in aquifer to form stable carbonate minerals. • Solubility trapping CO2 dissolves into the surrounding brine and thus stored within the brine. This study explores the fluid mechanics of the solubility trapping mechanism. A schematic of the process is shown in figure 1. After injection, the buoyant CO2 initially rises and forms a horizontal layer beneath an impermeable caprock. With time, free CO2 dissolves into the underlying brine across the two-phase interface and forms a downwardly growing diffusive boundary layer. Because the CO2-rich brine in the boundary layer is denser than the underlying brine, a gravitational instability eventually results in finger-like structures that break away from the boundary layer and convect CO2 downwards into the aquifer. Due to the ensuing natural convection, free CO2 dissolves more faster into the brine. A clear understanding of the physical 2 mechanisms behind natural convection is vital to the modelling of CO2 sequestration (Huppert & Neufeld, 2014; Riaz & Cinar, 2014). 1.2 Review A gravitational instability usually occurs when a heavier fluid is present above a lighter fluid. When the interface between the two fluids is perturbed, the fluids interpenetrate to produce buoyant unstable fingers. Such an instability, which is referred to in literature as the Rayleigh–Taylor instability (Drazin & Reid, 2004), is observable in a wide scale of systems ranging from a modern lava lamp to the outer atmosphere of the sun. Similar instabilities are also formed when unstable density distributions exist within a single fluid, as in the case of Rayleigh–Be´nard convection. In this classical scenario, a horizontal layer fluid is heated from be- low such that a linear temperature profile is formed within the fluid, leading to a buoyancy-driven convection. For many natural phenomena, convection occurs well before the formation of the linear temperature profile, because of strong unfavorable density gradients within the developing thermal boundary layer. This form of early onset of natural convection in a transient, developing boundary layer is a common feature of several important porous media applications such as CO2 sequestration and groundwater flow (Wooding et al., 1997) as well as pure fluid applications such as heat transfer devices (Goldstein, 1959). The onset of natural convection can be explored using analogous experiments in laboratory conditions (Blair & Quinn, 1969). In addition to experiments, they can 3 also be analyzed using various theoretical methods. One popular method is to exam- ine the eigenvalues of the operator or matrix that describes the physical evolution of instabilities. Instabilities are also investigated by performing a computational simu- lation of the nonlinear equations that govern the physics. This dissertation explores gravity-driven instabilities that lead to natural convection in transient, diffusive boundary layers using a combination of theoretical and computational methods. The archetypal problem of Rayleigh–Be´nard convection in which a horizontal fluid heated from below is well understood as the underlying thermal gradient or boundary layer is steady and linear. The stability characteristics are determined by the dimensionless Rayleigh number, which is the ratio of the destabilizing buoyancy forces that drive convection to the stabilizing effects of diffusion. Convection occurs when the thermal gradients are strong enough to drive motion. Similar behavior is also observed for porous media flows (Horton & Rogers, 1945; Lapwood, 1948). For the classical Rayleigh–Be´nard convection, there is excellent agreement between experiments and theory. In comparison to classical Rayleigh-Be´nard convection, the diffusive boundary layers of CO2 sequestration are unsteady and nonlinear. This renders the linear sta- bility operator non-autonomous. In comparison, popular stability operators in fluid mechanics such as the Orr-Sommerfeld operator are not a function of time (Reddy et al., 1993). Furthermore, the governing equations are also non-self-adjoint and the resultant eigenmodes are non-orthogonal. The superposition of non-orthogonal eigenmodes may result in transient growth (Schmid, 2007). Because of these effects, it is uncertain whether existing theoretical methods predict the physics accurately. 4 It is also unclear how theoretical and experimental methods of transient boundary layers compare with each other. Earlier experimental studies provide a sound qualitative understanding of the evolution of transient diffusive boundary layers (Blair & Quinn, 1969; Elder, 1968). In the early stages of the boundary layer formation, perturbations to the layer are damped due to an initial diffusive period during which the stabilizing effects of viscosity dominate the physics. Eventually a critical time is reached at which the vertical density gradient is sufficient to drive fluid motion (Foster, 1965; Goldstein, 1959) such that perturbations begin to grow. Computational simulations suggest that for small initial perturbations, nonlinear effects increase slowly and, linear mechanisms can dominate for considerable time beyond the critical onset time for instabilities (Farajzadeh et al., 2007; Riaz et al., 2006; Selim & Rees, 2007b). Within this linear regime, the flux of CO2 into the brine decreases monotonically, see figure 1.2. Eventually nonlinear mechanisms cause the flux of CO2 to increase from that predicted by linear theory such that there is a turning point across which flux starts to increase. This turning point is referred to as the onset time of natural convection. The linear stability of diffusive boundary layers has been studied using three main approaches. The quasi-steady-state approach (QSSA) approximates the bound- ary layer as steady in comparison to the rapid growth of perturbations. The method was first applied to convection in fluid layers by Morton (1957), Goldstein (1959), and Lick (1965), and subsequently to porous layers by Robinson (1976). The QSSA is considered to be invalid for small times when the boundary layer grows rapidly. The stability of diffusive boundary layers has also been studied using energy meth- 5 Time Flu xo fC O 2 ONSET OF NATURALCONVECTION Figure 1.2: Dissolution flux of CO2 (solid line) into brine. The diffusive flux (dashed line) corresponds to the linear regime. Nonlinear effects begin to dominate at the onset of natural convection (solid dot). ods (Caltagirone, 1980; Homsy, 1973; Kim & Choi, 2007; Slim & Ramakrishnan, 2010). These methods determine lower bounds below which all perturbations decay, but do not provide information for the subsequent evolution of unstable perturba- tions in the linear regime. The third, and most popular, approach avoids the QSSA by solving the non-autonomous, linear, initial value problem (IVP) numerically for given initial conditions. The IVP was first considered by Foster (1965) to study thermal convection in pure fluid media. Foster focused on white noise initial conditions that excited all ver- tical Fourier modes equally. Gresho & Sani (1971) noted several drawbacks to this approach. First, white noise conditions produce initial temperature perturbations that vary far outside the boundary layer. In experiments, however, perturbations originate in the boundary layer (Blair & Quinn, 1969; Elder, 1968; Green & Foster, 6 1975; Spangenberg & Rowland, 1961; Wooding et al., 1997). Second, initial condi- tions may play a secondary role in systems continuously forced by noise. Lastly, the onset time for convection is due to nonlinear effects and cannot be predicted by the linear IVP. Subsequently, Jhaveri & Homsy (1982) modelled noise by adding random forcing to the momentum equation and reported good agreement with experiments. The IVP for thermal boundary layers in porous media was first considered by Caltagirone (1980) and Kaviany (1984), and subsequently applied to CO2 seques- tration by Ennis-King & Paterson (2005). All three studies used initial white noise conditions that spanned the entire domain. However, these initial conditions con- tradict previous experimental observations that suggest that perturbations originate within the boundary layer. Consequently, two methods have emerged that suggest more appropriate initial conditions for the IVP. The first approach (Ben et al., 2002; Kim & Choi, 2011; Riaz et al., 2006) involves a similarity transformation of the lin- ear stability operator. As a result, the newly transformed eigenmodes are always localized within the boundary layer. Riaz et al. (2006) also developed the “dominant mode analysis” (DMA) that involved a projection of the dynamics onto a weighted Hermite polynomial. This technique produced analytical expressions for onset time that showed good agreement with corresponding results of IVP simulations. The second approach to determining initial conditions for the IVP uses non- modal stability theory (Farrell & Ioannou, 1996a,b; Schmid, 2007) to determine opti- mal initial perturbations with maximum amplification over a certain period of time. In this manner, Rapaka et al. (2008), using a singular value decomposition method, showed that optimal perturbations tend to be more localized within the boundary 7 layer than white noise conditions. Subsequently, Doumenc et al. (2010) studied non-modal growth in transient Rayleigh-Be´nard-Marangoni convection. This was performed using an adjoint-based method. An approach similar to that of Rapaka et al. (2008) or Doumenc et al. (2010) is to optimize for the growth rate by formu- lating an eigenvalue problem to determine the numerical abscissa. This approach has been taken by Slim & Ramakrishnan (2010) and Kim & Choi (2012). 1.3 Challenges and Objectives Though the stability of diffusive boundary layers has been studied extensively using multiple methods, it is still uncertain which methodology best reflects the physics of the problem. This uncertainty is best illustrated by examining recent works. Ghesmat et al. (2011) performed a QSSA eigenvalue analysis to study the effect of chemical reactions within the boundary layer . Elenius et al. (2012) em- ployed a self-similar QSSA eigenvalue analysis to investigate the effect of a capil- lary transition zone above the transient boundary layer. Meanwhile, Cheng et al. (2012) investigated the effect of permeability anisotropy using the method of Slim & Ramakrishnan (2010). Because of uncertainties in methodology, comparing ex- perimental and theoretical studies is not straightforward. In addition, there are other challenges in relating theoretical and experimental studies of transient diffusive boundary layers. For example, the linear stability analyses (Ennis-King et al., 2003; Riaz et al., 2006; Selim & Rees, 2007a) of CO2 solute boundary layers employ a constant viscosity. On the other hand, experimental 8 studies (Backhaus et al., 2011), due to practical limitations, use miscible fluids with large viscosity contrast. It is therefore important to characterize the effect of fluid viscosity within the boundary layer. An important measure for subsurface flows is the permeability of the porous medium. The magnitude of permeability signifies the ability of the fluid to move with ease through the porous medium. Typical aquifers are made of sedimentary rocks with grain sizes varying from coarser sand to finer clay, which in turn causes permeability to change within position (Bear, 1988). Depending on grain shape, permeability may also be anisotropic due to preferential flow in a particular direc- tion. Currently, there is no clear understanding on the effect of a spatially varying permeability field on the onset times for natural convection in transient boundary layers. Though the effect of permeability anisotropy on transient boundary layers have been addressed by some studies (Cheng et al., 2012; Ennis-King & Paterson, 2005; Rapaka et al., 2009; Xu et al., 2006), the linear stability of transient boundary layers in an aquifer with vertically (Rapaka et al., 2009) and horizontally varying permeability is less explored. This is partly because of complications in the theoreti- cal formulation of the linear stability problem especially in the case of a horizontally varying permeability field. The current study addresses these issues by • developing a robust physical basis from which to study the influence of physical effects, such as permeability heterogeneity and anisotropy as well as chemically active boundary layers. 9 • investigating the role of viscosity on transient boundary layers, and to relate experimental and theoretical studies. • exploring the effect of vertical and horizontal variations in the permeability field. 1.4 Outline of Dissertation Chapter 2 explores the onset of natural convection in homogeneous isotropic porous media using a combination of linear stability analysis and direct numerical simulations. Various linear stability approaches are investigated simultaneously in order to develop a robust analytical framework. A new optimization procedure to analyze the onset of instabilities is proposed. Chapter 3 examines the effect of viscosity contrast on the linear stability of transient, diffusive layers in porous media. This analysis helps evaluate experimental observations of various boundary layer models that are commonly used to study CO2 sequestration. Comparisons are also made with classical viscous fingering systems. Chapter 4 explores the stability of transient diffusive boundary layers in a horizontally layered porous medium. Permeability is assumed to vary periodically along the direction of the unstable gravity force. The behavior of instability is studied with respect to modes of vorticity production related to the coupling of perturbations with the magnitude and gradients of permeability. Chapter 5 analyzes the onset of natural convection in a vertically layered porous medium. The results are obtained using multi-dimensional eigenvalue prob- 10 lems and non-linear direct numerical simulations. Finally, Chapter 6 lists the journal publications produced by this study. Av- enues for future research are suggested. 11 Chapter 2: Natural convection in homogeneous porous media 2.1 Overview In this chapter, it is shown that classical optimal perturbation profiles that maximize perturbation amplification cannot lead to onset of convection in finite time. Rather, onset of convection results from the growth of “suboptimal” pertur- bations localized within the diffusive boundary layer. The chapter is divided into three broad categories: (i) analysis of the classical optimization procedure; (ii) de- velopment of a new methodology to study constrained optimal perturbations; and (iii) validation that the new approach predicts optimal perturbations closely related to physical systems. This chapter is organized as follows. The governing equations are presented in §2.2. The classical modal and nonmodal approaches are described in §2.3. The classical optimization results are presented in §2.4. The proposed modifications to the classical optimization procedure and associated results are presented in §2.5. DNS results are reported in §2.6. The main findings are summarized in §2.7. 12 (a) (b) 0 0.2 0.4 0.6 0.8 10 0.2 0.4 0.6 0.8 1 z c b Figure 2.1: (a) Sketch, not to scale, of the geometry considered in this study. (b) Base-state (2.3) for Ra = 500 and t = 0.1 (solid line), t = 1 (dashed line), and t = 10 (dash-dotted line). 2.2 Geometry and governing equations Due to the fundamental nature of the current study, the porous medium is modeled as being fluid-saturated, isotropic, homogeneous, of infinite horizontal ex- tent in the x and y directions, and of finite depth H in the vertical z direction, see figure 2.1(a). The vertical z direction is positive in the downward direction of gravity, g. The domain is bounded by an impermeable wall at z = H. The top boundary at z = 0 represents the fixed two-phase interface between CO2 gas and CO2-saturated brine. The porous medium is characterized by its permeability, K, dispersivity, D, and porosity, φ, respectively. Initially, the brine is quiescent with zero CO2 concentration, c = 0, and constant density, ρ = ρ0. At time t = 0, satu- rated brine is supplied at z = 0 with a constant concentration c = C1 and density ρ1. The fluid viscosity, µ, is assumed to be constant. The density difference ∆ρ = ρ1−ρ0 13 is assumed to be much less than ρ0, i.e. ∆ρ ρ0. Fluid flow and mass transport in the porous medium are governed by Darcy’s law and volume averaged forms of the continuity and advection-diffusion equations (Whitaker, 1999). The governing equations are written in nondimensional form as, v +∇p− cez = 0, ∇ · v = 0, ∂c ∂t + v · ∇c− 1 Ra ∇2c = 0, (2.1) using the characteristic length L = H, time T = φH/U , buoyancy velocity U = K∆ρg/µ, pressure P = ∆ρgH, and concentration C = C1. The dimensionless equations (2.1) have been obtained using the Boussinesq approximation and a linear fluid density profile, ρ = ρ0 + ∆ρ(c/C1). The Rayleigh number is defined as Ra = UH/(φD). The symbol v = [u, v, w] is the nondimensional velocity vector, c is the nondimensional concentration and p is the nondimensional pressure obtained from the dimensional pressure pˆ through the relation p = (pˆ − ρ0gz)/P. The symbol ez is the unit vector in the z direction. Equations (2.1) must satisfy the following boundary conditions, c ∣∣∣ z=0 = 1, ∂c ∂z ∣∣∣ z=1 = 0, w ∣∣∣ z=0 = w ∣∣∣ z=1 = 0, t ≥ 0. (2.2) Equations (2.1) admit the transient base state, vb = 0, cb(z, t) = 1− 4 pi ∞∑ n=1 1 2n−1sin [( n− 1 2 ) piz ] exp [ − ( n− 1 2 )2 pi2t Ra ] , (2.3) The linear stability of base-state (2.3) is studied with respect to small wavelike perturbations of the form, c˜ = ĉ(z, t)ei(αx+βy), v˜ = v̂(z, t)ei(αx+βy), p˜ = p̂(z, t)ei(αx+βy), (2.4) 14 where i = √ −1, α and β are wavenumbers in the x and y directions respectively, and ĉ(z, t), v̂(z, t) and p̂(z, t) are time-dependent perturbation profiles in the z direction. Following the standard procedure (see Riaz et al., 2006), the linear stability problem can be written as the following initial value problem for ĉ and ŵ, ∂ĉ ∂t + ŵ ∂cb ∂z − 1 Ra Dĉ = 0, Dŵ + k2ĉ = 0, (2.5) ĉ ∣∣∣ z=0 = 0, ∂ĉ ∂z ∣∣∣ z=1 = 0, ŵ ∣∣∣ z=0 = ŵ ∣∣∣ z=1 = 0, (2.6) where D = ∂2/∂z2 − k2 and k = √ α2 + β2. Because the base-state is transient, the boundary layer is sensitive to the time at which it is perturbed. The boundary layer is perturbed at time t = tp with the following initial perturbation profiles, ĉ ∣∣∣ t=tp = cp(z), ŵ ∣∣∣ t=tp = wp(z), (2.7) where cp and wp must satisfy equations (2.5)–(2.6). 2.3 Linear stability methods This study explores a wide range of complementary modal and nonmodal analysis methods. The various methods used for the linear stability analysis of equations (2.5)–(2.6) are discussed in this section. 2.3.1 Quasi-steady eigenvalue analysis The QSSA approach reduces equations (2.5)–(2.6) to an autonomous eigen- value problem. For a prescribed final time tf , this approach approximates the base- state, cb(z, t), as steady and decomposes perturbations into separable functions of 15 z and t, ĉ = ce(z; tf)e σ(tf)t, ŵ = we(z; tf)e σ(tf)t, (2.8) where σ(tf) is the instantaneous growth rate at t = tf . Substituting (2.8) into (2.5)– (2.6) produces an eigenvalue problem for eigenvalues σ and eigenfunctions ce and we, 1 Ra ( d2 dz2 − k2 ) ce + k 2∂cb ∂z ( d2 dz2 − k2 )−1 ce = σce, ce ∣∣∣ z=0 = dce dz ∣∣∣ z=1 = 0. (2.9) The QSSA eigenvalue problem can also be written in terms of we. In practice, however, solving (2.9) was more stable. 2.3.2 Classical Optimization The initial perturbation profiles, cp and wp, are determined so that the per- turbation amplification is maximized at some prescribed final time t = tf . Tan & Homsy (1986) and Doumenc et al. (2010) have observed that the perturbation amplification is sensitive to the perturbation flow field used to measure the pertur- bation magnitude. To investigate how different measures of perturbation magnitude influence nonmodal results, the perturbation magnitude at time t is defined as, E(t) = ∫ 1 0 [ A1ĉ(z, t) 2 + A2ŵ(z, t) 2 + A3û(z, t) 2 ] dz, (2.10) where A1,A2 and A3 are constants to be defined shortly. The following measures of perturbation amplification, Φ(t), are studied, Φc(t) = [ E(t) E(tp) ] 1 2 , A1 = 1, A2 = A3 = 0, (2.11a) 16 Φw(t) = [ E(t) E(tp) ] 1 2 , A2 = 1, A1 = A3 = 0, (2.11b) Φe(t) = [ E(t) E(tp) ] 1 2 , A1 = A2 = A3 = 1. (2.11c) Most previous studies of transient boundary layers measure amplification with re- spect to the perturbation’s concentration field, Φc (Caltagirone, 1980; Ennis-King et al., 2003; Kim & Kim, 2005; Rapaka et al., 2008; Tan & Homsy, 1986), or the vertical velocity field, Φw (Foster, 1965; Gresho & Sani, 1971; Kaviany, 1984; Tan & Homsy, 1986). In addition, Φe is introduced as a measure of perturbation energy that includes both the velocity and concentration fields. Φ(tf) is optimized using an adjoint procedure described by Doumenc et al. (2010) in which E(tf) is maximized subject to the constraint that E(tp) = 1. For this purpose, the Lagrangian is defined as, L(ĉ, c∗, ŵ, w∗, û, s) = E(tf)− s [ E(tp)− 1 ] − ∫ tf tp ∫ 1 0 w∗ ( Dŵ + k2ĉ ) dz dt − ∫ tf tp ∫ 1 0 c∗ ( ∂ĉ ∂t − 1 Ra Dĉ+ ŵ∂cb ∂z ) dz dt, (2.12) where s is a scalar Lagrange multiplier and the adjoint variables c∗(z, t) and w∗(z, t) are Lagrange multipliers dependent on z and t. The double integrals on the right- hand-side of (2.12) assure satisfaction of the governing IVP (2.5)–(2.7). To obtain first-order optimality conditions, the variational of the Lagrangian, δL, is set to 17 zero. Integrating by parts, δL can be written as, δL = ∫ 1 0 [ 2 (A1ĉ δĉ+ A2ŵ δŵ + A3û δû)− c∗ δĉ ] t=tf dz − ∫ 1 0 [ 2s (A1ĉ δĉ+ A2ŵ δŵ + A3û δû)− c∗ δĉ ] t=tp dz − ∫ tf tp ∫ 1 0 [ δĉ ( −∂c ∗ ∂t − 1 Ra Dc∗ + k2w∗ ) + δŵ ( Dw∗ + ∂cb ∂z c∗ )] dz dt + ∫ tf tp [ 1 Ra ( c∗ ∂δĉ ∂z − δĉ∂c ∗ ∂z ) − w∗∂δŵ ∂z + δŵ ∂w∗ ∂z ]z=1 z=0 dt = 0. (2.13) The optimality conditions are met when c∗ and w∗ satisfy the following adjoint problem, −∂c ∗ ∂t − 1 Ra Dc∗ + k2w∗ = 0, Dw∗ = −∂cb ∂z c∗, (2.14) c∗ ∣∣∣ z=0 = 0, ∂c∗ ∂z ∣∣∣ z=1 = 0, w∗ ∣∣∣ z=0 = w∗ ∣∣∣ z=1 = 0, (2.15) along with the following coupling conditions between the adjoint and physical vari- ables, 2 (A1ĉ δĉ+ A2ŵ δŵ + A3û δû) ∣∣∣ t=tf = c∗ δĉ ∣∣∣ t=tf , (2.16) 2s (A1ĉ δĉ+ A2ŵ δŵ + A3û δû) ∣∣∣ t=tp = c∗ δĉ ∣∣∣ t=tp . (2.17) The optimal initial perturbations are found using an iterative procedure. Given an initial guess for cp and wp, the IVP (2.5)–(2.7) is integrated forward in time to t = tf . Then the condition (2.16) is used to obtain a final condition for the adjoint IVP (2.14)–(2.15). The adjoint IVP is then integrated backwards in time to t = tp. Then using condition (2.17), one obtains improved initial profiles cp and wp. This procedure is repeated until satisfaction of the convergence criteria, ‖cnp − cn−1p ‖∞/‖cn−1p ‖∞ ≤ 10−4, where n is the iteration number. The iterative procedure is insensitive to the initial guess; however, the number of iterations is 18 reduced using c0p = ξexp(−ξ2) where ξ = z √ Ra/(4t). The IVPs are solved using standard second-order finite-difference methods. The application of the coupling conditions (2.16)–(2.17) depends on the def- inition of the perturbation amplification. When maximizing Φc, conditions (2.16)– (2.17) are satisfied for, 2ĉ ∣∣∣ t=tf = c∗ ∣∣∣ t=tf , 2sĉ ∣∣∣ t=tp = c∗ ∣∣∣ t=tp . (2.18) The derivation of (2.18) is described by Doumenc et al. (2010). When maximizing Φw or Φe, however, the application of the coupling conditions is less straightforward than in the case of Doumenc et al. (2010) because in the current study, the momen- tum equation lacks a temporal derivative. For those cases, the Neumann boundary conditions for ĉ and c∗ at the lower wall are replaced with the following Dirichlet condition, ĉ ∣∣∣ z=1 = c∗ ∣∣∣ z=1 = 0. (2.19) Consequently, when maximizing Φw, coupling conditions (2.16)–(2.17) are satisfied for, −2k2ŵ ∣∣∣ t=tf = Dc∗ ∣∣∣ t=tf , −2k2sŵ ∣∣∣ t=tp = Dc∗ ∣∣∣ t=tp . (2.20) When maximizing Φe, conditions (2.16)–(2.17) are satisfied for, ( k2 ∂2 ∂z2 − k4 −D2 ) ŵ ∣∣∣ t=tf = k2 2 Dc∗ ∣∣∣ t=tf , (2.21) s ( k2 ∂2 ∂z2 − k4 −D2 ) ŵ ∣∣∣ t=tp = k2 2 Dc∗ ∣∣∣ t=tp . (2.22) For the parameter space (tp, tf ,Ra, k) considered in the current study, the Dirichlet condition (2.19) is a valid approximation because the optimal perturbations 19 are concentrated near z = 0 and decay to zero outside the boundary layer such that they are not influenced by the lower wall (Rapaka et al., 2008; Slim & Ramakrishnan, 2010). The results are validated by directly optimizing the IVP (2.5)–(2.7), subject to the standard boundary conditions (2.6), using MATLAB routines. The adjoint- based method shows excellent agreement with direct optimization but is an order- of-magnitude faster. 