arxiv:1602.02736v4 [stat.co] 16 jan 2018

25
Probabilistic modeling and global sensitivity analysis for CO 2 storage in geological formations: a spectral approach Bilal M. Saad a,b , Alen Alexanderian c,* , Serge Prudhomme d , Omar M. Knio a,e a King Abdullah University of Science and Technology, Division of Computer, Electrical and Mathematical Sciences & Engineering, 4700 KAUST, Thuwal 23955–6900, Kingdom of Saudi Arabia b The Institute for Computational Engineering and Sciences, The University of Texas at Austin, Austin, TX, USA c Department of Mathematics, North Carolina State University, Raleigh, NC, USA d Department of Mathematical and Industrial Engineering, ´ Ecole Polytechnique de Montr´ eal, Montr´ eal, Canada e Department of Mechanical Engineering and Materials Science, Duke University, Durham, NC 27708, USA Abstract This work focuses on the simulation of CO 2 storage in deep underground formations under uncertainty and seeks to understand the impact of uncertainties in reservoir properties on CO 2 leakage. To simulate the process, a non-isothermal two-phase two-component flow system with equilibrium phase exchange is used. Since model evaluations are computationally intensive, instead of traditional Monte Carlo methods, we rely on polynomial chaos (PC) expansions for representation of the stochastic model response. A non-intrusive approach is used to determine the PC coefficients. We establish the accuracy of the PC representations within a reasonable error threshold through systematic convergence studies. In addition to characterizing the distributions of model observables, we compute probabilities of excess CO 2 leakage. Moreover, we consider the injection rate as a design parameter and compute an optimum injection rate that ensures that the risk of excess pressure buildup at the leaky well remains below acceptable levels. We also provide a comprehensive analysis of sensitivities of CO 2 leakage, where we compute the contributions of the random parameters, and their interactions, to the variance by computing first, second, and total order Sobol’ indices. Keywords: Carbon sequestration, Multiphase flow, Risk assessment, Parametric Uncertainty, Polynomial Chaos, Sensitivity analysis 1. Introduction Carbon capture and storage (CCS) is an important topic related to the reduction of CO 2 pollution in the atmosphere. In general, CCS is process of capture and long-term storage of CO 2 . Different variants for CO 2 storage are being explored, with the storage in deep underground formation such as oil fields, gas fields, abandoned mines, and saline formations being of highest interest. Various risks exist in CO 2 sequestration in deep underground formations, the most important being (i) CO 2 leakage through caprock failure, faults, and abandoned wells; (ii) structural failure due to large pressure peaks; and (iii) brine displacement and * Corresponding author Email addresses: [email protected] (Bilal M. Saad), [email protected] (Alen Alexanderian), [email protected] (Serge Prudhomme), [email protected] (Omar M. Knio) Preprint submitted to Elsevier Received: date / Accepted: date arXiv:1602.02736v4 [stat.CO] 16 Jan 2018

Upload: others

Post on 14-Mar-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Probabilistic modeling and global sensitivity analysis for CO2 storage ingeological formations: a spectral approach

Bilal M. Saada,b, Alen Alexanderianc,∗, Serge Prudhommed, Omar M. Knioa,e

aKing Abdullah University of Science and Technology, Division of Computer, Electrical and Mathematical Sciences &Engineering, 4700 KAUST, Thuwal 23955–6900, Kingdom of Saudi Arabia

bThe Institute for Computational Engineering and Sciences, The University of Texas at Austin, Austin, TX, USAcDepartment of Mathematics, North Carolina State University, Raleigh, NC, USA

dDepartment of Mathematical and Industrial Engineering, Ecole Polytechnique de Montreal, Montreal, CanadaeDepartment of Mechanical Engineering and Materials Science, Duke University, Durham, NC 27708, USA

Abstract

This work focuses on the simulation of CO2 storage in deep underground formations under uncertainty andseeks to understand the impact of uncertainties in reservoir properties on CO2 leakage. To simulate theprocess, a non-isothermal two-phase two-component flow system with equilibrium phase exchange is used.Since model evaluations are computationally intensive, instead of traditional Monte Carlo methods, we relyon polynomial chaos (PC) expansions for representation of the stochastic model response. A non-intrusiveapproach is used to determine the PC coefficients. We establish the accuracy of the PC representationswithin a reasonable error threshold through systematic convergence studies. In addition to characterizingthe distributions of model observables, we compute probabilities of excess CO2 leakage. Moreover, weconsider the injection rate as a design parameter and compute an optimum injection rate that ensures thatthe risk of excess pressure buildup at the leaky well remains below acceptable levels. We also provide acomprehensive analysis of sensitivities of CO2 leakage, where we compute the contributions of the randomparameters, and their interactions, to the variance by computing first, second, and total order Sobol’ indices.