2.4 Results Previously, Rapaka et al. (2008) reported optimal perturbations that maximize Φc for a fixed initial perturbation time, tp. That work is extended in the following manner. First, this study explores how the amplification measure (2.11) affects the optimization results. Second, the role of the initial perturbation time is investigated. Third, this study examines the influence of the final time on the initial optimal profiles. Fourth, in this study, the optimal perturbations are compared with quasi- steady eigenmodes. Finally, in §2.3, the current study assess the relevance of the optimal perturbations to physical experiments. 2.4.1 Effect of amplification measure Figure 2.2 presents optimization results for Ra = 500 and tp = 0.01 when maximizing Φc, Φw, and Φe. The Rayleigh number is set to a typical value for CO2 sequestration (Ennis-King & Paterson, 2005). The initial perturbation time is chosen to be one order-of-magnitude smaller than the critical time for instability, 20 (a) (b) 0 10 20 30 40 500 1 2 3 4 5 k t f 1 1e4 1e3 100 10 0.1 0.2 10 5 10 15 20 25 30 tf k max (c) (d) 0 0.05 0.1 0.150 0.2 0.4 0.6 0.8 1 z c b(z, t p), c p(z) 0 0.5 1 1.5 100 101 t Φ c (e) (f ) 0 0.5 1 1.5 100 101 t Φ w 0 0.5 1 1.5 100 101 t Φ e Figure 2.2: Optimization results for Ra = 500 and tp = 0.01 when maximizing Φc, Φw and Φe, respectively. (a) Isocontours of Φc (solid line), Φw (dashed line), and Φe (dash-dotted line) in the (k, tf) plane. (b) kmax, vs. tf , when maximizing Φc (solid line), Φw (dashed line), and Φe (dash-dotted line). (c) The optimal cp profiles when maximizing Φc (circles), Φw (crosses), and Φe (squares) for k = 30 and tf = 5. The base state cb(z, tp) is shown as a solid line. (d)–(f ) Amplifications Φc, Φw, and Φe vs. t when integrating the forward IVP (2.5)–(2.7) in time using the optimal initial cp profiles shown in panel (c), that maximize Φc (solid line), Φw (dashed line), and Φe (dash-dotted line). 21 tc, where the critical time is the time at which dΦ/dt = 0, after which Φ begins to increase. The critical time depends on the initial condition and choice of the amplification measure; however, previous analyses find the minimum critical time is on the order of tc ∼ O(0.1) for Ra = 500 (Riaz et al., 2006; Selim & Rees, 2007a; Slim & Ramakrishnan, 2010). Panel (a) illustrates optimal isocontours of Φc (solid lines), Φw (dashed lines), and Φe (dash-dotted lines) in the (k, tf) plane. The three amplification measures produce qualitatively similar behavior. The isocontours for Φc and Φe are visually indistinguishable. For much of the (k, tf) plane, Φw is marginally larger than Φc or Φe. The dominant wavenumber, kmax, is defined as the wavenumber for which the amplification is maximized at tf , Φmax(tf) = sup 0≤k<∞ Φ(tf , k). (2.23) Figure 2.2(b) illustrates the dominant wavenumbers that maximize Φc (solid line), Φw (dashed line), and Φe (dash-dotted line) for the final times 0.1 ≤ tf ≤ 2. The dominant wavenumbers for the three amplification measures are qualitatively sim- ilar. When tf ≤ 0.21, the dominant wavenumbers are zero. When tf > 0.21, the dominant wavenumbers jump discontinuously to values around kmax ≈ 25. The dominant zero-wavenumber perturbations were not reported by Rapaka et al. (2008) because they considered late values of tf for which kmax is non-zero. When comparing results of Rapaka et al. (2008) with the current study, one must note that Rapaka et al. (2008) nondimensionalized the problem with a diffusive time scale, while the current study uses an advective time scale. Consequently, the nondimensional times, 22 t, in this study are related to those of Rapaka et al. (2008), t(R), through the relation t(R) = t/Ra. Though maximizing different perturbation fields produces similar dominant wavenumbers, kmax, the corresponding optimal initial profiles, cp and wp, are sen- sitive to the amplification measure. Figure 2.2(c) illustrates the optimal cp profiles that maximize Φc (circles), Φw (crosses), and Φe (squares) at tf = 5 for k = 30. For visualization, the profiles have been scaled so ‖cp‖∞ = 1. The solid line shows the base-state at tp = 0.01. Figure 2.2(c) shows results for 0 ≤ z ≤ 0.15 because the profiles are concentrated near z = 0 and decay to zero before interacting with the lower wall z = 1. The profiles for Φe and Φc have maxima occurring around z = 0.05, while the profile for Φw has a maximum occurring around z = 0.01. Figure 2.2(d) illustrates the temporal evolution of Φc when the forward IVP (2.5)–(2.7) is integrated from tp = 0.01 to t = 2 using the three initial cp profiles illustrated in figure 2.2(c). The cp profiles that maximize Φc (solid line) and Φe (dash-dotted line) produce indistinguishable results in figure 2.2(d). The cp pro- file that maximizes Φw (dashed line), however, produces much lower values of Φc. This suggests that maximization of Φw occurs at the expense of Φc. Figure 2.2(e) illustrates the corresponding results for the evolution of Φw. The cp profiles that maximize Φc (solid line) and Φe (dash-dotted line) produce nearly indistinguishable results, while the profile that maximizes Φw (dashed line) produces marginally larger Φw. Finally, figure 2.2(f ) illustrates the corresponding results for the evolution of Φe. The initial profiles that maximize Φc and Φe again produce indistinguishable results. This indicates that maximizing the perturbation’s concentration field nat- 23 urally maximizes Φe, while maximizing Φw does so at the expense of Φc and Φe. Because Φc naturally maximizes Φe, hereinafter this study focusses on maximiz- ing Φc. Φc is preferred over Φe because the application of the coupling conditions (2.16)–(2.17) is much simpler for Φc. 2.4.2 Optimal perturbation structures Figure 2.3(a) illustrates the optimal amplifications Φc versus tf for tp = 0.01, and k = 0 (circles), k = 10 (crosses), k = 25 (squares), and k = 40 (diamonds). For small final times, tf < 0.1, all perturbations decay; however, the k = 25 perturba- tions are more damped than the k = 0 and k = 10 perturbations. Note that the k = 0 perturbations have a small constant damping rate. This occurs because the IVP (2.5)–(2.7) for k = 0 reduces to ∂ĉ ∂t − 1 Ra ∂2ĉ ∂z2 = 0, ŵ = 0. (2.24) Equation (2.24) can be solved analytically to show that the optimal perturbation is given by ĉ = sin (piz/2) exp (−pi2Ra−1t/4). In contrast to the k = 0 perturbations, finite wavenumber perturbations do not have constant growth rates. The k = 25 perturbations begin to grow around tf = 0.1 and eventually overtake the k = 10 and k = 0 perturbations. This explains the discontinuous jump in the dominant wavenumbers from kmax = 0 to kmax ≈ 25 illustrated in figure 2.2(b). The k = 40 perturbations experience greater damping, and consequently, never overtake the k = 25 perturbations. Figure 2.3(b) illustrates the base state, cb(z, tf) (solid line), and optimal pro- 24 (a) (b) 0 0.1 0.2 0.3 0.4 0.50.85 0.9 0.95 1 1.05 tf Φ c 0 0.2 0.4 0.6 0.8 10 0.2 0.4 0.6 0.8 1 z c b(z, t f), ĉ(z,t f) Figure 2.3: Dominant perturbations for Ra = 500 and tp = 0.01. (a) Φc vs. tf , for k = 0 (circles), k = 10 (crosses), k = 25 (squares), and k = 40 (diamonds). (b) cb(z, tf) (solid line) and ĉ(z, tf) for tf = 0.21, and k = 0 (circles), k = 10 (crosses), k = 20 (squares), and k = 30 (diamonds). files, ĉ(z, tf), at tf = 0.21 for k = 0 (circles), k = 10 (crosses), k = 20 (squares), and k = 30 (diamonds). The final time is chosen to be near to the discontinuous jump in kmax illustrated in figure 2.2(b). The optimal profile for k = 0 has a maximum at the lower boundary at z = 1, while the profiles for k = 10, 20, and 30 have maxima near z = 0. With increasing k, the optimal profiles become increasingly concentrated within the boundary layer. The results for Φc and cp illustrated in figure 2.3 can be explained physically by examining the competing effects of the stabilizing diffusive term, Dĉ/Ra, and the destabilizing convective term, ŵ∂cb/∂z, in equation (2.5). At small times, t  tc, the convective term ŵ∂cb/∂z has only a small effect because ∂cb/∂z is nonzero only within the thin boundary layer where ŵ necessarily tends to zero due to the no- penetration condition at z = 0. This explains why the boundary layer is stable at small times. The dominant wavenumber is initially zero because finite wavenumber 25 perturbations have additional damping due to the transverse diffusive term (k2/Ra)ĉ in equation (2.5). At later times, the growing boundary layer increases the influence of the destabilizing term ŵ∂cb/∂z such that non-zero wavenumber perturbations become unstable. This explains why dominant perturbations at late times tend to be increasingly concentrated in the boundary layer. 2.4.3 Sensitivity to initial perturbation time Due to the transient nature of the base-state, the optimal perturbations also depend on the time, tp, at which the boundary layer is perturbed. Figure 2.4 explores the sensitivity of the optimal amplifications Φc to the initial perturbation time tp for Ra=500. Panel (a) illustrates Φc versus tf for k = 30 and tp = 0.001 (solid line), tp = 0.1 (dashed line), and tp = 0.5 (dash-dotted line). Perturbations originating at tp = 0.001 have a long initial damping period and consequently have smaller amplifications than perturbations originating at tp = 0.1. Perturbations originating at the late time tp = 0.5 experience no damping, but have smaller amplifications than perturbations originating at tp = 0.001 and tp = 0.1 because those perturba- tions begin growing much earlier. At later times, tf > 0.5, the three curves have identical slopes, indicating that the perturbations have identical temporal growth rates. Figure 2.4(b) illustrates isocontours of Φc in the (k, tf) parameter plane for tp = 0.001 (solid line), tp = 0.1 (dashed line), and tp = 0.5 (dash-dotted line). As expected, perturbations originating at tp = 0.1 produce larger amplifications. The horizontal dash-dotted line indicates that perturbations originating at tp = 0.5 grow 26 immediately for 2 < k < 56. Figures 2.4(a) and 2.4(b) suggest that there exists an optimal initial pertur- bation time, top, that maximizes Φc. Perturbations originating prior to t o p cannot outgrow the optimal perturbation originating at top due to the initial damping pe- riod. From figure 2.4(a), top is expected to occur near the critical time, t = tc, because this minimizes the damping period. Note that for Ra = 500, Slim & Ramakrish- nan (2010) report that the minimum critical time is tc ≈ 0.096. The notion of an optimal initial perturbation time may appear counterintuitive because in physical systems the boundary layer is continuously perturbed beginning at tp = 0. Within the framework of a linear stability analysis, however, the response to this contin- uous forcing can be expressed as the infinite sum of many impulse responses to forcing at discrete initial times, tp. The optimal perturbation originating at top gives a theoretical upper bound for the amplification. Figure 2.4(c) illustrates the normalized amplifications, Φc/||Φc||∞, versus tp for tf = 1, and k = 10 (solid line), k = 30 (dashed line), and k = 50 (dash-dotted line). The amplifications have been normalized with respect to their maximum values to facilitate comparison between the results for different wavenumbers. As tp → 0, the amplifications asymptote to constant values. With increasing tp, the amplifications attain maxima near tp = tc and then decrease. Stronger sensitivity of Φc to tp occurs with increasing wavenumber. This behavior is similar to that observed in figure 2.3(a) for the sensitivity of Φc to the final time tf . The increasing sensitivity of Φc to both tp and tf at higher wavenumbers is likely due to the increase in transverse diffusive damping as noted in the previous section. Figure 2.4(d) 27 (a) (b) 0 0.2 0.4 0.6 0.8 1 100 tf Φ c 0 10 20 30 40 500 1 2 3 4 5 k t f 1 10100 1e3 (c) (d) 10−4 10−3 10−2 10−10.65 0.7 0.75 0.8 0.85 0.9 0.95 1 tp Φ c/|| Φ c|| ∞ 10−4 10−3 10−2 10−10.85 0.9 0.95 1 tp Φ c/|| Φ c|| ∞ Figure 2.4: Effect of initial perturbation time for Ra = 500. (a) Φc vs. tf for k = 30, and tp = 0.001 (solid line), tp = 0.1 (dashed line), and tp = 0.5 (dash-dotted line). (b) Isocontours of Φc in the (k, tf) plane for tp = 0.001 (solid line), tp = 0.1 (dashed line), and tp = 0.5 (dash-dotted line). (c) Φc/||Φc||∞, vs. tp for tf = 1, and k = 10 (solid line), k = 30 (dashed line), k = 50 (dash-dotted line). (d) Φc/||Φc||∞ vs. tp for k = 30 and tf = 1 (circles), tf = 2 (crosses), and tf = 3 (squares). 28 (a) (b) 0 1 2 3 4 50 5 10 15 20 25 30 tf k max 0 0.05 0.1 0.15 0.20 0.2 0.4 0.6 0.8 1 z c b(z, t p), c p(z) (c) (d) 0 0.05 0.1 0.15 0.2 0.25 0.3 0.350 0.2 0.4 0.6 0.8 1 z c b(z, t p), c p(z) 0 0.05 0.1 0.15 0.2 0.25 0.3 0.350 0.2 0.4 0.6 0.8 1 z c b(z, t p), c p(z) Figure 2.5: (a) Dominant wavenumbers, kmax vs. tf for tp = 0.001 (solid line), tp = 0.1 (dashed line), and tp = 0.5 (dash-dotted line). (b) Base- state, cb (solid line), and optimal cp profiles (dashed line) for tp = 0.001, tf = 5, and k = 30. (c) Same as panel (b) for tp = 0.5. (d) Same as panel (b) for tp = 1.5. With increasing tp, the cp profiles become increasingly concentrated in the boundary layer. 29 illustrates Φc/||Φc||∞ versus tp for k = 30 and tf = 1 (circles), tf = 2 (crosses), and tf = 3 (squares). The results for different tf are indistinguishable from each other. This occurs because, as demonstrated in figure 2.4(a), the perturbations have identical growth rates for tf > 1. Figure 2.5(a) illustrates the temporal evolution of the dominant wavenumbers, kmax, when tp = 0.001, (solid line), tp = 0.1 (dashed line), and tp = 0.5 (dash-dotted line). As expected from the discussion in §2.4.2, the dominant wavenumbers are initially zero when tp = 0.001. When tp = 0.1, however, the dominant wavenumber is initially kmax = 29.74 for tf = 0.12 and reaches a maximum at tf = 0.26 after which it decays monotonically. When tp = 0.5, kmax decreases monotonically with tf . Previously, Rapaka et al. (2008) only reported cases with a monotonic decay of kmax with tf . Figures 2.5(b)–2.5(d) illustrate the base state (solid lines) and optimal cp profiles (dashed lines), for tp = 0.001 (panel b), tp = 0.5 (panel c), and tp = 1.5 (panel d) for Ra = 500, k = 30, and tf = 5. As expected from the discussion in §2.4.2, the optimal profiles become increasingly concentrated within the boundary layer with increasing tp due to the destabilizing convective term. To explore the optimal initial perturbation time, the optimization procedure is repeated for a wide range of wavenumbers, initial times, and final times. Figure 2.6(a) illustrates the optimal amplifications Φc for tf = 1, Ra = 500, 10 ≤ k ≤ 50, and 0.01 ≤ tp ≤ 0.5. The maximum amplification, i.e. the peak of the Φc surface in figure 2.6(a), is defined as Φoc(tf) = sup 0≤k<∞ 0 1, the amplifications produced by COP and MOP in figure 2.12(b) have identical slopes. This occurs because of similar growth rates between the final states of the dominant wavenumber perturbations obtained using the COP and MOP schemes. The difference between the COP and MOP amplifications depends on the initial time, tp. To explore this, one can measure ∆Φmax = ΦCOP − ΦMOP ΦCOP , (2.40) where ΦCOP and ΦMOP are the maximum amplifications, Φmax, obtained using COP and MOP, respectively. Figure 2.12(c) illustrates ∆Φmax for tf = 4 as the initial perturbation time varies from tp = 10−3 to tp = 1. Note that the results are independent of final time tf when tf > 1. ∆Φmax tends to a maximum as tp → 0 because ΦMOP → 0, while ΦCOP converges to finite values, see figures 2.4(c)–2.4(d). With increasing tp, ∆Φmax decreases indicating better agreement between the COP and MOP amplifications. Note that the maxima of the optimal initial MOP profiles are always closer to the top boundary, z = 0, compared to the initial COP profiles. Recall from §2.4.4, that beyond certain final times, the initial cp profiles gen- erated by the COP scheme are insensitive to further increases in tf . To investi- gate this behavior for the MOP scheme, figure 2.12(d) illustrates the isocontours ∆cp/∆tf = 0.001, see equation (2.31), in the (k, tf) parameter plane generated using the COP (solid line) and MOP (dashed line) schemes for tp = 0.1 and Ra = 500. The 47 final times beyond which the initial MOP profiles do not change shape are much smaller than the COP profiles. This suggests that initial perturbations confined within the boundary layer rapidly converge to a common shape. As discussed in §2.4.3, due to the transient growth of the base-state, there exists an optimal combination of initial time and wavenumber, top and k o, that produces the subsequent optimal amplification Φoc. Figure 2.13(a) illustrates the MOP amplifications, Φc, for 10 ≤ k ≤ 50, 0.1 ≤ tp ≤ 0.5, tf = 1, and Ra = 500. The solid dot marks the peak of the surface, Φoc. Figure 2.13 demonstrates Φ o c (panel b), ko/Ra (panel c), and topRa (panel d) as functions of tfRa. The results for different Ra collapse as previously demonstrated for the COP scheme in figure 2.6. The following relationships for Φoc and the dimensional forms of wavenumber, k∗, and the initial time, t∗p are obtained, log Φoc = −5.550×10−8t∗f 2 ( U2 φ2D )2 + 0.001785t∗f U2 φ2D − 0.3967, (2.41) k∗ = U φD [ 0.1234− 0.02237 log ( t∗fU 2 φ2D )] , (2.42) t∗p = −5.107×10−7t∗f 2 U2 φ2D + 0.01086t∗f + 120.1 φ2D U2 . (2.43) For high permeability aquifers, K = 10−12 m2 (see §2.4.3), figure 2.13 predicts that the optimal perturbation wavelength and initial perturbation time vary in the range, 10 cm ≤ 2pi/k∗ ≤ 18 cm and 36 hours ≤ t∗p ≤ 51 hours as the final time varies between, 6 days ≤ t∗f ≤ 96 days. For low permeability aquifers, K = 10−14 m2, these parameters vary in the range 10 m ≤ 2pi/k∗ ≤ 18 m, 41 years ≤ t∗p ≤ 58 years, 165 years ≤ t∗f ≤ 2636 years. The optimal amplifications, Φoc, are approximately 50 % those produced by the COP scheme, see figure 2.6. ko agrees closely with 48 (a) (b) 10 20 30 40 500.20.4 0.5 1 1.5 2 2.5 k Φoc tp Φc tfRa Φo c 0 2000 4000 6000 8000100 105 1010 (c) (d) tfRa ko /R a 500 1000 2000 4000 8000 0.04 0.045 0.05 0.055 0.06 0.065 tfRa to pRa 0 2000 4000 6000 8000120 130 140 150 160 170 180 Figure 2.13: The optimal MOP point (Φoc, k o, top) as a function of tf and Ra. (a) Φc vs. tp and k for Ra = 500 and tf = 1. The solid dot marks (Φoc, k o, top). (b) Φ o c vs. tfRa for Ra = 500 (circles), Ra = 750 (crosses), and Ra = 1000 (squares). The dashed line shows relationship (2.41). (c) ko/Ra vs. tfRa for Ra = 500 (circles), Ra = 750 (crosses), and Ra = 1000 (squares). The dashed line shows relationship (2.42). (d) tpRa vs. tfRa for Ra = 500 (circles), Ra = 750 (crosses), and Ra = 1000 (squares). The dashed line shows relationship (2.43). 49 those produced, using the COP scheme. The optimal initial perturbation times, top, however, are roughly twice as large as those for the COP scheme due to the large initial damping periods experienced by the MOP perturbations. The optimal initial time, top, is also more sensitive to tf than the COP scheme. Recall from §2.4.3, that the optimal initial perturbation time would require a priori knowledge of the onset time of convection, i.e. tf = to. Because of the increased sensitivity of top to tf , the optimal MOP perturbations are expected to be more sensitive to initial perturbation amplitude, A∞, than the COP perturbations. 2.5.4 Comparison with eigenvalue and initial value problems In this section, the modified optimization procedure is compared to previously published linear stability methods that ensure perturbations are localized within the boundary layer. The first approach approximates the vertical domain as semi- infinite. In this case there is a similarity solution for the base-state, cb = 1− erf(ξ), where ξ(z, t) = z √ Ra/(4t) is the similarity variable. Riaz et al. (2006) demon- strated that a quasi-steady modal analysis with respect to the (ξ, t) space produces eigenmodes concentrated in the boundary layer. For convenience of notation, this eigenvalue problem is referred to as the QSSAξ problem. The eigenvectors of the QSSAξ problem are cξ and wξ. The second procedure considered is the solution of the forward IVP (2.5)–(2.7) using cp = cdm, where cdm is the “dominant mode” of Riaz et al. (2006) given by, cdm(z) = ξe −ξ2 . (2.44) 50 (a) (b) 0 0.005 0.01 0.015 0.02 0.025 0.030 0.2 0.4 0.6 0.8 1 z c b, c p, cξ , c dm 0 0.005 0.01 0.015 0.02 0.025 0.030 0.2 0.4 0.6 0.8 1 c b, c p, cξ , c dm z (c) (d) 0 1 2 3 4 510−1 100 101 102 103 Φ max tf 0 1 2 3 4 5 30 40 50 tf k max Figure 2.14: Comparison of modified optimization with QSSA in self- similar space and IVP with cp = cdm for Ra = 500. (a) Base-state, cb (solid line), initial MOP profile (circles), dominant QSSAξ eigenmode, c ξ (squares), and cdm perturbation (2.44) (crosses) at tp = 0.01 for k = 10. (b) Same as in panel (a) for k = 50. (c)–(d) Temporal evolution of Φmax and kmax for tp = 0.01 using MOP (solid line), QSSAξ (dashed line), and initial condition (2.44) (dash-dotted line). 51 Initial condition (2.44) is the leading-order term of a Hermite polynomial expansion in the (ξ, t) space and has been used in numerous previous studies (Ben et al., 2002; Elenius et al., 2012; Kim & Choi, 2011, 2012; Pritchard, 2004; Riaz et al., 2006; Selim & Rees, 2007a; Wessel-Berg, 2009). Figure 2.14(a) illustrates the initial perturbation concentration profiles, ĉ(z, tp), produced by the MOP (circles), dominant QSSAξ eigenmode (squares), and initial condition (2.44) (crosses), for tp = 0.01, Ra = 500, and k = 10. Figure 2.14(b) repeats figure 2.14(a) for the larger wavenumber, k = 50. In both figures, the base-state is shown as a solid line. For both wavenumbers, the three methodologies produce qualitatively similar profiles. The profiles produced by QSSAξ and cdm are indistinguishable, while the MOP profiles have maxima closer to z = 0. Note that the MOP profiles support slightly larger initial magnitudes, A∞, than the QSSAξ and cdm profiles, without producing negative net concentrations, cnet. Figure 2.14(c) illustrates results for Φmax versus tf obtained using the MOP (solid line), QSSAξ (dashed line), and initial condition (2.44) for tp = 0.01, 0.03 ≤ tf ≤ 5, and Ra = 500. The three procedures again produce similar results, though initial condition (2.44) produces marginally larger amplifications. Note that the QSSAξ amplifications are obtained by first transforming the dominant QSSAξ growth rates to the (z, t) coordinates using a L2 norm, before integrating equation (2.33). Figure 2.14(d) illustrates the corresponding dominant wavenumbers, kmax, of the three procedures. The results produced by initial condition (2.44) and the MOP are indistinguishable. 52 2.6 Direct Numerical Simulations Two-dimensional direct numerical simulations (DNS) of the nonlinear govern- ing equations (2.1)–(2.2) are performed using a traditional pseudospectral method with spectral spatial accuracy (Peyret, 2002). The horizontal domain is truncated to x ∈ [0, L] with periodic boundary conditions on x = 0 and x = L. Equa- tions (2.1)–(2.2) are then discretized spatially using Chebyshev polynomials in the vertical z direction and a Fourier expansion in the horizontal x direction. The advection-diffusion equation is discretized temporally using a third-order, semi- implicit, backwards-difference scheme (Peyret, 2002). This temporal discretization is chosen for its favorable stability and allows us to investigate small initial times, tp → 0, The initial concentration field is prescribed at t = tp as cdns(z, x) = cb(z) + A∞ ci(x, z) ‖ci‖∞ , (2.45) where A∞ is the initial perturbation magnitude measured with respect to the infinity norm of the perturbation concentration field, ci. 2.6.1 Simulations of physical systems To emulate physical experiments, DNS is performed in which the boundary layer is simultaneously perturbed with all wavenumbers resolved numerically, ci(x, z) = N/2−1∑ m=0 amcos ( 2pim L x ) G(z)F (z), (2.46) where N is the number of collocation points in the x direction, and −1 ≤ F (z) ≤ 1 is a random function generated using Fortran’s random number generator. The co- 53 Case ζc σ Symbol 1 0.50 0.05 Circle 2 0.50 0.10 Square 3 0.50 0.15 Cross 4 0.25 0.10 Diamond 5 0.75 0.10 Plus Table 2.2: The parameters used for the Gaussian, G(z). efficients am are computed to ensure that each horizontal Fourier mode is perturbed with equal energy. The following values, L = 4pi and N = 1024, are used in order to resolve wavenumbers, k = 0, 0.5, 1, ... , 255. To ensure that ci satisfies the bound- ary condition at z = 0 and remains concentrated within the boundary layer, the Gaussian function is employed, G(z) =    0 if z = 0, exp ( −12 ( ζ−ζc σ )2) if 0 < z ≤ δ, 0 if δ < z ≤ 1, (2.47) where ζ = z/δ, ζc is the mean and σ is the standard deviation. For example, when ζc = 0.5, the peak of the Gaussian function is located midway between z = 0 and z = δ. This study varies the peak location, ζc, and the width, σ, to recreate several experimental possibilities listed in table 2.2. Figure 2.15(a) illustrates the temporal evolution of the dominant wavenum- 54 (a) (b) 0 1 2 3 4 50 10 20 30 40 50 t k max 0 1 2 3 4 5 25 30 35 40 45 t k max Figure 2.15: (a) Temporal evolution of dominant wavenumbers, kmax, produced by DNS (symbols, see table 2.2), COP (solid line), and MOP (dashed line) for Ra = 500 and tp = 0.01 (b) Same as in panel (a) for tp = 0.2. bers, kmax, produced by COP (solid line), MOP (dashed line), and five DNS recre- ating the experimental conditions in table 2.2, for Ra = 500 and tp = 0.01. All simulations are run using the initial amplitude A∞ = 10−4 to produce a long linear regime, to > 5, to facilitate comparison of the dominant wavenumbers predicted by COP, MOP and DNS. Excellent agreement between the dominant wavenumbers produced by the MOP and DNS are observed, while those predicted by the COP show poor agreement. Figure 2.15(b) repeats figure 2.15(a) for the initial perturbation time tp = 0.2, chosen to be near the optimal perturbation time, top. Note that the DNS results for kmax have a much wider spread than those for tp = 0.01. This likely occurs because the initial damping period is much shorter for tp = 0.2. Overall, MOP shows much better agreement with DNS than COP. For cases 1, 2, and 3 (see table 2.2) the agreement between MOP and DNS is excellent. For cases 4 and 5, MOP 55 underpredicts kmax, though it still outperforms COP. The improved agreement for cases 1, 2, and 3 may stem from the fact that the boundary layer was perturbed near z = 0.5δ in these cases. In cases 4 and 5, the layer was perturbed near z = 0.25δ and z = 0.75δ, respectively. 2.6.2 Extent of linear regime and onset of convection In the remaining section, it is shown that there exists a well-defined linear regime preceding onset of convection. The onset time to is measured for different values of A∞ and tp by specifying the following initial concentration field, cdns(x, z) = cb(z) + A∞ cos(kx) cp(z) ||cp||∞ , (2.48) where cp are the optimal initial profiles determined by COP or MOP. Motivated by experiments (Blair & Quinn, 1969; Kaviany, 1984), the current study defines to as the time at which dJ/dt = 0, where J is the mean flux of CO2 into the brine given by, J(t) = − 1 L ∫ L 0 1 Ra ∂cdns ∂z ∣∣∣ z=0 dx. (2.49) Note from (2.49) that perturbations oscillating sinusoidally in the horizontal direc- tion have no net effect on J . Consequently, during the linear regime, the net flux is due to pure diffusion of the base-state, i.e. J = Jb. The deviation of the DNS results for J from Jb is due to the growth of a zero-wavenumber mode, k = 0, due to nonlinear interactions (Jhaveri & Homsy, 1982). To further quantify the duration of the linear regime, this study also measures the time, t = tl, for which J/Jb = 1.01. Figure 2.16(a) presents DNS results for J using the optimal cp profile produced 56 (a) (b) 0 5 10 15 200 0.01 0.02 0.03 0.04 0.05 t J A∞ = 10−1 A∞ = 10−3 A∞ = 10−5 A∞ = 10−7 10−8 10−6 10−3 10−10 5 10 15 20 A∞ t l, t o (c) (d) 10−8 10−6 10−3 10−10 5 10 15 20 A∞ t o 10−8 10−6 10−3 10−11 10 20 30 A∞ t o Figure 2.16: DNS results for Ra = 500 and k = 30 (a) the flux due to base-state, Jb (solid lines), and the flux from DNS, J , (dashed lines) using the MOP cp profile at tp = 0.1. The crosses denote tl while the solid dots denote to. (b) tl (crosses) and to (solid dots) vs. A∞ using the MOP cp profile at tp = 0.1. (c) to vs. A∞ using the COP (crosses) and MOP (solid dots) cp profiles at tp = 0.1. (d) to vs. A∞ using the MOP cp profiles at tp = 0.01 (crosses), tp = 0.05 (plus signs), and tp = 0.2 (circles). Note that a log scale has been used for to. 57 by MOP for tp = 0.1, tf = 5, k = 30, and Ra = 500. Note that the MOP cp profiles are insensitive to the final time when tf > 1. The solid line shows the temporal evolution of the flux due to the base-state, Jb, while the dashed lines show DNS results for J when A∞ = 10−1, 10−3, 10−5, and 10−7. The times, tl and to, are marked with solid dots and crosses respectively. The flux J initially agrees with Jb and then deviates after t = tl due to nonlinear effects. The initial linear regime exists even in the case of large initial amplitude A∞ = 10−1. Figure 2.16(b) illustrates tl (crosses) and to (solid dots) for various perturbation amplitudes A∞. Figure 2.16(c) illustrates to versus A∞ using the optimal profiles produced by COP (crosses) and MOP (solid dots) for tp = 0.1, tf = 5, k = 30, and Ra = 500. The COP scheme produces negative net concentration fields, cnet, for all finite perturbation amplitudes, see table 2.1. For illustration purposes, the maximum amplitude for COP to is arbitrarily set A∞ = 10−6 for which cminnet = −4.1 × 10−7. In this case, COP produces onset times as low as to = 7.29 for A∞ = 10−6. Note, however, that the onset times predicted by COP cannot be realized in physical systems because of cminnet < 0, and are shown for illustration purposes only. In comparison, the MOP supports finite initial amplitudes as large as A∞ = 10−1 for which to = 1.21. This study concludes that the perturbations produced by the MOP are more likely to trigger onset of convection in physical systems. Figure 2.16(d) illustrates to versus A∞ using the MOP cp profiles at tp = 0.01 (crosses), tp = 0.05 (plus signs), and tp = 0.2 (circles) for tf = 5, Ra = 500, and k = 30. Onset of convection occurs later for smaller tp due to the strong initial damping periods. Note that a log scale has also been used for to to highlight the 58 difference for larger A∞. For large amplitude perturbations, the onset of convection can occur around to ≈ 1. For typical aquifer conditions (see §4.3), with permeability K = 10−14 m2 and height, H = 51 m, this corresponds to a dimensional onset time of t∗o ≈ 165 years. 2.7 Conclusions and summary This chapter investigated the linear stability of gravitationally unstable, tran- sient, diffusive boundary layers in isotropic, homogeneous porous media. First, a classical optimization procedure (COP) was performed to determine optimal per- turbations with maximum amplifications. Previously, Tan & Homsy (1986) and Doumenc et al. (2010) have observed that perturbation amplification is sensitive to the perturbation flow field used to measure perturbation magnitude. Because this sensitivity has not been addressed for applications to CO2 sequestration, this study considered three different measures of perturbation amplitude that maximize either the perturbation concentration field, vertical velocity field, or the sum of the per- turbation velocity and concentration fields, which is referred to as the total energy. It was determined that maximizing the perturbation concentration field naturally maximizes the total energy. Maximizing the perturbation velocity field, however, does so at the expense of the concentration field and total energy. Consequently, this dissertation focusses on perturbations that maximize the concentration field because these are expected to be the dominant trigger for onset of nonlinear convection. As the final time, tf , increases, the optimal initial perturbations eventually 59 converge to a fixed shape and cease to vary with increasing tf . This occurs be- cause the final perturbations at t = tf rapidly tend to the dominant quasi-steady eigenmode. In fact, for the current problem, the quasi-steady modal analysis is a good approximation to the COP. Both methods produce nearly identical amplifica- tions and dominant wavenumbers. This suggests that the deviation of the optimal perturbations from the dominant eigenmodes at small times may be primarily due to the transient base-state, rather than the nonorthogonality of the quasi-steady eigenmodes. This is in stark contrast to wall-bounded shear flows for which non- orthogonal eigenmodes often play a dominant role. To judge the relevance of optimal perturbations to physical systems, it is shown that every perturbation has a maximum allowable initial amplitude above which the sum of the base-state and perturbation produces unphysical negative concentrations. This study demonstrates that the optimal initial perturbations predicted by the COP produce unphysical negative concentrations for all finite initial amplitudes. Consequently, onset of convection in physical systems is more likely triggered by suboptimal perturbations that support finite amplitudes. To explore this alternate path to onset of convection, a modified optimization procedure (MOP) is developed that constrains the initial perturbations to be concentrated within the boundary layer. An integral characteristic of the MOP is the concept of a filter function, Ψ(z), that effectively filters out perturbations with concentration fields extending beyond the boundary layer, see equation (2.35). The choice of filter function is not unique, and determines both the maximum allowable initial perturbation amplitude as well 60 as the subsequent perturbation amplification. Filter functions that concentrate the initial perturbation close to z = 0 support large initial amplitudes, but produce small subsequent amplifications. Filter functions that concentrate the perturbations near the boundary layer depth support small initial amplitudes, but produce large sub- sequent amplifications. This raises the possibility that there exists an optimal filter function that balances the effects of the initial amplitude and subsequent amplifica- tion in order to minimize the onset time for convection. This is an avenue for future work. This study focussed on perturbations produced by Ψ = c−1b because this nat- urally concentrates perturbations in regions of large base-state concentration, and because it shows good agreement with corresponding DNS of physical systems. The alternate path to onset of convection taken by the MOP features smaller amplifications and larger dominant wavenumbers than the COP, especially at small initial perturbation times, tp  top. This occurs because the dominant MOP pertur- bations are concentrated within the boundary layer, and consequently experience more initial damping than the COP perturbations. It is shown that the results produced by MOP agree well with the “dominant mode” approach of Riaz et al. (2006) as well as quasi-steady modal analyses performed in the similarity space of the base-state (Riaz et al., 2006; Selim & Rees, 2007a; Wessel-Berg, 2009). To emulate physical experiments, DNS is performed in which the boundary layer is simultaneously perturbed with all wavenumbers resolved by the simulations. The perturbations have a random structure but are concentrated within the bound- ary layer. The DNS results confirm that physical systems follow the alternate path to convection predicted by the MOP scheme and show poor agreement with COP. 61 Furthermore, the MOP perturbations support large initial amplitudes, A∞ ∼ 10−1, and produce early onset times for nonlinear convection. In contrast, the COP per- turbations support neither finite amplitudes nor finite onset times. 62 Chapter 3: Effect of viscosity contrast on gravity-driven instabilities in porous media This chapter examines the effect of viscosity contrast on the linear stability of gravitationally unstable, transient, diffusive layers in porous media. The analysis presented in this chapter helps evaluate experimental observations of various bound- ary layer models that are commonly used to study the sequestration of CO2 in brine aquifers. It is shown that models that allow the interface between CO2 and brine to move, in an effort to capture the effect of dissolution, can more unstable compared to conventional, fixed interface models. Also, diffusive layers are generally found to be more unstable when viscosity decreases with depth within the layer compared to when viscosity increases with depth. This behavior is in contrast to the classical understanding of gravitationally unstable diffusive layers subject to mean flow. For that case, greater instability is associated with the displacement of a more viscous fluid in the direction of gravity by a less viscous fluid. The new phenomenon can be explained as a special case of the classical displacement problem that depends on the relative magnitude of the displacement and buoyancy velocities. For such clas- sical flows, there exists a critical viscosity ratio that determines whether the flow is buoyancy dominated or displacement dominated. The new behavior is explained in 63 terms of the interaction of vorticity components related to gravitational and viscous effects. 3.1 Overview During CO2 sequestration, free CO2 is trapped as it dissolves into brine across a two-phase interface, see figure 3.1. Once dissolved, the trapped CO2 diffuses downwards to form a solute boundary layer. These diffusive boundary layers have been studied with the help of various models. These models often assume that CO2 dissolves into brine across the two-phase interface at constant pressure and temperature. The concentration of dissolved CO2 at the interface across which CO2 dissolves into brine is therefore taken to be constant. The interface motion resulting from dissolution is considered to be small by one popular model in comparison with other relevant time scales in the problem. The interface location is therefore considered to be fixed. This model is referred to as in this chapter as the fixed interface model (Ennis-King et al., 2003; Riaz et al., 2006; Slim & Ramakrishnan, 2010). This model is a popularly used in many theoretical and computational studies and also in previous chapter of this study. An alternative model of the diffusive boundary layer attempts to incorporate the motion of the interface by considering a diffused layer that separates two initially quiescent, miscible fluids. For this model, a non-monotonic density-concentration relationship is used to produce both stable and unstable regions within the boundary layer (Backhaus et al., 2011; Ehyaei & Kiger, 2014; MacMinn et al., 2012; Neufeld et al., 2010). The overall result is the 64 apparent motion of the diffusive layer after the onset of nonlinear convection. This setup as the moving interface model. Because of the relative ease of laboratory setup, the moving interface model has gained more popularity with experimental studies compared with the fixed interface model. The moving interface model however has not been well explored and a fundamental insight regarding the physical behavior is lacking. Consequently, it is unclear how the stability characteristics of the two models compare with each other. From a practical stand point, differences in the viscosities of CO2-brine solu- tion and CO2-free brine are small (Bando et al., 2004; Kumagai & Yokoyama, 1999). However, because of physical constraints, the fluids employed by experimental stud- ies lead to very different viscosity contrasts than what is expected in practice. In some cases, the viscosity of the experimental fluid representing CO2-brine solution is about 20 times larger than that of the fluid representing CO2-free brine solution Backhaus et al. (2011). Since the fixed and moving interface models are frequently used as analogs for physical systems, the effect of viscosity contrast on stability behavior needs to be understood to properly interpret corresponding experimental observations. In order to facilitate such an understanding, this study draws com- parison with the closely related problem of the gravitationally unstable diffusive layer subject to mean flow. The moving interface model is a special case (with zero mean flow) of this displaced interface problem. For this classical problem the viscosity contrast, density difference and mean flow, all interact to affect stability behavior (Manickam & Homsy, 1995). Evaluation of such interactions is expected to facilitate a deeper understanding of relevant physical mechanisms for the moving 65 Figure 3.1: Sketch of CO2 sequestration. Dissolution of CO2 into brine occurs across the two-phase interface, indicated by pairs of counter- pointing arrows. The gravitationally unstable CO2 layer within brine plays a vital role in determining the interfacial dissolution rate. interface problem, and also by extension for the fixed interface problem. The main highlights of this chapter are; (i) investigation for the effect of the viscosity contrast for the fixed and moving interface models and (ii) exploration of the interaction of mean flow, buoyancy velocity and viscosity contrast for un- derstanding the transition from displacement dominated to buoyancy dominated behavior. These two features are explored by means of a quasi-steady-state eigen- value approach in self-similar space of the diffusive boundary layer. The suitability of this approach is confirmed by our findings in Chapter 2. The work is divided as follow. The geometries and governing equations are explained in §3.2. The results are discussed in §3.3 along with conclusions in §3.4. 66 Figure 3.2: (a) Non-monotonic density-concentration profile employed in a MI model (b) Monotonic viscosity-concentration profiles for various log mobility ratios, R = ln(µ0/µ1). 3.2 Governing equations In order to evaluate experimental setups based on the moving interface (MI) model of the diffusive boundary layer, this study uses a non-monotonic density profile, ρ∗, of the form illustrated in figure 3.2(a). This density profile can be represented as, ρ∗ = ρ0 + ∆ρF (c), (3.1) where the function F (c) = ∑4 n=1 anc n, determines how density varies with con- centration c. The end point densities related to c = 0 and c = 1 are ρ0 and ρ1, respectively, and ρm is the maximum density. Note that the fluid with c = 1 lies above the fluid with c = 0. The quantity, ∆ρ = ρm − ρ0, indicates the strength of unstable density stratification. The function F (c) is normalized to the maximum 67 value of one. The density profile is linear when a1 = 1 and an = 0 for n = 2, 3, 4. Fol- lowing previous works, this study employs a monotonic viscosity profile illustrated in figure 3.2(b), µ∗ = µ1 exp (R (1− c)), (3.2) where R = ln(µ0/µ1) is the log mobility ratio, µ1 is the viscosity of the fluid with c = 1, and µ0 is the viscosity of the fluid with c = 0. For the experimental study of Backhaus et al. (2011) based on the moving in- terface model, water and propylene glycol were used as the lighter and heavier fluids, respectively. For that system, the location of the density peak occurs at a concen- tration of c ≈ 0.38. A log mobility ratio, R ≈ −3, fits the viscosity-concentration relationship at a temperature of 120 ◦F. Neufeld et al. (2010), MacMinn et al. (2012) and Ehyaei & Kiger (2014) also employ a moving interface model, using methanol/ethylene glycol (MEG) mixtures and water as the lighter and heavier fluids respectively. The location of the density peak and the viscosity differences depend on the composition of the MEG mixture. Typical values of the peak density vary in the range, 0.2 < c < 0.55 (Huppert et al., 1986), while the log mobility ratios vary approximately in the range, −1.5 < R < 1 MacMinn et al. (2012). An- other experimental study by Slim et al. (2013) employs a setup that is closer to the fixed interface model. The authors employed potassium permanganate (KMnO4) in water as an analogous model for CO2 in brine. The KMnO4-water mixture approx- imately satisfies a linear density profile and a log mobility ratio of R ≈ 0.04 fits the viscosity-concentration relationship at 77 ◦F (Jones & Fornwalt, 1936). 68 The porous aquifer is modeled as isotropic, homogeneous, and of infinite hor- izontal extent and finite depth H. The vertical coordinate, z, is positive in the direction of gravity, g. The porous medium is characterized by permeability, K, dispersivity, D, and porosity, φ. The characteristic values are H for length, µ1 for viscosity, K∆ρg/µ1 for velocity, µ1H/K∆ρgφ for time and ∆ρgH for pressure. Us- ing these characteristics values, one obtains the following non-dimensional governing equations, µ(c)v +∇p− F (c)ez = 0, ∇ · v = 0, ∂c ∂t + v · ∇c− 1 Ra ∇2c = 0. (3.3) The Rayleigh number is defined as Ra = K∆ρgH/φDµ1. The symbol v = [u, v, w] is the nondimensional velocity vector, and p is the nondimensional pressure obtained from the dimensional pressure pˆ through the relation p = (pˆ − ρogz)/∆ρgH. The symbol ez is the unit vector in the z-direction. The boundary conditions for (3.3) depend on the model. For the fixed interface (FI) model, the boundary conditions for (3.3) are, c ∣∣ z=0 = 1, ∂c ∂z ∣∣∣∣ z=1 = 0, w ∣∣ z=0 = w ∣∣∣ z=1 = 0. (3.4) Equations (3.3) and (3.4) admit the concentration base state, cFb (z, t) = erfc(z √ Ra/4t), see figure 3.3(a) for illustration. The velocity base-state is vb = 0. For the MI model, the study uses boundary conditions that allow diffusion in two opposite directions, ∂c ∂z ∣∣∣∣ z=−1 = ∂c ∂z ∣∣∣∣ z=1 = 0, w ∣∣∣ z=−1 = w ∣∣∣ z=1 = 0. (3.5) Equations (3.3) and (3.5) admit the base-states, cMb (z, t) = erfc(z √ Ra/4t)/2, see 69 Figure 3.3: Concentration base-states for Rayleigh number, Ra = 500, at various instants of time t. (a) Base-state for FI model , cFb = erfc(z √ Ra/4t). (b) Base-state for MI model, cMb = 0.5 erfc(z √ Ra/4t). figure 3.3(b), and vb = 0. These expressions of the base-states, cFb and c M b , are valid as long as the boundary layer remains far away from the boundaries at z=1 for the FI model and z = ±1 for the MI model respectively. This holds true when √ Ra/4t > 3.(Riaz et al., 2006) The linear stability of various diffusive boundary layer models is studied with respect to small wavelike perturbations of the form, c˜ = ĉ(z, t)ei(αx+βy), v˜ = v̂(z, t)ei(αx+βy), (3.6) where i = √ −1, α and β are wavenumbers in the x- and y-directions respectively, and ĉ(z, t) and v̂(z, t) are time-dependent perturbation profiles in the z-direction. Substituting c = cb + c˜ and v = vb + v˜ into equation (3.3) and linearizing about base-states, one obtains the following initial value problem for ĉ and ŵ, ( ∂ ∂t − 1 Ra ∂2 ∂z2 − k2 ) ĉ+ ∂cb ∂z ŵ = 0, (3.7) 70 ( ∂2 ∂z2 −R∂cb ∂z ∂ ∂z − k2 ) ŵ +Gk2ĉ = 0, (3.8) where k = √ α2 + β2, G = 1/µ(cb) ∂F (cb)/∂cb+UR and cb refers to either of the FI or MI base states defined above. Homogeneous Dirichlet boundary conditions for perturbation variables are specified at z = 1 and z = ±1 for the FI and MI models, respectively. The symbol, U = U∗µ1/K∆ρg, refers to the fluid displacement velocity, U∗, scaled with the buoyancy velocity, K∆ρg/µ1. It indicates the relative strength of the mean flow with respect to buoyancy velocity and is effective only when viscosity varies in the boundary layer, R 6= 0. When U = 0, equations (3.7)–(3.8) represent either the FI or MI models. When U 6= 0, the equations represent the displaced in- terface model. The coordinate system for the displaced interface model is such that it moves with velocity, Uez. The associated linear stability equations were obtained by first performing coordinate transformations to equations (3.3) and (3.5) before carrying out an expansion using normal modes. Note that due to coordinate trans- formation, the boundary conditions and resulting base-state for displaced interface model are same as that of the MI model.(Manickam & Homsy, 1995) Equations (3.7)–(3.8) are analyzed using a quasi-steady-state (QSSA) eigen- value formulation in the self-similar (ξ, t) space, where ξ = az and a = √ Ra/4t.(Riaz et al., 2006) The resulting eigenvalue problem may be expressed as σce = ξ 2 ∂ce ∂ξ + 1 Ra ( a2 ∂2 ∂ξ2 − k2 ) ce − a ∂cb ∂ξ we, (3.9) ( a2 ∂2 ∂ξ2 − a2R∂cb ∂ξ ∂ ∂ξ − k2 ) we = −Gk2ce, (3.10) with homogeneous Dirchlet boundary conditions for the eigenmodes, ce and we, 71 Figure 3.4: Isocontours of σ = 0 produced by an FI model. Arrows point toward the unstable region, σ > 0. Solid dots mark the critical points, (kc, tc). and the eigenvalue, σ, represents the growth rate. The least stable perturbation is defined as the eigenmode with the maximum real value for σ. The growth rates obtained in the self-similar space (ξ, t) are equivalent to the growth rates calcu- lated in the regular space (z, t) when perturbation amplitudes are based on the L∞ norm.(Tilton et al., 2013) The equations (3.9)–(3.10) are discretized using standard second-order finite difference schemes. For given parameters of k, t and Ra, the gen- eralized eigenvalue problem is solved using function ‘eig’ in MATLAB. The onset time for linear instability, t = tc, is defined as the time at which the growth rate of a perturbation eigenmode first becomes positive. The corresponding wavenumber is called the critical wavenumber, k = ko. 72 3.3 Results and discussion This section examines the effect of viscosity contrast on the onset of instabil- ity in a gravitationally unstable, transient, diffusive boundary layer. The moving interface model is compared extensively with the fixed interface model and is fur- ther explored by considering various types of non-monotonic density-concentration distributions. This section also explores how mean flow and viscosity contrast deter- mines the shift in stability features from that of displacement dominated to those of buoyancy dominated behavior. For the remainder of the study, the Rayleigh number is fixed at Ra = 500. Linear stability behavior at other value of Ra > 50 can be obtained by a simple rescaling.(Riaz et al., 2006; Tilton et al., 2013) 3.3.1 The fixed interface model The diffusive boundary layer in the fixed interface (FI) model adopts a linear density-concentration relationship such that F (c) = c. Figure 3.4 illustrates the isocontours of growth rates, σ = 0, in the (k, t) parameter plane for log mobility ratios, R = −1 (solid line), R = 0 (dashed line), and R = 1 (dash-dotted line). The lowest point of the σ = 0 isocontour corresponds to the critical parameters (kc, tc). For R = −1, the critical point is at k = 66.8 and t = 0.1 respectively. The arrows point towards the unstable zone where the growth rates are greater than zero, σ > 0. For small times, all perturbation wavenumbers are stable, after which a band of wavenumbers become unstable. When R = 0, one recovers the constant viscosity case(Riaz et al., 2006) with the critical point at (34.7,0.3). The R = 0 case 73 has a smaller band of unstable wavenumbers compared to R = −1. The unstable region shrinks further when the viscosity ratio is increased to R = 1, resulting in smaller kc and larger tc. A similar effect of R on transient diffusive boundary layers has also been observed by Meulenbroek et al. (2013). The dominant wavenumber, kmax, is defined as the wavenumber at a given time t for which the largest growth rate is observed, σmax(t) = sup 0≤k<∞ σ(t, k). (3.11) Figure 3.5(a) illustrates the temporal evolution of σmax for R = −1 (solid line), R = 0 (dashed line), and R = 1 (dash-dotted line). The R = −1 perturbations have larger σmax compared to R = 0 and R = 1 perturbations. When t < 1, the R = −1 perturbations attain growth rates as large as σmax ≈ 6. In contrast, the R = 1 perturbations are stable for the same period with σmax < 0. Figure 3.5(b) illustrates the dominant wavenumbers, kmax, for R = −1 (solid line), R = 0 (dashed line), and R = 1 (dash-dotted line). The dominant wavenumbers, kmax, monotonically decrease with increasing time. The R = −1 perturbations have larger values of kmax compared to the R = 1 perturbations. It is found that the increasing strength of the instabilities for decreasing R directly correlates to the perturbation’s instantaneous vorticity field, Ωe = k µ(cb) ce − R k ∂cb ∂z ∂we ∂z . (3.12) The individual contributions to vorticity are examined using, I = ∫ Ωe dz = I1 + I2, (3.13) 74 Figure 3.5: Results produced by FI model. (a) Maximum growthrates, σmax, vs. t. (b) Dominant wavenumbers, kmax, vs. t. The solid points represent the critical point of instability, (kc, tc). where I1 = k ∫ exp[−R (1− cb)]ce dz, I2 = − R k ∫ ∂cb ∂z ∂we ∂z dz. (3.14) The first integral, I1, measures the contribution to vorticity arising from the buoyant forces. When density gradients are unstable, I1 is positive. The second integral I2 depends both on the base-state and the velocity perturbation. Table 3.1 illustrates the values of the integral quantities, I1 and I2, and the growth rate σ for values of R ranging from -3 to 3. The eigenmodes are normalized such that maximum value of ce is one. The values of I2 are consistently smaller than I1. When R = 3, one observes the smallest values for I1. This indicates that the buoyancy velocities produced due to the unstable density gradients are small, and consequently, tend to have weak destabilizing effects. With decreasing R, values of I1 increases till a maximum value is reached at R = −3. Larger values 75 R I1 I2 I1 + I2 σ −3 12.67 −5.40 7.27 23.39 −2 5.37 −1.75 3.63 8.99 −1 2.49 −0.45 2.04 1.70 0 1.26 0.00 1.26 −2.23 1 0.68 0.13 0.80 −4.41 2 0.38 0.13 0.51 −5.59 3 0.23 0.09 0.32 −6.21 Table 3.1: Vorticity values and growth rate produced by the FI model for k = 30 and t = 0.2 Figure 3.6: Base-state and least stable eigenmodes produced by the FI model as a function of self-similar coordinate, ξ, for k = 30, and t = 0.2. Viscosity values are presented at the top axis. (a) R = −1.5. (b) R = 1.5 76 of the vorticity integral, I, produce higher growth rates, and consequently, produce stronger perturbation fields that promote the formation of instabilities. Figure 3.6 illustrates the base-state, cFb (solid line), concentration eigenmode, ce (dashed line), and vertical velocity eigenmode, we (dash-dotted line) for R = −1.5 (panel a) and R = 1.5 (panel b) respectively. The numbers along top axis represent fluid viscosity obtained using µ = exp (R(1− cFb )). When R = −1.5, one observes a large vertical velocity perturbation field, we, that is nearly of the same magnitude as the concentration perturbation field ce. The ratio of the maximum value of we and the maximum value of ce is 0.92. For large value of the log mobility ratio, R = 1.5, weaker vertical velocity fields are observed. The ratio of the maximum values drops to 0.13. The drop in magnitudes is because we fields associated with R = 1.5 are formed in regions of higher viscosity (more resistance to fluid flow) compared to R = −1.5. The viscosity variation within the boundary layer also explains why perturbations are concentrated away from z = 0 for R = −1.5 compared to R = 1.5. When R > 0, viscosity increases with depth and perturbation peaks are closer to z = 0, compared to R < 0 where viscosity decreases with depth and perturbation peaks are concentrated in regions away from z = 0. Note that the classical behavior of greater instability associated with higher R is due to a source of perturbations arising from the background mean flow, as explained in more detail in section 3.3.4. 77 Figure 3.7: Comparison between MI and FI models. (a) Nonmonotonic function F as a function of concentration c, see equation (3.1). The coefficients of F (c) are: a1 = 1.06, a2 = 17.31, a3 = −39.35, a4 = 12.28. (b) cMb vs. z for t = 10. The density gradients are destabilizing only when z > γ (arrow). (c) tc vs. R. (d) kc vs. R. 78 Figure 3.8: Linear stability results for FI with R = 0 (solid line) and MI with R = 0.48 (dashed line). (a) σmax vs. t. (b) kmax vs. t. (c) ce/‖ce‖∞ vs. ξ for k = 30 and t = 1. (d) we/‖we‖∞ vs. ξ for k = 30 and t = 1. 79 3.3.2 The moving interface model This section explores the effect of viscosity contrast on the moving interface (MI) model. A non-monotonic density profile is employed such that the function F (c), see equation (3.1), corresponds to aqueous propylene glycol mixtures , with c = 0 and c = 1 representing concentrations of pure propylene glycol and wa- ter respectively. Figure 3.7(a) illustrates the function F (c). The positive density gradients when c < 0.38 promote the formation of instabilities. Figure 3.7(b) il- lustrates the location of the zone of unstable density gradients within the diffusive layer. Unlike an FI model, where the entire boundary layer is unstable, the unstable density gradients in an MI model exist only when z > γ(t) (or ξ > 0.21), where γ(t) = 0.21 √ 4t/Ra. Figure 3.7(c) illustrates the onset times, tc, versus the log mobility ratio, R, for the MI (circles) and FI (crosses) models. As expected, the onset times produced by the two models increase with increasing R. A large difference in the onset times occurs especially for small values of R. When R = −2, the FI model has onset times, tc, that are approximately 500% larger than those produced by the MI model. With increasing R, the onset times produced the two models agree better. For R ≈ 1.8, the FI and MI models have nearly identical onset times. Figure 3.7(d) illustrates the critical wavenumbers, kc, versus R for the MI (circles) and FI (crosses) models. As expected, kc decreases with increasing R. For R < 1.36, the MI model has larger kc compared to the FI model. The linear stability behavior of the FI and MI models can be made to coincide 80 when different viscosity contrasts are used with each model. For example, when R = 0 for the FI model and R = 0.48 for the MI model, the onset times, tc, predicted by the two models agree with each other. For these parameters, good agreement is obtained even for time, t > O(tc). This is depicted in figures 3.8(a)–3.8(b) where the temporal evolution of the dominant growth rates, σmax, (panel a) and dominant wavenumbers, kmax, (panel b) are plotted for FI model with R = 0 (solid line) and MI model with R = 0.48 (dashed line). Figure 3.8(c) illustrates ce/‖ce‖∞ versus ξ for k = 30 and t = 1. The ce profile produced by the MI model (dashed line) for R = 0.48 is identical to that of the FI model (solid line) for R = 0 except for a small narrow region across z = 0. Figure 3.8(d) illustrates the corresponding normalized we for k = 30 and t = 1. The we profiles are similar for ξ > 0. The FI domain (solid line) does not exist when ξ < 0. When we < 0 as for the MI model (dashed line), the region associated with it becomes stabilizing, see term we∂cb/∂z in equations (3.9)–(3.10). This region does not contribute to perturbation growth. To further illustrate the above argument, figure 3.9(a) shows the base-state, cMb (solid line), least stable ce (dashed line), and we (dash-dotted line), for k = kc, t = tc when R = 0 in the MI model. The vertical line represents the location of the peak density, ξ = 0.21 or z = γ. The perturbations, ce and we, are concentrated in the destabilizing zone where we > 0. Due to momentum transport, the point (solid dot) where velocity changes sign from a negative to a positive value does not coincide with the location of zero gradient of density, ξ = 0.21. Figure 3.9(b) depicts the normalized we profiles for k = 15, t = 1 and log mobility ratios, R = 0 (solid line), R = 1 (dashed line), and R = 2 (dash-dotted line). With increasing R, 81 there is less momentum transport across the location of peak density at ξ = 0.21. Consequently, the solid dots move to the right, indicating a decrease in the width of the boundary layer corresponding with lesser instability, as shown in figure 3.7(c). The following explains why the MI model is more unstable at small R and less unstable at large R, as seen in figure 3.7(c). Unlike the FI model, it is found that for the MI model, momentum transports across the region linking the zones of stable and unstable density gradients at z = γ or ξ = 0.21. FI model offers no such mechanism. The magnitude of momentum transport in the MI model is proportional to the strength of the we fields within the boundary layer. Consequently, the width of the unstable portion of the boundary layer depend on R. The MI model is more unstable than the FI model when the width of the unstable region is large enough to produce stronger instabilities than an FI model. With increasing R, the width of the destabilizing zone becomes narrower, producing weaker instabilities compared to the FI model. 3.3.3 Effect of non-monotonic density profiles The features of the non-monotonic density profile are varied in order to inves- tigate its effect on the results produced by the MI model. The density profile used in §3.3.2 is referred to as ρA. To investigate how varying the location of the maximum density may affect the onset times, this study uses two additional profiles, ρB and ρC , by modifying the function F (c), see equation (3.1). Figure 3.10(a) illustrates the function F (c) for density profiles, ρA (circles), ρB (crosses), and ρC (squares). 82 Figure 3.9: Eigenmodes associated with the MI model (a) Base-state, cMb (solid line), least stable ce (dashed line), and we (dash-dotted line) at critical point (kc, tc) when R = 0. The vertical line is drawn at the location of maximum (turning point) in the density profile. (b) Normalized we profiles at k = 15 and t = 1 for log mobility ratios, R = 0 (solid line), R = 1 (dashed line), and R = 2 (dash-dotted line). The dots emphazise the points where we = 0. 83 Figure 3.10: Five different non-monotonic density profiles in the MI model. (a)–(b) Effect of changing the location of the maximum value of density. (c)–(d) Effect of changing the density values of the saturated fluid, c = 1. 84 The maximum value of density for the ρA profile occurs at c = 0.38, while it is at c = 0.5 for ρB and at c = 0.25 for ρC . As peak value of density moves closer to c = 1, the width of the zone of unstable density gradients increases but the magnitude of positive density gradients decreases. Figure 3.10(b) illustrates the onset time, tc, versus log mobility ratio, R, ob- tained using the MI model with density profiles, ρA (circles), ρB (crosses), and ρC (squares). The onset times produced by the three density profiles increase with in- creasing R. When R = −2, the onset time, tc, produced by ρB is 1.8 times greater than those produced by ρC . This suggests that the onset times obtained using ρB are closer to those predicted by FI model compared to other density profiles. For R ≈ −0.2, three density profiles produce nearly equal tc. With increasing R, the ρC profile tends to produce the largest onset times, tc. Figure 3.10(c) illustrates F (c) for peak density location at c = 0.375 (solid dot) and for different values of density at c = 1. Modifying the density values of the saturated fluid changes the slope of the negative density gradients. The density profile, ρE (squares), has larger gradients compared to ρA (circles) and ρD (crosses). Figure 3.10(d) illustrates the corresponding onset times as function of R. The onset times for the different profiles are identical. The stability characteristics observed in figures 3.10(b) and 3.10(d) can be explained by examining the relative location of the velocity eigenmode, we, and the fluid viscosity. Figure 3.11(a) illustrates we versus ξ for R = 0, k = 30, t = 1, and density profiles, ρA (solid line), ρB (dashed line), ρC (dash-dotted line). The viscosity values are mentioned along the top axes in figure 3.11. The crosses 85 Figure 3.11: Effect of density profiles on velocity eigenmode we. The top axis represents viscosity at corresponding depthwise coordinate ξ. The location of the maximum density associated with each profile is marked with a cross. (a) we/‖we‖∞ vs. ξ for R = 0, k = 30, t = 1, and ρA (solid line), ρB (dashed line), ρC (dash-dotted line). (b) Same as in panel (a) for R = −2. (c) Same as in panel (a) for R = 2. (d) Same as in panel (a) for ρA (solid line), ρD (dashed line), ρE (dash-dotted line). 86 denote the location of maximum value of density. Perturbations are predominantly concentrated in regions to the right of the crosses where unstable density gradients exist. When maximum density occurs at c = 0.5 as in ρB, one observes that the peak of we shifts to the right in comparison to we related to ρA for which maximum density is at c = 0.38. When maximum density is at c = 0.25 as for ρC , the peak of we shifts to the left. The change in perturbation structures occurs due to the shifting of unstable regions within the boundary layer. Though the perturbations produced by ρA, ρB, and ρC peak at three different locations, the corresponding onset times for R = 0, shown in figure 3.10(b), are similar. Let us now consider the effect of the spatial variation of viscosity within the boundary layer. Figure 3.11(a) is repeated for R = −2 in figure 3.11(b) and for R = 2 in figure 3.11(c) respectively. For R = −2, one finds that we fields remain shifted with respect to each other. Because viscosity decreases with depth when R < 0, the perturbations produced by ρC profiles are now located in lower viscous regions compared to perturbations produced by ρA or ρB. Consequently, perturba- tions produced by ρC have much earlier onset times compared to other profiles when R < 0, see figure 3.10(b). For R = 2, the perturbations produced by ρB profiles are located in lower viscosity regions, and therefore, have earlier onset times. Density profiles ρD and ρE have insignificant effects on perturbation structures, see we pro- files in figure 3.11(d). The momentum transport across peak density locations has increased slightly when decreasing the magnitude of the negative density gradients. This is reflected in the upward movement of the crosses. However, these changes were small and did not affect the onset times for instability shown previously in 87 figure 3.10(d). 3.3.4 Effect of uniform flow In previous sections, it is observed that gravitationally unstable boundary lay- ers are less unstable for larger values of R. This behavior contrasts with classical displacement problems where instability increases with R. To explain this difference in behavior, this study considers a displaced interface model, which is an extension of the MI model where the lighter fluid is displaced by the heavier fluid with a uniform velocity, U , along the direction of gravity. For U = 1, the dimensional displacement velocity is equal to the buoyancy velocity of K∆ρg/µ1, see §3.2 for details. To facilitate comparison with previous studies on displacement based in- stabilities (Manickam & Homsy, 1995), a linear density profile is first considered in addition to non-monotonic density profiles illustrated in figure 3.10(a). Figure 3.12(a) illustrates onset times, tc, versus log mobility ratio, R, for displacement velocities, U = 0 (circles), U = 0.5 (squares), and U = 1(crosses), using a linear density profile. As expected, when U = 0, one observes that tc increases with increasing R. For U = 0.5, tc increases with R until it attains a maximum value at R ≈ 0. Beyond this R, tc decreases with R. When U = 1, similar behavior is observed with the maximum value of tc occurring at R = −1. Figure 3.12(b) illustrates corresponding critical wavenumbers, kc, versus R for U = 0 (circles), U = 0.5 (squares), and U = 1(crosses). For U = 0, kc monotonically decreases with R. For U = 0.5, kc decreases with increasing R until it reaches a 88 Figure 3.12: Effect of viscosity contrast on the displaced interface model. (a) Onset (critical) time tc vs. log mobility ratio R, for U = 0 (circles), U = 0.5 (squares), and U = 1 (crosses). Solid dot marks Rc when U = 1. Solid dot marks Rc when U = 1.(b) Critical wavenumber kc vs. R, for U = 0 (circles), U = 0.5 (squares), and U = 1 (crosses). (c) Critical viscosity ratio, Rc, vs. U for linear density profile (diamond), ρA (circles) , ρB (crosses) , ρC (squares). (d) Rc vs U for a linear density profile. The shaded region represents the space of Rc and U for which no instabilities are formed. 89 minimum at R = 0 and increases afterwards. For U = 1, the minimum point is located at a lower value of R. Similar qualitative trends were also observed for tc and kc produced by non-monotonic density profiles. Figures 3.12(a) and 3.12(b) depict the existence of qualitatively different sta- bility behaviors for the critical parameters, tc and kc. The value of R at which the instability characteristics of tc changes is referred to as the critical viscosity ratio, Rc. When R < Rc, tc increases with increase in R. This feature is in sharp contrast with classical displacement behavior and is more in accordance with buoyancy insta- bilities demonstrated in previous sections. When R > Rc, tc decreases with increase in R, depicting the dominance of displacement-related instabilities. It is also ob- served that the point of maximum tc and minimum kc do not coincide but are close to each other. This may be because of sensitivity issues regarding measurement of perturbation quantities during small times, t < O(tc).Tilton et al. (2013) Figure 3.12(c) illustrates Rc versus U for density profiles, ρA (circles), ρB (crosses), ρC (squares), and a linear density profile (diamonds), for mean flow in the range, 0 < U < 2. For any given density profile, when displacement velocity tends to zero, U → 0, the critical log mobility ratio, Rc, approaches infinity, Rc → ∞. With increasing U , Rc decreases. The rate at which Rc decreases is larger for a linear density profile followed by the density profiles, ρB, ρA and ρC respectively. This suggests that the decay rate is proportional to the width of the zone of positive density gradients within the boundary layer, see section 3.3.3. By increasing the displacement velocity beyond U = 2, Rc splits into two branches. This is depicted in figure 3.12(d) for a linear density profile. The split at U ≈ 2.7 is due to the 90 R I1 I2 I3 I1 + I2 + I3 σ −3 15.90 −3.29 −4.89 7.71 19.79 −2 7.16 −0.76 −3.55 2.85 4.88 −1 3.48 −0.11 −1.92 1.44 0.85 0 1.90 0.00 0.00 1.90 3.00 1 1.19 −0.07 1.86 2.97 7.36 2 0.82 −0.31 3.66 4.18 12.17 3 0.62 −0.74 5.43 5.31 16.59 Table 3.2: Vorticity integral values and growth rates for U = 1, k = 30, t = 0.2 and Ra = 500. formation of a stable region when U is larger than a certain critical value. To gain a deeper insight into relevant physical mechanisms, the instantaneous perturbation vorticity field is defined as, Ωe = k µ(cb) ce − R k ∂cb ∂z ∂we ∂z + kRUce. (3.15) Equation (3.15) is integrated to obtain a measure of the vorticity field given by, I = ∫ Ωe dz = I1 + I2 + I3, (3.16) where I1 = k ∫ exp (−R(1− cb)) ce dz, I2 = − R k ∫ ∂cb ∂z ∂we ∂z dz, I3 = kRU ∫ ce dz. (3.17) 91 Compared to equation (3.13) in §3.3, there is an extra component to the vorticity field, I3, that depends on the uniform flow, U . For R > 0, I3 is positive and destabilizing, and vice-versa for R < 0. Uniform flow plays an insignificant role on the stability of the diffusive front when R tends to zero, R→ 0. Table 3.2 illustrates the values of the vorticity integrals, I1, I2, and I3, and the growth rate, σ, using a linear density profile for U = 1, k = 30, and t = 0.2. The eigenmodes are normalized such that the maximum value of ce is one. The smallest values of I and σ are observed when the log mobility ratio is close to its critical value, R = −1 (R ≈ Rc). With increasing or decreasing R, the magnitude of I increases, and consequently, the growth rate σ. For smaller values of R, the major contribution to the vorticity comes from the buoyancy term I1. Though the uniform flow contribution, I3 is stabilizing, it is not strong enough to overcome the density mechanisms. For larger values when R > 0, the term I1 has small magnitudes, and therefore the major contribution to the vorticity comes from the flow term I3. The vorticity integrals also explain the presence of the stable zone in figure 3.12(d). Within the stable zone where R < 0 and U > 0, the stabilizing effects of I3 are stronger than destabilizing effects of buoyant term, I1. The stabilizing effect of U when R < 0 has also been previously reported by Manickam and Homsy.(Manickam & Homsy, 1995) Note that when fluids move against the direction of gravity, U < 0, the perturbation characteristics are similar to those of buoyancy dominated instabil- ities. This is because when U < 0, both I1 and I3 becomes increasingly destabilizing with decreasing R. For non-monotonic density profiles, the strength of displacement dominated 92 Figure 3.13: Effect of displacement velocity on the velocity eigenmode we. Viscosity values are mentioned at the top. The crosses denote the location of the maximum in the density profile, ρA. (a) we/‖we‖∞ vs. ξ for R = −1, k = 30, and t = 1. (b) Same as in panel (a) for R = 1. instabilities would affect the momentum transport across the interface of zero den- sity gradient. Figure 3.13(a) illustrates the normalized vertical velocity profiles, we/‖we‖∞, for U = 0 (solid line), U = 1 (dashed line) , and U = 2 (dashed-dotted line) when R = −1, k = 30, and t = 1. The crosses correspond to the peak density location at ξ = 0.21 for ρA profile. With increasing values of U , the strength of the we fields associated with R = −1 decreases due to stabilizing effects of the uniform flow. Therefore, the momentum transport across ξ = 0.21 also decreases. On the other hand, R > 0 results in greater momentum transport. This is illustrated in figure 3.13(b) for R = 1. In this case the velocity perturbations increase in strength due to increased destabilizing effects of the uniform flow. The critical Rc is also an excellent indicator of long term linear viscous char- acteristics for time, t  tc, even though Rc was defined with respect to the onset 93 Figure 3.14: Long term linear stability characteristics, σmax vs. R, using a linear density profile for (a) U = 0.5 (b) U = 2. time of linear instability, tc. To demonstrate this, figure 3.14(a) illustrates σmax vs R for U = 0.5 using a linear density profile. The vertical dashed line represents R = Rc. The solid dots denote the points at which ∂σmax/∂R = 0. Across this minima, there is a reversal of instability characteristics associated with σmax. At t = 0.1, σmax has a nonmonotonic behavior with R with the local minima at R = 0. At t = 1 and t = 10, the local minima is still at R = 0. The viscosity ratio at which local minima occurs for late times are in excellent agreement with Rc. Figure 3.14(b) repeats figure 3.14(b) for U = 2. Large values of t were used because of smaller perturbation growth rates. The locations of local minima (solid dots) are within 10 percent of Rc even for late times, t < 1000. 94 3.4 Conclusions This chapter examines the effect of viscosity contrast on the linear stability of gravitationally unstable, transient boundary layers. To interpret experimental ob- servations, the current work considered two physical models that are characterized by the specific depth-wise concentration profiles and density-concentration relation- ships. If laboratory studies are carried out with fluids for which R ≈ 0, our study indicates that the onset times of instability determined by the MI and FI models can differ by a factor of 3. For R < 0, the MI model becomes even more unstable and gives rise to much earlier onset times compared to the FI model. By adopting fluids with R ≈ 0.5 for the MI model and R = 0 for the FI model, the MI and FI models produce similar linear stability characteristics. The exact value of R for MI model that provides the agreement would also depend on the specific profiles of density-concentration relationship. Previous works (Ennis-King et al., 2003; Riaz et al., 2006) demonstrate that the onset times scale proportionately to the Rayleigh number. Therefore, the ratio of the onset times for the fixed and moving interface models is not a function of the Rayleigh number. Diffusive layers are more unstable in general when viscosity decreases with depth within the layer compared to when viscosity increases with depth. This behavior is in contrast to the classical understanding of gravitationally unstable, transient diffusive layers. For that case, greater instability is associated with the displacement of a more viscous fluid in the direction of gravity by a less viscous fluid. To place this behavior in context, it is shown how instability mechanisms 95 depend on the critical value of the log mobility ratio, Rc. When R < Rc, density gradients govern instability. When R > Rc, perturbations derive energy from the background flow. When the magnitude of displacement velocity that is scaled with buoyancy velocity exceeds a certain threshold value, Rc splits into two branches due to the formation of an intermediate stable zone. The mechanisms are explained in terms of the interaction of vorticity components related to gravitational and viscous effects. Available data on the viscosity-concentration relationship for the CO2-water system indicates some uncertainty with regards to whether the viscosity of the mix- ture would increase or decrease upon dissolution of CO2. Different studies report both positive and negative values of R (Bando et al., 2004; Kumagai & Yokoyama, 1998, 1999), though all studies report viscosity differences to be small, |R| ≈ O(0.1). Therefore it is determined that viscosity contrast would not substantially alter the stability results based on the R = 0 case for practical problems. However, the viscosity contrast between the solution and the solvent for the fluids employed by experimental studies is large. For example, the experimental study of Backhaus et al. (2011) was based on the moving interface model where the viscosity contrast was about R ≈ −3. This implies a much greater level of instability compared with the case of R ≈ 0. For the experimental study of Slim et al. (2013) based on the fixed interface model, on the other hand, R ≈ 0.04, which is closer to what is ex- pected in practice. However, it is uncertain whether the model employed in that study exactly corresponds to either the fixed or the moving interface models. This is because the authors observed bubbles at the top boundary for certain experimental 96 runs. This may suggest that the presence of small velocity perturbations at the top boundary. Such FI models usually have much smaller onset time of instabilties compared to standard FI models (Elenius et al., 2012). In general, modeling and theoretical studies based on the assumption of R = 0 cannot be used directly to interpret results from experimental observations of systems with a large viscosity contrast. Stability analysis for such systems need to account for these differences. Most experimental studies report the time for the onset of nonlinear convec- tion. Fully resolved nonlinear simulations (Farajzadeh et al., 2013; Rapaka et al., 2008; Tilton & Riaz, 2014) demonstrate that the time for the onset of nonlinear effects depends on amplitude of initial perturbations and Ra, and is usually much greater than to. Our conclusions with regards to differences in the level of instabil- ity depending on the viscosity contrast and different models are expected to have a proportionate effect on the onset of convection. 97 Chapter 4: Natural convection in horizontally layered porous media 4.1 Overview This chapter explores the onset of natural convection in horizontally layered aquifers where permeability varies in the direction of gravity. Natural convection associated with CO2 sequestration in deep saline aquifers occurs due to the unsta- ble density stratification of diffusive boundary layers. The density-driven instability plays a key role in the rapid dissolution of CO2 into brine (Orr, 2009). The primary objective of this chapter is to develop a basic understanding of the instability mecha- nisms that govern the interaction of perturbations with permeability heterogeneity. Permeability is assumed to vary periodically across the thickness of the aquifer. Such a variation of permeability is relevant for saline aquifers (Bickle et al., 2007; Chadwick et al., 2005; Green & Ennis-King, 2010; Neufeld & Huppert, 2009) and is used to identify the general mechanisms of instability in heterogeneous porous me- dia. The periodic variation of permeability may also facilitate comparison with the classical Rayleigh-Be´nard convection in layered and periodic permeability porous media (Gjerde & Tyvand, 1984; Hewitt et al., 2014; McKibbin & O’Sullivan, 1980, 1981; Nield & Bejan, 2006). Previous studies of miscible flows in heterogeneous porous media highlight the 98 importance of the interactions between fluid flow and permeability heterogeneity. These studies relate to displacement processes for oil recovery (Christie et al., 1993; Tchelepi et al., 1993; Zhang et al., 1997), the flow of a contaminant jet into an aquifer (Schincariol, 1998; Schincariol et al., 1997) and the problem of natural con- vection due to a gravitationally unstable boundary layer (Prasad & Simmons, 2003; Rapaka et al., 2009; Simmons et al., 2001). The presence of correlated variability of permeability leads to complex regimes. As the variance level and the correlation length of heterogeneity increase, the details of the permeability field dictate the flow paths. (Camhi et al., 2000; Chen & Meiburg, 1998; Tan & Homsy, 1992; Tchelepi & Orr, 1994). A resonant amplification of instability was reported by DeWit & Homsy (1997) and Riaz & Meiburg (2004) when the length scale of the viscous instability is close to the correlation scale of the permeability variation. The onset of natural convection depends on the growth rate of perturbation amplitude within the linear regime. For the problem of CO2 sequestration, the transient nature of the diffusive boundary layer imparts an evolutionary character to the amplitude that depends on the instantaneous coupling between perturbations and the boundary layer. In the presence of heterogeneity, the coupling involves additional terms related to the permeability field, the effect of which has not been investigated for the onset of convection in the past. This study examines such effects using a combination of linear stability analysis and direct numerical simulations. The chapter is organized as follows. The governing equations and the perme- ability model are described in §4.2. The critical times for the onset of instability and the physical justification are presented in §4.3. The onset of nonlinear convection is 99 explained in §4.4. Finally, a summary of our findings are presented in §4.5. 4.2 Governing equations The mathematical model governing the flow of dissolved CO2 in a porous saline aquifer can be expressed in dimensionless form as ∇ · u = 0 , ct = −u · ∇c+ 1 Ra ∇2c , u = −k (∇p− czˆ) , (4.1) where, u(x, t) is the Darcy velocity, c(x, t) is the concentration, p(x, t) the pressure. In equation (4.1), Ra and k(z) are dimensionless quantities corresponding to the Rayleigh number and permeability field. The unit vector, zˆ, acts in the direction of gravity. The initial condition is the quiescent state, u = 0. Periodic boundary conditions are imposed on the vertical boundaries. The boundary conditions on the horizontal boundaries are, c ∣∣ z=0 = 1, cz ∣∣ z=1 = 0, w ∣∣ z=0 = w ∣∣ z=1 = 0 . (4.2) Equation (4.1) was non-dimensionalized as in Riaz et al. (2006) with the following characteristic values for length, time, velocity, and pressure, L = H , T = φµH k¯∆ρg , U = k¯∆ρg µ , P = µUH k¯ , (4.3) where k¯ is a characteristic value of permeability, φ is the porosity, D is the diffusion coefficient, and H the thickness of the aquifer. Concentration is scaled with respect to its maximum value found at z = 0. The Rayleigh number is defined as Ra = UH/(φD). The driving mechanism of density difference is represented by, ∆ρ = 100 ρ1 − ρo, where ρo and ρ1 refer, respectively, to the density of pure brine and CO2 saturated brine. The difference in density compared with the density of brine is small, ∆ρ/ρo ∼ ∆ρ/ρ1  1 T˙his justifies the use of the Boussinesq approximation and a linear density profile, ρ = ρo+∆ρc, for obtaining (4.1). Viscosity, µ, is assumed to be a constant (Bando et al., 2004; Fleury & Deschamps, 2008). The dimensionless permeability field to assumed to vary periodically in the vertical direction according to k(z) = exp [√ 2σ2 cos(mz + γ) ] . (4.4) The term, √ 2σ2, indicates the amplitude of the spatial permeability oscillation, where σ2 is the variance of ln(k). The wavelength of permeability oscillation is m, and γ indicates the phase with respect to z = 0. The cosine permeability variation allows the spatial distribution to be smooth with a constant wavelength, 2pi/m. When k(z) is normalized according to ∫ 1 0 k(z) dz = 1, as in this work, k¯ corresponds to the average of the permeability field. The permeability model given by (4.4) includes the essential features of the spatial correlation of the permeability field and its variance (Delhomme, 1979; Freeze, 1975). The variance associated with the spatial distribution of the natural logarithm of the permeability field has been reported to be within the range, 0.5 to 5 (Clifton & Neuman, 1982), while the spatial correlation scales are often observed to be large horizontally and much smaller vertically (Dagan, 1984), indicating a layered permeability structure. Porosity variation is not considered in this study because it is generally much less than the permeability variation (Hoeksema & Kitanidis, 1985). 101 A linear stability analysis is used to identify the structure and growth of finite amplitude perturbations that lead to the onset of nonlinear effects. To perform a linear stability analysis, perturbations are imposed in the form of vertical shape functions that vary periodically in the x and y directions, (u, p, c)(x, t) = co(z, t) + (u˜, p˜, c˜)(z, t) exp [i(`xx+ `yy)] , (4.5) where co is the concentration base state and u˜, p˜ and c˜ denote the perturbation profiles in the z-direction. Perturbation wavenumbers in the x- and y-direction are denoted by `x and `y, respectively, and i = √ −1. The base state, co(z, t), is the solution of the diffusion equation, ct = czz/Ra, that follows from the initial condition, u = 0, and the concentration boundary condi- tions, (4.2). This time varying base state represents a diffusive boundary layer that originates at z = 0 and diffuses downwards. By substituting (4.5) into (4.1), elimi- nating pressure and collecting linear terms, the following system for perturbations, c˜ and w˜, are obtained, c˜t − 1 Ra ( c˜zz − `2c˜ ) = −cozw˜ , (4.6) w˜zz − (ln k)zw˜z − `2w˜ = −k`2c˜ , (4.7) where, ` = √ `2x + `2y. The boundary conditions are c˜ ∣∣ z=0 = c˜z ∣∣ z=1 = w˜ ∣∣ z=0 = w˜ ∣∣ z=1 = 0 . (4.8) The above set of equations represents a linear initial value problem (IVP) with respect to the perturbations, c˜ and w˜. The IVP can be expressed more compactly 102 as ∂c˜ ∂t = A(t; `)c˜ , c˜(z, to) = c˜o(z) ; z ∈ {0, 1} , (4.9) where to is the time at which perturbations are introduced and A(t; `) is the linear operator that results from the elimination of w˜ between (4.6) and (4.7). The IVP can be marched in time to determine the evolution of the initial condition, c˜o(z). As in chapter 3, the IVP (4.9) is studied by spectrally transforming the under- lying operator A. Using the transformation, ξ = zα and tˆ = t, where α = √ Ra/4t and ξ ∈ {0,∞}, equations (4.6) and (4.7) are expressed in (ξ, t)-space which gives the new IVP ∂c˜ ∂t = B(t; `)c˜ , c˜(ξ, to) = c˜o(ξ) ; ξ ∈ {0,∞} . (4.10) The base state in (ξ, t)-space is self-similar, co = erfc(ξ), where erfc is the complementary error function. Homogeneous Dirichlet boundary conditions are employed for both c˜ and w˜. The IVP in (ξ, t)-space, (4.10), is equivalent to the IVP in (z, t)-space, (4.9), provided α > 3. When α > 3, the two initial value problems, (4.9) and (4.10), produce identical results. 4.3 Linear perturbation growth This section explores the effect of permeability on the onset of linear instability. The onset of instability is defined as the time at which the perturbation growth rate first becomes positive, ω > 0. The critical time for the onset of instability is defined as the minimum onset time of instability over all perturbation wavenumbers. A 103 perturbation wavenumber tim e 5 30 55 80 105 130 0 0.2 0.4 0.6 0.8 1 σ 2 =1σ2=0 σ2=0.1 (a) perturbation wavenumber tim e 5 30 550.2 0.4 0.6 0.8 1 σ2=0 σ2=1 σ2=0.35 (b) Figure 4.1: Marginal stability curves for Ra = 500 and m = 6pi, based on ω∞ (lines) and ωDMA (open symbols). (a) γ = 0 and (b) γ = pi. Critical conditions for the onset of instability, (tc, `c), are indicated by solid dots. Increasing the variance for γ = 0 results in greater instability, shifting the onset of instability to earlier times and larger wavenumbers. For the case of γ = pi, an increase in variance has the opposite effect. marginal state of stability corresponds to ω = 0. The parameters of interest are the permeability variance, σ2, phase, γ, the permeability wavenumber, m, in addition to the perturbation wavenumber, `, and time, t. The analysis is conducted for a constant Ra = 500. Results for different Ra are related through simple scaling relationships. 4.3.1 Influence of permeability variance and phase The effect of permeability variance and phase on the marginal stability curve is shown in figure 4.1 for Ra = 500 and m = 6pi. Two different values of the phase, γ = 0 and γ = pi, are shown in plots (a) and (b), respectively. The marginal stability curve, for which ω = 0, maps the boundaries of the stable (negative growth rate) 104 and unstable (positive growth rate) regions in the t-` plane. The lowest point on the curve, highlighted with a dot, indicates the critical conditions, (tc, `c), for the onset of instability, i.e., the earliest time, tc, at which the critical wavenumber, `c, becomes unstable. The critical time is greater than zero because perturbations are initially damped. The marginal stability curves also show the longwave cutoff, indicating that wavenumbers smaller than a critical value are damped. The marginal stability curves based on the growth rate, ω∞, the maximum eigenvalue of B, are indicated with lines in figure 4.1. Symbols represent the growth rate, ωDMA, obtained with the dominant mode analysis (DMA) of Riaz et al. (2006), namely, ωDMA = − 1 4t ( 4 + β2 ) + 1 a(t) √ pit ∫ ∞ 0 exp(−ξ2)w˜dξ , (4.11) where, β = `/α, α = √ Ra/4t and a(t) is the coefficient in the expansion, c˜(ξ, t) = a(t)ξe−ξ 2 . For problems with homogeneous permeability, when σ2 = 0, an analytical solution for w˜ can be used to obtain the following relation, ωDMA = − 1 4t ( 4 + β2 ) + 1 16 √ pi t β2eβ 2/2 [ 1− erf ( β 2 )]2 . (4.12) For heterogeneous cases, where σ2 > 0, w˜ can only be determined numerically. The integration in (4.11) is then carried out numerically to determine ωDMA. The results based on ωDMA are plotted to demonstrate the validity of DMA for the heterogeneous case. Figure 4.1(a) shows that the marginal stability curves based on ωDMA and ω∞ agree very well for wavenumbers up to the critical wavenumber. Note that the time required for calculating critical conditions using DMA is atleast an order of magnitude smaller than other methods. 105 Figure 4.1(a) shows that when γ = 0, larger values of permeability variance, σ2, lead to greater instability, i.e., smaller critical times, tc, larger critical wavenumbers, `c, and a larger spectrum of unstable wavenumbers. Figure 4.1(b) indicates that the effect of variance is reversed when γ = pi. The critical time, tc, shifts to larger values with increasing variance while the critical wavenumber, kc, decreases. The entire spectrum of unstable wavenumbers also shrinks. 4.3.2 Stability mechanisms A physical understanding of the effect of heterogeneity can be gained by con- sidering the vorticity, ∇× u, which is obtained from (4.1) as ∇× u = k∇c× zˆ +∇lnk × u . (4.13) The two terms on the right hand side of (4.13) indicate that vorticity is produced or diminished by the interaction of two effects. The first term is related to the misalignment of the concentration gradient with the direction of the gravity vector, zˆ. The second term is associated with the misalignment of the velocity field with the gradient of lnk. For the homogeneous case, σ2 = 0, only the first term is active. For σ2 > 0, both terms contribute to either amplify or diminish vorticity. Note that the vorticity production due to variable viscosity is given by R∇c × u, where R is the log mobility ratio (see Chapter 3). Although this term is similar to vorticity production by permeability gradients (equation 4.13), the physical effect should be different because of the transient nature of the concentration field as opposed to the fixed spatial variation of permeability. 106 z 0 0.1 0.20 0.5 1 (a) z 0 0.1 0.20 0.5 1 (b) z 0 0.1 0.20 1 2 (c) c˜/ ||c˜ || ∞ w˜ /|| c˜|| ∞ k Figure 4.2: Perturbation and permeability profiles for ` = 40, m = 6pi and t = 0.5. Lines represent, σ2 = 0 (solid), σ2 = 0.1, γ = 0 (dashed) and σ2 = 0.1, γ = pi (dash-dot). The base state, co, is plotted in (a) with symbols. It indicates that perturbations rapidly decay to zero outside the boundary layer. To interpret vorticity in the linear regime for a 2-D disturbance field, this study considers the y-component of vorticity, η(x, z, t) = iη˜(z, t) exp(i`x), which can be expressed as η˜ = `kc˜− 1 ` (lnk)′w˜z . (4.14) The vorticity, η˜, reflects the net effect of the coupling between perturbation and permeability profiles. The amplitude of perturbations, c˜ and w˜, is governed by the history of perturbation growth. In order to focus on the effect of the instantaneous coupling between perturbation and permeability profiles, the concentration pertur- bation, c˜, in (4.14) will be scaled with ||c˜||∞. The velocity perturbation, w˜, then also scales with ||c˜||∞ because the governing equations (4.6) and (4.7) are linear and do not contain the time derivative of w˜. Note that the resulting instantaneous value of η˜/||c˜||∞ is useful for comparing the effect of different permeability profiles for a fixed value of `. Note that this approach is independent of the specific scale, ||c˜||p, 107 for c˜ and w˜. Our interpretation and conclusions do not change when perturbations are scaled with either ||c˜||1 or ||c˜||2. The spatial profiles for c˜/||c˜||∞, w˜/||c˜||∞ and k for Ra = 500, are plotted in figure 4.2 for ` = 40, m = 6pi and t = 0.5. Perturbation profiles shown are based on the initial condition (2.44). The base state, co, is also plotted in figure 4.2(a) to indicate that c˜ and w˜ decay rapidly to zero outside the boundary layer. Figure 4.2(a) shows that the profiles for c˜/||c˜||∞ do no change appreciably when heterogeneity is introduced. Therefore, the differences in the corresponding velocity perturbations, w˜/||c˜||∞, plotted in figure 4.2(b), are mainly due to the differences in the corresponding permeability profiles, shown in figures 4.2(c). Figure 4.2(b) shows that compared with the homogeneous case, the amplitude of w˜/||c˜||∞ increases for γ = 0 but decreases when γ = pi. This behavior can be explained with the help of (4.7) where k appears as a source term and sets the amplitude of w˜. The perturbation-permeability interactions are shown in figure 4.3 with the help of, ηˆc = `kc˜/||c˜||∞ , ηˆw = w˜z(lnk)′/`||c˜||∞ , and ηˆ = ηˆc − ηˆw , (4.15) for Ra = 500, ` = 40, m = 6pi and t = 0.5. Figure 4.3(a) shows that ηˆc increases with an increase in the variance for γ = 0. This is because larger values of k coincide with the peak of c˜/||c˜||∞ when σ2 is increased, as shown in figures 4.2(a,c). Conversely, ηˆc decreases with an increase in σ2 when γ = pi, as shown in figure 4.3(d). This is because smaller values of k coincide with the peak of c˜/||c˜||∞ with an increase in σ2, as shown in figures 4.2(a,c). The profiles of ηˆw, shown in figures 4.3(b,d), are much 108 z 0 0.1 0.20 40 80 (a) z 0 0.1 0.2-4 0 4 8 (b) z 0 0.1 0.20 40 80 (c) z 0 0.1 0.20 40 80 (d) z 0 0.1 0.2-4 0 4 8 (e) z 0 0.1 0.20 40 80 (f) ηˆc ηˆw ηˆ ηˆc ηˆw ηˆ Figure 4.3: Spatial variation of vorticity contributions and net vorticity, ηˆc, ηˆw and ηˆ, defined in (4.15), for ` = 40, m = 6pi and t = 0.5. Lines represent, σ2 = 0 (solid), σ2 = 0.1 (dashed) and σ2 = 0.5 (dash-dot). The two cases, γ = 0 and γ = pi are shown, respectively, in plots (a,b,c) and (d,e,f). smaller in magnitude compared with ηˆc. The primary contribution to ηˆ therefore comes from ηˆc. To sum up the spatial variation of vorticity production, the net vorticity is defined as ∫ 1 0 ηˆ dz = ` ∫ 1 0 kc˜/||c˜||∞ dz − 1 ` ∫ 1 0 w˜z(lnk) ′/||c˜||∞ dz . (4.16) The first term on the right hand side of (4.16), Ic, is positive because c˜ and k are both positive, as shown in figures 4.2(a,c). The second term on the right hand side of (4.16), Iw, shown in figures 4.3(b,e), depends in a more complicated manner on the coupling between the spatial variations of (lnk)′ and w˜z. The values of Ic, Iw 109 γ = 0 σ2 Ic Iw I ω∞ 0 2.397 0.000 2.396 1.052 0.1 2.876 0.095 2.778 2.838 0.5 3.393 0.263 3.129 4.571 1.0 3.692 0.402 3.290 5.427 γ = pi σ2 Ic Iw I ω∞ 0 2.397 0.000 2.396 1.052 0.1 1.933 -0.042 1.975 -0.705 0.5 1.451 -0.026 1.477 -2.509 1.0 1.166 0.013 1.152 -3.460 Table 4.1: Net vorticity contributions, Ic, Iw and I, as a function of σ2 for ` = 40, m = 6pi and t = 0.5. The corresponding growth rates, ω∞, indicate that larger positive values of I are associated with greater insta- bility. Net vorticity, I, is determined primarily by Ic for this parameter set because Ic/|Iw|  1. and I (the left hand side of equation 4.16) can b found in Table 4.1 for Ra = 500, ` = 40, m = 6pi and t = 0.5. Table 4.1 indicates that I increases with an increase in the variance when γ = 0 and decreases with an increase in the variance when γ = pi. Corresponding values of the growth rate listed in Table 4.1 indicate that larger values of I are associated with greater instability and smaller values with less instability. Table 4.1 further indicates that Ic is an order of magnitude greater than |Iw| and therefore sets the magnitude of I. Hence, with an increase in σ2 the increase in I for γ = 0 and the decrease in I for γ = pi is related, respectively, to larger and smaller values of k in the boundary layer, as shown in figure 4.2(c). Permeability gradients play a relatively minor role for the parameter set considered in figure 4.1. 110 wavenumber tim e 0 25 50 75 100 125 0 0.5 1 1.5 (b) wavenumber tim e 0 25 50 75 100 125 0 0.5 1 1.5 (a) m 8pi 10pi 10.4pi 12pi σ2 0.9 1.0 1.1 1.2 Figure 4.4: Marginal stability contours for γ = 0. (a) Influence of σ2 for m = 10pi. (b) Influence of m for σ2 = 1. 4.3.3 Effect of correlation length For larger values of m, the contribution of Iw to net vorticity is expected to increase because of larger gradients of permeability within the boundary layer. To investigate this effect, figure 4.4(a) plots the marginal stability curves for Ra = 500, γ = 0, m = 10pi and different value of σ2. The overall stability behavior, as a function of σ2, is quantitatively different compared with the case of m = 6pi considered in figure 4.1(a). In the case of m = 10pi considered in figure 4.4(a), instability is found to decrease with an increase in σ2. This effect is more obvious at times greater than about t > 0.4, which indicates that the influence of the two modes of vorticity production changes in time due to the change in the thickness of the boundary layer. For σ2 = 1.1, the marginal stability curve splits into two separate unstable regions that continue to diminish with a further increase in the 111 z 0 0.1 0.20 0.5 1 (a) z 0 0.1 0.20 0.5 1 (b) z 0 0.1 0.20 1 2 3 (c) z 0 0.1 0.20 40 80 (d) z 0 0.1 0.2-20 -10 0 10 20 (e) z 0 0.1 0.2 0 40 80 (f) c˜/ ||c˜ || ∞ w˜ /|| c˜|| ∞ k ηˆc ηˆw ηˆ Figure 4.5: Spatial variation of perturbation and vorticity contributions for ` = 40, m = 10pi and t = 0.5. Lines represent, σ2 = 0 (solid), σ2 = 0.1 (dashed) and σ2 = 0.5 (dash-dot). variance. figure 4.4(b) plots the effect of m for σ2 = 1. As m is increased from 8pi to 10pi the marginal stability curve shrinks substantially. With a further increase in m to 10.4pi, the marginal stability curve again splits up, similar to the case for m = 10pi and σ2 = 1.1 shown in figure 4.4(a). The lower branch disappears entirely with a further increase in m to 12pi. The physical basis for reduced instability at larger values of m and σ2, indi- cated in figure 4.4, is explained with the help of figure 4.5. The profiles for scaled concentration perturbations observed in figure 4.5(a) are not sensitive to σ2. The corresponding peak velocity perturbations are similarly not sensitive to σ2, but the profiles move slightly inwards into the boundary layer when σ2 is increased, as 112 shown in figure 4.5(b). Permeability profiles plotted in figure 4.5(c) indicate that k undergoes a relatively smaller increase in the neighborhood of the peak of c˜/||c˜||∞ with an increase in σ2 compared with the case for m = 6pi shown in figure 4.2(c). Consequently, the increase in the corresponding vorticity contribution, ηˆc, is rela- tively smaller for m = 10pi as shown in figure 4.5(d) compared with the case for m = 6pi shown in figure 4.3(a). The permeability gradients on the other hand are now much larger and lead to significantly larger values of ηˆw with an increase in σ2, as shown in figure 4.5(e). The effect on ηˆ, plotted in figure 4.5(f), shows that vorticity is now confined to a narrower region of the boundary layer. This results in a decrease in the net vorticity, I = (2.40, 2.32, 2.23), with an increase in variance, σ2 = (0.1, 0.5, 1.0), which coincides with a corresponding decrease in the growth rate, ω∞ = (1.43, 1.22, 0.70). The overall effect of the interaction of the two mechanisms of vorticity pro- duction described above is summarized in figure 4.6, which plots contours of tc in the plane of σ2 and m. These results are obtained with the DMA, given by (4.11). Figure 4.6 indicates that tc decreases with an increase in σ2 for small values of m and increases with an increase in m for larger values of m. A transition from one type of behavior to the other is observed to occur at intermediate values of m. A sharp jump in the critical time, tc, occurs for large values of σ2 because the lower of the two stability branches, shown in figure 4.4, vanishes, and tc jumps to the larger value on the upper branch. For smaller values of σ2 the transition is gradual. Such transitions are associated with the relative strength of the vorticity produc- tion mechanisms, see equation (4.14), related to the magnitude and the gradient of 113 Figure 4.6: Effect of σ2 and m, on the critical time for the onset of insta- bility, tc, determined by the DMA, (4.11). Numbers indicate the mag- nitude of tc. Gray scale emphasizes the direction of increasing and de- creasing values of tc, with lighter (darker) color indicating higher (lower) values. permeability. The mode of instability switches from being dominated by the mag- nitude of permeability for smaller values of m to being dominated by the gradients of permeability for large values of m. Note that the effect of mode switching is most prominent when high permeability regions are located closer to the top boundary, z = 0, for −3pi/4 < γ < 3pi/8. In this range, the effects of mode switching observed in figures 4.4 and 4.6 extend to the onset of nonlinear convection, as described next in §4.4. 4.4 Onset of natural convection The onset of natural convection that marks the beginning of a period of en- hanced dissolution, occurs in response to perturbation amplification within the linear 114 regime. At the time of the onset of convection, the flux of CO2 into brine transi- tions from a diffusive to a convective regime. The critical time for the onset of convection is defined as the minimum onset time for the onset of convection over all perturbation wavenumbers. The onset of convection is associated with the onset of nonlinear effects that are triggered when the perturbation amplitude becomes sufficiently large. With the help of 2-D numerical simulation, this study explores the dependence of the onset time of nonlinear convection on the amplification of perturbations that are predicted using linear theory. It is shown that due to the dependence, the mode switching behaviors in the linear regime are also observed in the onset of convection. A direct numerical simulation (DNS) based on the vorticity streamfunction formulation is used to solve equations (4.1) (Riaz et al., 2006). The numerical method uses a spectral discretization in the x-direction, compact finite differences in the z-direction and a 3rd order Runga-Kutta, explicit time integration. Periodic boundary conditions are used in the x-direction. Initial conditions for the DNS are set as, u = w = 0 and c(x, z, to) = c o(z, to) + ε cos(`x)c˜o(z) , (4.17) where to is the time at which the system is perturbed, co(z, to) = erfc(z √ Ra/4to) is the base state for the linear stability analysis, ε is the initial perturbation amplitude, ` is the perturbation wavenumber and c˜o(z) is the initial perturbation defined in (2.44). The minimum amplitude is ε = 0, which cannot lead to instability because of the quiescent initial state, u = w = 0. As in Chapter 2, the maximum value 115 time flu x 0 2 4 6 80 0.01 0.02 0.03 0.04 0.05 (a) perturbation wavenumber tim e 10 20 30 401 2 3 4 5 6 7 8 (a) 8.5 629.29 68.23 ε=10−1 10−2 10−3 ε =10−1 10−2 10−3 Figure 4.7: (a) Flux, J(t), for Ra = 500, ` = 20, σ2 = 0, ar = 2pi/` and to = 0.01 for various values of the perturbation amplitude. Symbols represent the flux associated with the diffusive boundary layer (b). Time for convection onset, tn, as a function of ` obtained from DNS (solid lines). Amplification, Φ, at tn is only a function of ε. The critical onset time, tnc, (symbols) occurs on the path of maximum amplification, Φ∗(t), (dashed line) in the linear regime. of ε is constrained by the requirement, c(x, z, to) > 0; ∀(x, z), to ensure that no unphysical negative values of concentration occur at t = to and beyond. Tilton et al. (2013) derived an asymptotic scaling that showed how the onset time of convection approaches infinity as the amplitude goes to zero. 4.4.1 Critical time of convection onset The time at which convection initiates for the first time is typically associated with the point of deviation of the flux from the diffusive behavior. The flux of CO2 into brine is obtained by integrating the normal concentration gradient across the 116 boundary at z = 0 from x = 0 to the domain width, x = ar, J(t) = − 1 arRa ∫ ar 0 ∂c ∂z ∣∣∣∣ z=0 dx . (4.18) The time at which the onset of natural convection occurs is taken to be the time, tn, when dJ/dt = 0. The flux, J(t), obtained from the DNS for various values of the initial amplitude, ε, is shown in figure 4.7(a) for Ra = 500, ` = 20, σ2 = 0, ar = 2pi/` and to = 0.01. Symbols represent the diffusive flux in the linear regime, 1/ √ piRat. Figure 4.7(a) shows that the flux follows the diffusive path initially. Eventually, nonlinear interactions between the perturbed mode and its harmonics modify the mean concentration to deviate the flux from the diffusive behavior (Tilton & Riaz, 2014). The onset time tn is plotted as a function of ` in figure 4.7(b). A critical time for the onset of convection, tnc, can be defined as the minimum value of tn over all perturbation wavenumbers. In order to associate tnc with the dynamics in the linear regime, the amplification attained by a given perturbation over a specific time interval, {to, t}, is defined as, Φ(t; `) = ‖c˜(z, t; `)‖2 ‖c˜(z, to)‖2 = exp ∫ t to ω2(t; `)dt . (4.19) Note that the growthrates, ω2 (IVP), and the perturbation profile, c˜, are obtained using linear theory, while the onset of convection is calculated via DNS. The term on the right in (4.19) indicates explicitly that the amplification measures the combined effect of the instantaneous growth rates, summed over a specific period of time. Clearly, Φ(t; `) is a function of to, see Chapter 2. The initial time is to = 0.01 which is about an order of magnitude smaller than the onset time for instability in 117 homogeneous media. The maximum amplification at any given time can then be defined as, Φ∗(t) = sup ` {Φ(t; `)} . (4.20) Tilton & Riaz (2014) show that the amplification based on the linear analysis accu- rately predicts the nonlinear amplification at the onset of convection because of the weak nonlinearity at tnc. The critical time for the onset of convection thus occurs along the path followed by Φ∗(t) in (t, `)-space, as shown in figure 4.7(b) with the dashed line. The minimum level of amplification needed to trigger the onset is also noted for different values of ε. Next, it is considered how the specific profile of the path of Φ∗(t) and its magnitude influence the time for the onset of convection. 4.4.2 Influence of mode switching on convection onset The dependence of tnc on ε, Φ∗(t) and the heterogeneity parameters is exam- ined in figure 4.8 for Ra = 500, to = 0.01 and ar = 2pi/`. Figure 4.8(a) shows the effect of variance for three different values of σ2 and ε = 10−2. In each case, tnc transitions from a set of small to large values with an increase in m. The transition is gradual when σ2 = 0.1 but becomes progressively steeper for larger values of σ2. Figure 4.8(a) further shows that with an increase in the variance, tnc decreases to the left of the transition region but increases to the right. This transition is similar to the one observed in figure 4.6 with respect to the critical time for the onset of instability, tc. The behavior of tnc depicted in figure 4.8(a) is a consequence of the mode switching mechanisms in the linear regime as discussed in §4.3. 118 In order to understand the effect of mode switching on the behavior of tnc observed in figure 4.8(a), let us consider the temporal evolution of Φ∗(t) in figure 4.8(b-d). Figure 4.8(b) illustrates Φ∗(t) as a function of time for m = 4pi and two different values of σ2 = 0.1 and 0.5. The critical time for the onset of convection, tnc, corresponding to ε = 10−3, 10−2 and 10−1 is marked with symbols. Figure 4.8(b) shows that for all values of ε the amplification, Φ∗(t), is greater when the variance is larger. Therefore, the minimum amplification required for the onset is attained more quickly for the case of σ2 = 0.5 resulting in an earlier onset time. This explains the behavior of the decrease in tnc with an increase in σ2 for small values of m, as shown in figure 4.8(b). Note further in figure 4.8(b) that the amplification needed to trigger the onset of convection for a fixed value of ε is similar to that in the homogeneous case shown in figure 4.7(b). To examine the transition region noted in figure 4.8(b), let us consider the case of m = 6.5pi in figure 4.8(c). In this case, Φ∗(t) is greater for σ2 = 0.5 until about t = 3. A point of crossover occurs beyond this time where the amplification is greater for the case of σ2 = 0.1. As a consequence of the crossover, the critical onset time, tnc, for various values of σ2 depend additionally on ε. The crossover can be attributed to the mode switching effects that are brought about in this case by the shift in the relative strength of vorticity contributions as a result of boundary layer growth. For larger values of ε, tnc occurs earlier for σ2 = 0.5. On the other hand, for smaller values of ε, tnc occurs later for σ2 = 0.5 because the level of amplification needed to trigger the onset of convection is attained beyond the point of crossover. Finally, in the case of a large value of m = 12pi shown in figure 4.8(d), the crossover 119 m (2π) cr itic al on se t t im e 0 2 4 6 80 2 4 6 8 (a) σ2=0.1 0.25 0.5 ε=10−2 ε=10−3 time m ax im um a m pl ific at io n 10-2 10-1 100 101 10-1 100 101 102 103 104 (b) ε=10−1 time m ax im um a m pl ific at io n 10-2 10-1 100 101 10-1 100 101 102 103 104 (c) time m ax im um a m pl ific at io n 10-2 10-1 100 101 10-1 100 101 102 103 104 (d) Figure 4.8: (a) Influence of permeability variance, σ2 on the critical onset time, tnc, for Ra = 500, to = 0.01 and ε = 10−2. (b), (c) and (d) show Φ∗(t) vs. time for m = 4pi, 6.5pi and 12pi, respectively. Solid symbols mark tnc for three different values of ε. 120 occurs relatively early in time. Consequently, Φ∗(t) attains the threshold for the onset well beyond the point of crossover for all values of ε. This explains why tnc is delayed with an increase in variance for large m, as observed in figure 4.8(a). Note that in this chapter, the problem is studied with respect to a single value of the Rayleigh number. This is because a simple rescaling based on an internal length scale formed with buoyancy velocity and diffusivity can be used to obtain corresponding results at other values of the Rayleigh number, see §2.4.3 in Chapter 2. This type of rescaling is also appropriate for characterizing the onset of the nonlinear convection as long as the boundary layer does not interact with the bottom aquifer boundary. 4.5 Conclusions This study explores the instability behavior with respect to interactions be- tween the permeability and perturbation profiles within the transient boundary layer. Two main vorticity modes are discussed. The first mode is related to the coupling between the concentration field and the magnitude of permeability within the boundary layer. The second mode of vorticity is associated with the coupling of the velocity perturbation with the gradients of permeability within the bound- ary layer. The gradients of permeability become important when the wavelength of permeability oscillation is small compared with the thickness of the boundary layer. Perturbations in the velocity field either enhance or diminish the vorticity produced by the first mode, depending on the permeability phase. The time required by any 121 perturbation to amplify sufficiently and trigger the onset of nonlinear convection depends on the instantaneous contribution of each vorticity mode, which changes in time with an increase in the boundary layer thickness. Vorticity production can transition quickly from one mode the other in response to small changes in the per- meability field when the variance is high, resulting in abrupt changes in the onset time for convection. Our description of physical mechanisms provides a framework for the evalu- ation of practical problems. For example, heterogeneity may be characterized by multiple length scales and sharp changes in the permeability field whose effect on the onset of convection can be understood with respect to the interaction of individual vorticity modes. In the case of multiple permeability lengths scales, mode switching effects are expected to be present depending upon the magnitude of permeability gradients in the boundary layer. In the limit of high permeability and high variance, the periodic model can also account for abrupt changes in permeability. 122 Chapter 5: Natural convection in vertically layered porous media 5.1 Overview This chapter explores the onset of natural convection in a vertically layered porous medium in which permeability varies perpendicular to the direction of grav- ity. Past studies that have dealt with steady diffusive boundary layers (McKibbin, 1986; Nield, 1986) model the horizontally varying permeability field using discrete slabs. These studies also report higher instability compared to the constant perme- ability case. The distinguishing feature of the present stability problem, in addition to the transient nature of the base-state, is that the shape of the eigenmodes in the horizontal direction can no longer be represented by pure Fourier modes. Due to interaction with permeability field, perturbation modes are no longer independent of each other. The stability analysis therefore, cannot be performed using the clas- sical normal mode decomposition. Instead, this chapter utilizes a multi-dimensional eigenvalue problem (Theofilis, 2011). The chapter is outlined as follows. The governing equations are presented in §5.2. The linear stability results are shown in §5.3 and the onset of nonlinear convection in §5.4. The conclusions are discussed in §5.5. 123 5.2 Governing equations and methodology The geometry and modeling assumptions are similar to those in previous chap- ters. This study considers a two-dimensional isotropic porous medium of infinite width and height H. To analyze the instability characteristics, a periodic perme- ability is adopted, K∗ = Kg exp [a cos (kx)] , (5.1) where Kg = (KmaxKmin)0.5 is the geometric mean, Kmax and Kmin refer to the maximum and minimum values of permeability, a is the permeability amplitude, k is the permeability wavenumber, and x is the horizontal direction perpendicular to the direction of gravity. Note that permeability phase is not modeled because linear stability results are independent of phase. The variations in dispersivity, D, and porosity, φ, are assumed to be small compared to the permeability variations. Viscosity is treated as a constant and the density profile is linear of the form, ρ = ρ0 + ∆ρc, where ρ0 is the density of pure brine, ∆ρ is the density difference of CO2- saturated brine and pure brine. Because density differences are small, ∆ρ/ρ  1, Boussinesq approximation is applied. Using the characteristic length, Lc = H, time Tc = H/(φUc), permeability Kc = Kg, buoyancy velocity Uc = Kg∆ρg/µ and pressure Pc = ∆ρgH, one obtains the non-dimensional form of the governing equations, K v +∇p− cez = 0, ∇ · v = 0, ∂c ∂t + v · ∇c− 1 Ra ∇2c = 0, (5.2) 124 These equations are the Darcy’s law and the volume averaged forms of the continuity and concentration advection-diffusion equations respectively. The Rayleigh number is defined as Ra = UcH/(φD). The symbol v = [u,w] is the nondimensional velocity vector, c is the nondimensional concentration and p is the nondimensional pressure obtained from the dimensional pressure pˆ through the relation p = (pˆ − ρ0gz)/P. The symbol ez is the unit vector in the z direction. The boundary conditions for (5.2) are, c ∣∣ z=0 = 1, ∂c ∂z ∣∣∣ z=1 = 0, w ∣∣ z=0 = w ∣∣ z=1 = 0, t ≥ 0. (5.3) Equations (5.2) admit the transient base state, vb = 0, cb(z, t) = 1− 4 pi ∞∑ n=1 1 2n−1sin [( n− 1 2 ) piz ] exp [ − ( n− 1 2 )2 pi2t Ra ] , (5.4) The linear stability of base-state (5.4) is studied with respect to infinitesimal perturbations. The linear stability problem is formulated in terms of perturbations of the concentration field ĉ(x, z, t) and of the stream function ψ(x, z, t), ∂ĉ ∂t + ∂ψ̂ ∂x ∂cb ∂z − 1 Ra Dĉ = 0, Dψ̂ − 1 K ∂K ∂x ∂ψ̂ ∂x −K ∂ĉ ∂x = 0, (5.5) ĉ ∣∣∣ z=0 = 0, ∂ĉ ∂z ∣∣∣ z=1 = 0, ψ̂ ∣∣∣ z=0 = ψ̂ ∣∣∣ z=1 = 0, (5.6) ĉ ∣∣ x=0 = ĉ ∣∣ x=xp , ψ̂ ∣∣ x=0 = ψ̂ ∣∣ x=xp , (5.7) where D = ∂2/∂z2 +∂2/∂x2, and xp is the cut-off length in the horizontal direction. The concentration and stream function perturbations are split into spatial and temporal components, ĉ = ce(x, z)e σt, ψ̂ = ψe(x, z)e σt. (5.8) 125 Substituting (5.8) in (5.5)–(5.11) and after transforming the regular (x, z, t) space to the self-similar (x, ξ, t) space, where ξ = za is the self-similar variable of the diffusive boundary layer with a = √ Ra/4t, the resultant generalized eigenvalue problem may be expressed as, σce − ξ 2 ∂ce ∂ξ + a ∂ψe ∂x ∂cb ∂ξ − 1 Ra ( a2 ∂2 ∂ξ2 + ∂ ∂x2 ) ce = 0, (5.9) ( a2 ∂2 ∂ξ2 + ∂ ∂x2 ) ψe − 1 K ∂K ∂x ∂ψe ∂x −K∂ce ∂x = 0, (5.10) ce ∣∣ ξ=0 = ce ∣∣ ξ=ξc = 0, ψe ∣∣ ξ=0 = ψe ∣∣ ξ=ξc = 0, (5.11) ce ∣∣ x=0 = ce ∣∣ x=xp , ψe ∣∣ x=0 = ψe ∣∣ x=xp , (5.12) where ξc = 8 is the cut off length chosen such that the perturbations tend to zero at ξ = ξc, and xp = 2pi such that discrete integer wavenumbers are resolved in the x direction. The spatial derivatives in (5.9)–(5.10) are discretized using fourth order finite difference schemes. The two-dimensional (2D) eigenmodes are obtained using the sparse matrix eigenvalue solver ‘eigs’ in MATLAB. For a given t and Ra, the eigen- mode with the largest eigenvalue, σ, represent the dominant perturbation structure. The corresponding eigenvalue represents the dominant perturbation growth rate. Note that growth rate, σ, obtained in the self-similar (x, ξ, t) space is equivalent to perturbation growth rate in the regular (x, z, t) space when the perturbation amplitude is measured using the L∞ norm (Tilton et al., 2013). 126 (a) 0 1 2 3 4 5 6−1 −0.5 0 0.5 1 c e,K x (b) 0 1 2 3 4 5 6−1 −0.5 0 0.5 1 c e,K x Figure 5.1: Concentration eigenmode, ce (solid line), and permeability, K (dashed line), in the horizontal x direction for Ra = 500 and t = 2 for (a) Homogeneous porous media of a = 0. (b) Heterogeneous porous media of a = 0.1 and k = 10. 127 5.3 Linear growth characteristics 5.3.1 Quasi-steady 2D eigenmodes Figure 5.1 illustrates the concentration eigenmode, ce (solid line), and per- meability, K (dashed line), in the spanwise x direction for t = 2 and Ra = 500. Note that these concentration shapes are identical at any horizontal slice within the boundary layer. Panel (a) illustrates the profile shapes for homogeneous porous media with a = 0. The least unstable eigenmode varies sinuisoidally in the x direc- tion with wavenumber of 25. This is in excellent agreement with results obtained using a normal mode decomposition, see figure 2.44(d) in Chapter 2. Figure 5.1(b) illustrates the profiles for heterogeneous porous media with permeability amplitude, a = 0.1, and permeability wavenumber, k = 10. The concentration eigenmodes, ce, has a irregular wavy structure consisting of multiple sinusoidal or Fourier modes. It is unclear whether these modes are in phase with each other. To further investigate the structure of the quasi-steady eigenmodes, a Fourier transform is applied in order to examine the individual Fourier components of the eigenmode. Figure 5.2 illustrates the Fourier sine coefficients, al (solid line), and cosine coefficients, bl (dashed line), for t = 2, Ra = 500, and permeability amplitudes and wavenumbers of: a = 0 (panel a), a = 0.1 and k = 1 (panel b), a = 0.1 and k = 10 (panel c), and a = 0.1 and k = 50 (panel d). Figure 5.2(a) shows that for a homogeneous porous medium, a = 0, a single Fourier component is recovered at l = 25. This case corresponds to the eigenmode illustrated in figure 5.1(a). 128 (a) (b) 0 10 20 30 40 500 0.2 0.4 0.6 0.8 1 a l,b l l 0 10 20 30 40−0.2 −0.1 0 0.1 0.2 a l,b l l (c) (d) 0 10 20 30 40 50 −0.5−0.4 −0.3−0.2 −0.10 0.1 a l,b l l 0 20 40 60 80−1 −0.8 −0.6 −0.4 −0.2 0 a l,b l l Figure 5.2: Fourier sine coefficients, al (solid line), and cosine coefficients, bl (dashed line), for the horizontal variation of ce for t = 2 and Ra = 500 in homogeneous porous media (panel a) and in heterogeneous porous media with a = 0.1, and k = 1 (panel b), k = 10 (panel c), and k = 50 (panel d) respectively. 129 For a = 0.1 and k = 1 (figure 5.2b), the least stable eigenmode consist of pure modes in the frequency band, 18 < l < 35, clustered around the homogeneous dominant wavenumber of l = 25. The presence of both al and bl components indicates that the eigenmode contains Fourier modes or harmonics that are not in phase with each other. With increasing permeability wavenumber, k = 10 (figure 5.2c), the sinusoidal modes are located wider apart in the spectrum. Interestingly, the Fourier component with the maximum magnitude is still located close to the homogenous dominant wavenumber, l = 25. The other modes located at l = 14, 34, and 44 are much smaller in magnitude. This case corresponds to the eigenmode illustrated in figure 5.1(b). When permeability wavenumber is increased to k = 50 (figure 5.2d), there is only one Fourier component at l = 25. This indicates that with increasing permeability wavenumber, the quasi-steady 2D eigenmodes recover the dominant perturbation structures observed for homogeneous porous media. 5.3.2 Perturbation growth Figure 5.3 illustrates the effect of permeability wavenumber on the temporal evolution of growth rate, σ, for Ra = 500. Figure 5.3(a) illustrates σ versus t for the homogeneous case of a = 0 (circles) and the heterogeneous cases of a = 0.1 and k = 1 (crosses), k = 10 (squares), and k = 50 (diamonds) respectively. Perturbations experience an initial decay period, σ < 0, before experiencing positive growth rates, σ > 0, due to destabilizing effects. The smallest values for the growth rate are observed when a = 0, i.e. for homogeneous porous media. For a = 0.1, the largest 130 (a) (b) 0 0.5 1 1.5 2−1 −0.5 0 0.5 1 1.5 2 2.5 σ t 0 0.5 1 1.5 2−1 0 1 2 3 4 5 σ t Figure 5.3: Effect of permeability wavenumber on the temporal evolution of growth rate, σ, for (a) a = 0.1 and k = 1 (crosses), k = 10 (squares), and k = 50 (diamonds). (b) a = 0.5 and k = 10 (crosses), k = 50 (squares), and k = 100 (diamonds). The results for homogeneous porous medium for Ra = 500 is shown using circles. growth rates, σ, are observed for a smaller permeability wavenumber, k = 1. When permeability wavenumber is increased to k = 10, the magnitude of the growth rates decreases but remains larger than the homogeneous growth rates, a = 0. When k = 50, better agreement is observed with the homogenous case. With decreasing permeability length scales, the linear stability characteristics of a vertically layered porous medium are similar to that of a homogeneous medium. Figure 5.3(b) illustrates σ versus t for a larger permeability amplitude, a = 0.5, for the heterogeneous cases with k = 10 (crosses), k = 50 (squares), and k = 100 (diamonds) respectively and the homogeneous case of a = 0 (circles). For k = 10, one observes the largest deviation of the growth rate from the homogeneous case . With increasing permeability wavenumbers, k = 50 and 100, better agreement is found with the homogeneous case. With increasing permeability amplitude, the 131 (a) (b) 0 20 40 60 80 100 101 σ k 20 40 60 80 1001.7 1.8 1.9 2 2.1 2.2 σ k Figure 5.4: (a) Effect of permeability amplitude, a, on σ vs. k for t = 1 and Ra = 500 for a = 0 (solid line) , a = 0.2 (circles) , a = 0.5 (squares), and a = 1 (crosses). (b) Effect of time on σ vs k curves forRa = 500 and t = 1 (crosses), t = 2 (squares) and t = 4 (circles). instabilities are found to be more stronger. Our results suggest that the effect of horizontal permeability is to increase the instability of transient diffusive boundary layers. This instability enhancing effect of a horizontally varying permeability field is similar to the previous studies of McKibbin (1986) and Nield (1986) on steady diffusive boundary layers. Studies on viscous fingering in heterogeneous media (DeWit & Homsy, 1997; Riaz & Meiburg, 2004) have reported a resonance behavior when the length scale of the viscous fingering instability is close to that of the heterogeneous instability. To examine whether similar effects exist for the case of a gravitational instability, let us look at the perturbation growth rate for a fixed time, t = 1, for Ra = 500 and vary the length scale of the horizontal heterogeneity field. Figure 5.4(a) illustrates growth rate, σ, for the permeabiity wavenumbers, 1 < k < 100, for a = 0 (solid line) and a = 0.2 (circles), a = 0.5 (squares), and a = 1 (crosses). As observed earlier, 132 the maximum growth rates are bounded by the maximum value of permeability and occurs for large permeability length scales. In the limit of k approaching zero, k → 0, the growth rate is equal to that of a homogeneous porous media with a constant permeability of Kg exp(a). Resonance behavior is observed for heterogeneous media when k is large. For a = 0.2, after the initial decay, the growth rate starts to increase at k = 42, reaches a local maxima at k = 59, and begins to decrease again. The local maxima is approximately twice the value of the dominant perturbation wavenumber, `max = 28, for a homogeneous media of permeability Kg. For larger values of the permeability amplitude a, the resonance behavior occurs for higher permeability wave numbers k. Because of the unsteady nature of the boundary layer, these resonance effects are also time dependent. Figure 5.4(b) illustrates σ versus k curves for a = 0.1 and t = 1 (crosses), t = 2 (squares), and t = 4 (circles). Note that the maximum growth rates are observed at t = 2. It is known that the dominant wavenumbers for a homogeneous porous medium decrease monotonically with time, see figure 2.44(d). Consequently, the location of the resonance behavior, which is approxi- mately around permeability wavenumbers that are twice in magnitude to that of the homogeneous dominant perturbation wavenumbers, shifts with time. With in- creasing time, resonance occurs for smaller permeability wavenumbers. In §5.4, it is shown how these resonance behaviors observed in the linear regime affect the onset of nonlinear convection. 133 0 100 200 300 400 500 0 2 4 6 8 10x 10−3 σRa −1 tRa Figure 5.5: σRa−1 vs. tRa, for a = 0.5 and combinations of Ra = 500 and k = 10 (circles), Ra = 250 and k = 5 (crosses), and Ra = 750 and k = 15 (squares). 5.3.3 Scaling with Rayleigh number Figure 5.5 illustrates the temporal evolution of the perturbation growth rate for a fixed perturbation amplitude of a = 0.5 and different combinations of Ra = 500 and k = 10 (circles), Ra = 250 and k = 5 (crosses), and Ra = 750 and k = 15. The growthrates and times are scaled such that the plot shows curves of σRa−1 versus tRa. The results for different Ra and k collapse to a single curve. Consequently, all the results in this study can be generalized using the following scaling. If (σ, t, k) is known for a given Ra, then the corresponding values for Ra∗ are (σRa∗/Ra, tRa/Ra∗, kRa∗/Ra). 134 5.4 Onset of natural convection Two-dimensional direct numerical simulations (DNS) are performed using a vorticity stream function formulation of the nonlinear governing equations (5.2)– (5.3). This scheme uses a spectral discretization in the x-direction, compact finite differences in the z-direction and a 3rd order Runga-Kutta, explicit time integration. The horizontal domain is truncated to x ∈ [0, 2pi] with periodic boundary conditions on x = 0 and x = 2pi. The initial conditions are u = w = 0 with initial concentration field prescribed at t = 0.1, c(x, z) = cb(z) + ε cos(`x+ γ) ci(z) ‖ci(z)‖∞ , (5.13) where cb is the base state from the linear stability analysis, ε = 0.1 is the initial perturbation amplitude, ` is the perturbation wavenumber, γ is the perturbation phase, and ci(z) = z exp(−z2 √ Ra/(4t)), is the initial perturbation. This initial condition was found previously in Chapter 2 to be a good approximation of the optimal shape for linear growth. As in previous chapters, the onset time of natural convection, to, occurs when dJ/dt = 0, where J is the mean flux of CO2 into the brine given by, J(t) = − 1 2pi ∫ 2pi 0 1 Ra ∂c ∂z ∣∣∣ z=0 dx. (5.14) Simulations are first performed to measure to as a function of the perturbation wavenumber ` for a fixed permeability wavenumber k and perturbation phase γ. The critical time for the onset of convection is defined as follows tmino (k, γ) = min 0≤`<∞ to(`, k, γ). (5.15) 135 Figure 5.6 illustrates the critical onset times of natural convection, tmino (ver- tical axis) versus the permeability wavenumbers (horizontal axis) for Ra = 500 and for perturbation phases of γ = 0 (circles) and γ = pi/2 (crosses). The dashed line without symbols measures the onset times of linear instabilities, tc, i.e the time cor- responding to σ = 0, obtained using the 2D eigenvalue problem. The linear onset time, tc, decreases with increase in permeability wavenumber. As explained earlier in §5.3.2, the maximum linear growth rates are observed for k → 0. For permeability amplitude of a = 0.1 (panel a), the onset of instabilities is close to the onset time predicted by linear theory for homogeneous media, tc ∼ 0.34. The values for critical nonlinear onset times are much greater with tmino ∼ 1. For a = 0.5 (panel b), the boundary layer is more unstable. When k < 50, the values for the linear onset time, tc, and the nonlinear onset time, tmino , are smaller compared to a = 0.1. When k > 50, the values for tc are similar to the case of a = 0.1, but the corresponding values for nonlinear onset times, tmino ∼ 0.9, are smaller than the previous case of a = 0.1. The earlier onset times for tmino for larger permeability wavenumbers may be due to stronger resonance between instability and permeability structures, see §5.3.2. At small times, resonance effects are observed for much larger permeability wavenumbers while for large times, they are observed for smaller permeability wavenumbers. Because to occurs much later than tc, perturbation structures are more prone to resonance effects when k > 50. Resonance effects are notably enhanced when the permeability amplitude is increased to a = 1 (panel c). Nonlinear onset times, tmino , when k > 60, occur even earlier even though linear onset times, tc, have been delayed compared to the 136 a = 0.5 case. The variation of onset times due to different perturbation phase, γ, may signify the resonance interactions taking place between perturbation and permeability structures. 137 (a ) (b ) (c ) 20 40 60 80 00.20.40.60.81 20 40 60 80 00.20.40.60.81 20 40 60 80 00.20.40.60.81 F ig ur e 5. 6: tm in o (v er ti ca l ax is ) vs . p er m ea bi lit y w av en um b er s, k (h or iz on ta l ax is ) fo r R a = 50 0 an d fo r γ = 0 an d γ = pi /2 (c ro ss es ). T he da sh ed lin e (n o sy m b ol s) co rr es p on ds to th e on se t ti m e of lin ea r in st ab ili ti es ob ta in ed us in g th e 2D ei ge nv al ue pr ob le m . T he th re e fig ur es co rr es p on d to ca se s of p er m ea bi lit y am pl it ud e of (a ) a= 0. 1, (b ) a= 0. 5, an d (c ) a= 1 . T he on se t ti m e of in st ab ili ti es qu al it at iv el y re pr es en ts th e tr en ds in th e on se t of na tu ra l co nv ec ti on 138 5.5 Conclusions This chapter explores the linear instability of transient, diffusive boundary layers in a vertically layered isotropic porous media using a multi-dimensional eigen- value formulation. Due to interactions between permeability and perturbation struc- tures, the shape of dominant quasi-steady eigenmodes in the transverse direction may consist of multiple sinusoidally varying modes. This is in contrast to previous linear stability analysis on transient, diffusive boundary layers. The boundary layer is more unstable in a vertically layered porous medium and is bounded by the max- imum local value of permeability. The perturbation growth rates are always larger than the corresponding values for homogeneous media. Furthermore, resonant inter- actions between the permeability and perturbation fields are found to enhance the growth rates for large permeability wavenumbers. Because of the transient bound- ary layer, these resonant interactions are also time-dependent. The location around which these effects act vary from large permeability wavenumbers at small times to small permeability wave numbers at large times. Our results suggest that resonant interactions play an important role in determining the onset of nonlinear convection in aquifers with small permeability length scales. 139 Chapter 6: Contributions and Future work 6.1 Contributions The following journal articles have been published/submitted as part of this work on gravitationally driven instabilities in porous media: • “Optimal perturbations of gravitationally unstable, transient boundary lay- ers in porous media”, D. Daniel, N. Tilton, and A. Riaz, Journal of Fluid Mechanics, 727, 456-487, 2013. • “Effect of viscosity contrast on gravitationally, unstable, diffusive boundary layers”, D. Daniel and A. Riaz, Physics of Fluids, 26, 116601, 2014. • “Onset of natural convection in layered saline aquifers”, D. Daniel, A.Riaz, and H. Tchelepi, Journal of Fluid Mechanics, in review. • “Onset of natural convection in a vertically layered porous medium”, D. Daniel and A. Riaz, in preparation. In addition to the above works, the dissertation author D. Daniel has made scientific contributions via the following articles: 140 • “The linear transient period of gravitationally unstable, diffusive boundary layers developing in porous media”, N. Tilton, D. Daniel, and A. Riaz, Physics of Fluids, 25, 092107, 2013 Contribution: In the development and analysis of modal and nonmodal linear stability procedures. • “An empirical theory for gravitationally unstable flow in porous media”, R. Farajzadeh, B. Meulenbroek, D. Daniel, A. Riaz, J. Bruining, Computational Geosciences, 17, 515-527, 2013 Contribution: In the analysis of nonlinear computational simulations. 6.2 Future work The following areas have been identified for future exploration: • In chapter 2 of this study, an optimization procedure was employed to exam- ine the perturbation structure with the maximum amplification in the linear regime. Similarly, one could carry out an optimization procedure to find the optimal initial profile that results in the earlier onset of nonlinear convection. The new optimization procedure would consist of a weakly nonlinear analysis that can efficiently predict the onset of nonlinear convection. • This study predominantly explores gravity-driven instabilities of CO2 seques- tration using a fixed interface model which assumes a fixed two-phase interface between CO2 and underlying brine. On the other hand, experimental studies 141 of CO2 sequestration, see Chapter 3, model a moving two-phase interface. The degree to which either the fixed, or the moving interface model corresponds to the actual physical problem of two-phase flow remains to be determined because of the difficulty of setting up such an experiment and making quan- titative measurements. Further work in developing suitable models of CO2 sequestration is vital to understand CO2 storage mechanisms in subsurface saline aquifers, and consequently, improve estimates for the onset times of natural convection. • This study provides a robust foundation on the effect of horizontal and vertical variation of permeability on the stability of transient, diffusive boundary layers in porous media. Characterizing the stability behavior in random permeability fields is an avenue for future work. • This study explores the time scales associated with the onset of linear insta- bilities and the onset of natural convection. A further exploration into the regime beyond the onset of natural convection is beckoning and is currently missing in literature. 142 Bibliography Backhaus, S., Turitsyn, K. & Ecke, R., E. 2011 Convective instability and mass transport of diffusion layers in a Hele-Shaw geometry. Phys. Rev. Lett. 106, 104501. Bando, S., Takemura, F., Nishio, M., Hihara, E. & Akai, M. 2004 Vis- cosities of aqueous NaCl solutions with dissolved CO2 at (30-60) C and (10 to 20) MPa. J. Chem. Eng. Data 49, 1328–1332. Bear, J. 1988 Dynamics of Fluids in Porous Media. Dover Publications. Ben, Y., Demekhin, E. A. & Chang, H.-C 2002 A spectral theory for small- amplitude miscible fingering. Phys. Fluids 14 (3), 999–1010. Bickle, Mike, Chadwick, Andy, Huppert, Herbert E., Hallworth, Mark & Lyle, Sarah 2007 Modelling carbon dioxide accumulation at Sleipner: Implications for underground carbon storage. Earth. Planet. Sci. Lett. 255 (1-2), 164–176. Blair, L. M. & Quinn, J. A. 1969 The onset of cellular convection in a fluid layer with time-dependent density gradients. J. Fluid Mech. 36 (02), 385–400. Caltagirone, J. P. 1980 Stability of a saturated porous layer subject to a sudden rise in surface temperature: Comparison between the linear and energy methods. Q. J. Mech. Appl. Math. 33 (1), 47–58. Camhi, E, Meiburg, E & Ruith, M 2000 Miscible rectilinear displacements with gravity override. Part 2. Heterogeneous porous media. J. Fluid Mech. 420, 259–276. Chadwick, R. A., Arts, R. & Eiken, O. 2005 4D seismic quantification of a growing CO2 plume at Sleipner, North Sea. In 6th Petroleum Geology Conference, Geological Society London, , vol. 6, p. 13851399. Geological Society London. Chen, CY & Meiburg, E 1998 Miscible porous media displacements in the quar- ter five-spot configuration. Part 2. Effect of heterogeneities. J. Fluid Mech. 371, 269–299. 143 Cheng, P., Bestehorn, M. & Firoozabadi, A. 2012 Effect of permeability anisotropy on buoyancy-driven flow for CO2 sequestration in saline aquifers. Wa- ter Resourc. Res. 48 (9). Christie, M. A., Muggeridge, A. H. & Barley, J. J. 1993 3D simulation of viscous fingering and wag schemes. SPE Reservoir Eng. 8 (1), 19–26. Clifton, P. M. & Neuman, S. P. 1982 Effects of kriging and inverse modeling on conditional simulation of the Avra valley aquifer in Southern Arizona. Water Resourc. Res. 18 (4), 1215–1234. Dagan, G. 1984 Solute transport in heterogeneous porous formations. J. Fluid Mech. 145, 151–177. Delhomme, J. P. 1979 Spatial variability and uncertainty in groundwater-flow parameters - geostatistical approach. Water Resourc. Res. 15 (2), 269–280. DeWit, A & Homsy, GM 1997 Viscous fingering in periodically heterogeneous porous media .1. Formulation and linear instability. J. Chem. Phys. 107 (22), 9609–9618. Doumenc, F., Boeck, T., Guerrier, B. & Rossi, M. 2010 Transient Rayleigh– Be´nard–Marangoni convection due to evaporation: a linear non-normal stability analysis. J. Fluid Mech. 648, 521–539. Drazin, PG & Reid, WH 2004 Hydrodynamic Stability . Cambridge University Press. Ehyaei, D. & Kiger, K. T. 2014 Quantitative velocity measurement in thin-gap Poiseuille flows. Exp. Fluids 55 (4), 1–12. Elder, J. W. 1968 The unstable thermal interface. J. Fluid Mech. 32 (01), 69–96. Elenius, M. T., Nordbotten, J. M. & Kalisch, H. 2012 Effects of capillary transition on the stability of a diffusive boundary layer. IMA J. Appl. Math. 77 (6), 771–787. Ennis-King, J. & Paterson, L. 2005 Role of convective mixing in the long-term storage of carbon dioxide in deep saline formations. SPE J. 10, 349–356. Ennis-King, J., Preston, I. & Paterson, L. 2003 Onset of convection in anisotropic porous media subject to a rapid change in boundary conditions. Phys. Fluids 17, Article no. 084107. Farajzadeh, R., Meulenbroek, B., Daniel, D., Riaz, A. & Bruining, J. 2013 An empirical theory for gravitationally unstable flow in porous media. Computat. Geosci. 17 (3), 515–527. 144 Farajzadeh, R., Salimi, H., Zitha, P. L. J. & Bruining, H. 2007 Numerical simulation of density-driven natural convection in porous media with application for CO2 injection projects. Int. J. Heat Mass Tran. 50, 5054–5064. Farrell, B. F. & Ioannou, P. J. 1996a Generalized stability theory. Part I: Autonomous operators. J. Atmos. Sci. 53 (14), 2025–2040. Farrell, B. F. & Ioannou, P. J. 1996b Generalized stability theory. Part II: Nonautonomous operators. J. Atmos. Sci. 53 (14), 2041–2053. Fleury, M. & Deschamps, H. 2008 Electrical conductivity and viscosity of aque- ous NaCl solutions with dissolved CO2. J. Chem. Eng. Data 53 (11), 2505–2509. Foster, T. D. 1965 Stability of a homogeneous fluid cooled uniformly from above. Phys. Fluids 8 (7), 1249–1257. Freeze, R. A. 1975 Stochastic-conceptual analysis of one-dimensional groundwater flow in nonuniform homogeneous media. Water Resourc. Res. 11 (5), 725–741. Ghesmat, K., Hassanzadeh, H. & Abedi, J. 2011 The impact of geochemistry on convective mixing in a gravitationally unstable diffusive boundary layer in porous media: CO2 storage in saline aquifers. J. Fluid Mech. 673, 480–512. Gjerde, K. M. & Tyvand, P. A. 1984 Thermal convection in a porous medium with continuous periodic stratification. Int. J. Heat Mass Transfer 27, 2289–2295. Goldstein, A. W. 1959 Stability of a horizontal fluid layer with unsteady heating from below and time-dependent body force. Tech. Rep. NASA-TR-R-4. NASA. Green, C. P. & Ennis-King, J. 2010 Effect of vertical heterogeneity on long-term migration of CO2 in saline formations. Transport Porous Med. 82, 31–47. Green, L. L. & Foster, T. D. 1975 Secondary convection in a Hele Shaw cell. J. Fluid Mech. 71 (04), 675–687. Gresho, P. M. & Sani, R. L. 1971 The stability of a fluid layer subjected to a step change in temperature: Transient vs. frozen time analyses. Int. J. Heat Mass Tran. 14 (2), 207 – 221. Hewitt, D. R., Neufeld, J. A. & Lister, J. R. 2014 High Rayleigh number convection in a porous medium containing a thin low-permeability layer. J. Fluid Mech. 75, 844–869. Hoeksema, R. J. & Kitanidis, P. K. 1985 Analysis of the spatial structure of properties of selected aquifers. Water Resourc. Res. 21 (4), 563–572. Homsy, G. M. 1973 Global stability of time-dependent flows: impulsively heated or cooled fluid layers. J. Fluid Mech. 60, 129–139. 145 Horton, C. W. & Rogers, F. T. 1945 Convection currents in a porous medium. J. Appl. Phys. 16, 367–370. Huppert, H. E. & Neufeld, J. A. 2014 The fluid mechanics of carbon dioxide sequestration. Annu. Rev. Fluid Mech. 46 (1). Huppert, H. E., S., Turner J., N., Carey S., Stephen R., Sparks J. & A., Hallworth M. 1986 A laboratory simulation of pyroclastic flows down slopes. J. Volcanol. Geotherm. Res. 30, 179199. Jhaveri, B. S. & Homsy, G. M. 1982 The onset of convection in fluid layers heated rapidly in a time-dependent manner. J. Fluid Mech. 114, 251–260. Jones, G. & Fornwalt, H. J. 1936 The viscosity of aqueous solutions of elec- trolytes as a function of the concentration. III. Cesium iodide and potassium permanganate. J. Am. Chem. Soc. 58 (4), 619–625. Kaviany, M. 1984 Onset of thermal convection in a saturated porous medium: experiment and analysis. Int. J. Heat Mass Tran. 27 (11), 2101 – 2110. Kim, M.C. & Choi, C.K. 2011 The stability of miscible displacement in porous media: nonmonotonic viscosity profiles. Phys. Fluids 23 (8), 084105. Kim, M.C. & Choi, C.K. 2012 Linear stability analysis on the onset of buoyancy- driven convection in liquid-saturated porous medium. Phys. Fluids 24 (4), 044102. Kim, M.C. & Kim, S. 2005 Onset of convective stability in a fluid-saturated porous layer subjected to time-dependent heating. Int. Commun. Heat Mass 32, 416 – 424. Kim, M. C. & Choi, C. K. 2007 Relaxed energy stability analysis on the onset of buoyancy-driven instability in the horizontal porous layer. Phys. Fluids 19, 088103. Kumagai, A. & Yokoyama, C. 1998 Viscosities of aqueous solutions of CO2 at high pressures. Int. J. Thermophys. 19. Kumagai, A. & Yokoyama, C. 1999 Viscosities of aqueous nacl solutions con- taining co2 at high pressures. J. Chem. Eng. Data 44, 227–229. Lapwood, E. R. 1948 Convection of a fluid in a porous medium. Proc. Cambridge Philos. Soc. 44, 508–521. Lick, Wilbert 1965 The instability of a fluid layer with time-dependent heating. J. Fluid Mech. 21 (03), 565–576. MacMinn, C. W., Neufeld, J. A., Hesse, M. A. & Huppert, H. E. 2012 Spreading and convective dissolution of carbon dioxide in vertically confined, hor- izontal aquifers. Water Resour. Res 48 (11). 146 Manickam, O. & Homsy, G. M. 1995 Fingering instabilities in vertical miscible displacement flows in porous media. J. Fluid Mech. 288, 75–102. McKibbin, R. 1986 Heat transfer in a vertically-layered porous medium heated from below. Transport Porous Med. 1, 361–370. McKibbin, R. & O’Sullivan, M.J. 1980 Onset of convection in a layered porous medium heated from below. J. Fluid Mech. 96, 375–393. McKibbin, R. & O’Sullivan, M.J. 1981 Heat transfer in a layered porous medium heated from below. J. Fluid Mech. 111, 141–173. Meulenbroek, B., Farajzadeh, R. & Bruining, H. 2013 The effect of in- terface movement and viscosity variation on the stability of a diffusive interface between aqueous and gaseous CO2. Phys. Fluids 25, 074103. Morton, B. R. 1957 On the equilibrium of a stratified layer of a fluid. J. Mech. Appl. Math. 10, 433–447. Neufeld, J. A., Hesse, M., A., Riaz, A., Hallworth, M., A., Tchelepi, H. A. & Huppert, H. E. 2010 Convective dissolution of carbon dioxide in saline aquifers. Geophys. Res. Lett. 37, L22404. Neufeld, Jerome A. & Huppert, Herbert E. 2009 Modelling carbon dioxide sequestration in layered strata. J. Fluid Mech. 625, 353–370. Nield, D. A. 1986 Convective heat transfer in porous media with columnar struc- ture. Transport Porous Med. 2, 177–185. Nield, D. A. & Bejan, A. 2006 Convection in Porous Media, 3rd edn. Springer, New York. NOAA-ESRL 2014 CO2 data from Mauna Loa Observatory, http://co2now.org. Orr, F. M. 2009 Onshore geologic storage of CO2. Science 325, 1656–1658. Peyret, R. 2002 Spectral Methods for Incompressible Viscous Flows . Springer- Verlag. Prasad, A. & Simmons, C. T. 2003 Unstable density-driven flow in heteroge- neous porous media: A stochastic study of the Elder [1967b] short heater problem. Water Resourc. Res. 39 (1). Pritchard, D. 2004 The instability of thermal and fluid fronts during radial in- jection in a porous medium. J. Fluid Mech. 508, 133–163. Rapaka, S., Chen, S., Pawar, R. J., Stauffer, P. H. & Zhang, D. 2008 Non-modal growth of perturbations in density-driven convection in porous media. J. Fluid Mech. 609, 285–303. 147 Rapaka, S., Pawar, R. J., Stauffer, P. H., Zhang, D. & Chen, S. 2009 Onset of convection over a transient base-state in anisotropic and layered porous media. J. Fluid Mech. 641, 227–244. Reddy, S. C., Schmid, P. J. & Henningson, D. S. 1993 Pseudospectra of the Orr-Sommerfeld Operator. SIAM J. Appl. Math. 53 (1), 15–47. Riaz, A. & Cinar, Y. 2014 Carbon dioxide sequestration in saline formations: Part I- Review of the modeling of solubility trapping. J. Petrol. Sci. Eng. . Riaz, A., Hesse, M., Tchelepi, H. A. & Orr, F. M. 2006 Onset of convection in a gravitationally unstable diffusive boundary layer in porous media. J. Fluid Mech. 548, 87–111. Riaz, A. & Meiburg, E. 2004 Vorticity interaction mechanisms in variable- viscosity heterogeneous miscible displacements with and without density contrast. J. Fluid Mech. 517, 1. Robinson, J. L. 1976 Theoretical analysis of convective instability of a growing horizontal thermal boundary layer. Phys. Fluids 19, 778–791. Schincariol, R. A. 1998 Dispersive mixing dynamics of dense miscible plumes: natural perturbation initiation by local-scale heterogeneities. J. Contaminant Hy- drology 34 (3), 247–271. Schincariol, R. A., Schwartz, F. W. & Mendoza, C. A. 1997 Instabilities in variable density flows: Stability and sensitivity analyses for homogeneous and heterogeneous media. Water Resourc. Res. 33 (1), 31–41. Schmid, P. J. 2007 Nonmodal stability theory. Annu. Rev. Fluid Mech. 39, 129– 162. Selim, A. & Rees, D. A. S. 2007a The stability of developing thermal front in a porous medium. I. Linear theory. J. Porous Media 10, 1–16. Selim, A. & Rees, D. A. S. 2007b The stability of developing thermal front in a porous medium. II. Nonlinear evolution. J. Porous Media 10 (1), 17–34. Simmons, C. T., Fenstemaker, T. R. & Sharp, J. M. 2001 Variable- density groundwater flow and solute transport in heterogeneous porous media: approaches, resolutions and future challenges. J. Contaminant Hydrology 52 (1- 4), 245–275. Slim, A.C. & Ramakrishnan, T.S. 2010 Onset and cessation of time-dependent, dissolution-driven convection in porous media. Phys. Fluids 22 (12), 124103. Slim, A. C., Bandi, M. M., Miller, Joel C., & Mahadevan, L. 2013 Dissolution-driven convection in a Hele-Shaw cell. Phys. Fluids 25, 024101. 148 Spangenberg, W. G. & Rowland, W. R. 1961 Convective circulation in water induced by evaporative cooling. Phys. Fluids 4, 743–750. Tan, C. T. & Homsy, G. M. 1986 Stability of miscible displacements in porous media: Rectilinear flow. Phys. Fluids 29, 3549–3556. Tan, C T & Homsy, G M 1992 Viscous fingering with permeability heterogeneity. Phys. Fluids A 4 (6), 1099–1101. Tchelepi, H. A. & Orr, F. M. 1994 Interaction of viscous fingering, permeability heterogeneity, and gravity segregation in 3 dimensions. SPE Reservoir Eng. 9, 266–271. Tchelepi, H. A., Orr, F. M., Rakotomalala, N., Salin, D. & Woumeni, R. 1993 Dispersion, permeability heterogeneity, and viscous fingering - Acoustic experimental-observations and particle-tracking simulations. Phys. Fluids A 5 (7), 1558–1574. Theofilis, V. 2011 Global linear instability. Annu. Rev. Fluid Mech. 43, 319–352. Tilton, N., Daniel, D. & Riaz, A. 2013 The initial transient period of grav- itationally unstable diffusive boundary layers developing in porous media. Phys. Fluids 25, 092107. Tilton, N. & Riaz, A. 2014 Nonlinear stability of gravitationally unstable, tran- sient, diffusive boundary layers in porous media. J. Fluid Mech. 745, 251–278. Wessel-Berg, D. 2009 On a linear stability problem related to underground CO2 storage. SIAM J. Appl. Math. 70 (4), 1219–1238. Whitaker, S. 1999 The Method of Volume Averaging . Kluwer Academic Publish- ers. Wooding, R. A., Tyler, S. W. & White, I. 1997 Convection in groundwater below an evaporating salt lake: 1. Onset of instability. Water Resour. Res. 33 (6), 1199–1217. Xu, X., Chen, S. & Zhang, D. 2006 Convective stability analysis of the long-term storage of carbon dioxide in deep saline aquifers. Adv. Water Resour. 29. Zhang, H. R., Sorbie, K. S. & Tsibuklis, N. B. 1997 Viscous fingering in five-spot experimental porous media: New experimental results and numerical simulations. Chem. Eng. Sci. 52, 37. 149