Keywords: Carbon sequestration, Multiphase flow, Risk assessment, Parametric Uncertainty, PolynomialChaos, Sensitivity analysis

1. Introduction

Carbon capture and storage (CCS) is an important topic related to the reduction of CO2 pollution in theatmosphere. In general, CCS is process of capture and long-term storage of CO2. Different variants for CO2storage are being explored, with the storage in deep underground formation such as oil fields, gas fields,abandoned mines, and saline formations being of highest interest. Various risks exist in CO2 sequestrationin deep underground formations, the most important being (i) CO2 leakage through caprock failure, faults,and abandoned wells; (ii) structural failure due to large pressure peaks; and (iii) brine displacement and

∗Corresponding authorEmail addresses: [email protected] (Bilal M. Saad), [email protected] (Alen Alexanderian),

[email protected] (Serge Prudhomme), [email protected] (Omar M. Knio)

Preprint submitted to Elsevier Received: date / Accepted: date

arX

iv:1

602.

0273

6v4

[st

at.C

O]

16

Jan

2018

infiltration into drinking water aquifers. Quantification of the risks is of paramount importance for deci-sion makers when evaluating the storage approaches before this technology can be implemented on largescale projects. In the case of deep geological storage of CO2, there have been significant research effortsdealing with mathematical and numerical models for simulating the CO2 injection processes into geologicalformations. Nordbotten et al. [1, 2, 3], presented in a series of papers the development of a semi-analyticalmodel to describe the space and time evolution of CO2 plumes and the leakage through abandoned wells. Areduced spatial dimension model based on vertical equilibrium was discussed by Nilsen et al. [4]. Ebigboet al. [5] set up benchmark examples in order to compare different modeling approaches such as numericaland semi-analytical models, for the problem of CO2 leakage. Class et al. [6] published a benchmark study,comparing a number of mathematical and numerical models with different complexities for problems relatedto CO2 storage in geologic formations.

CO2 sequestration is a complex multiphysics process, in which multiphase multicomponent flows playa critical role. The fact that the CO2 should be stored for many thousands of years implies that full scaleexperiments are not possible, and computer simulation is the main approach for exploring the feasibilityof different CO2 storage options. However the mathematical models of underground CO2 storage involvemany sources of geological uncertainties [7, 8]. These uncertainties are due to the limited knowledge aboutreservoir properties such as porosity and permeability. These sources of uncertainty lead to large variabil-ities in the predictive modeling of subsurface processes. Hence, one needs to propagate such uncertaintiesthroughout the calculations to quantify their impact on results of computer simulations. This requires theuse of stochastic modeling approaches.

Survey of literature on uncertainty quantification (UQ) for CO2 storage. In [9] the authors utilize a stochas-tic response surface method for assessment of leakage detectability for CO2 sequestration, by parameterizingthe spatially heterogeneous reservoir permeability using Karhunen–Loeve expansion. However, they usedthe analytical solution developed by Nordbotten et al. [2] to generate the pressure distribution at the injec-tion zone, which is then used to calculate the leakage flux into a confined aquifer using Darcy’s law. Theanalytical solution in [2] assumes that the phase saturations and fluid viscosities are constant within eachzone, that the capillary effects are small, and that vertical equilibrium applies to the entire flow system.In [10], the authors use polynomial chaos (PC) expansions for probabilistic analysis of the CO2 leakage ratein the CO2 benchmark presented by Class et al. [6]. In that article, the authors use a number of simplifyingassumptions to set up the mathematical model: fluid properties such as density and viscosity are constant, allprocesses are isothermal, CO2 and brine are immiscible phases, capillary pressure is negligible and mutualdissolution is neglected.

The article [11] provides estimates of the risk of brine discharge into freshwater aquifers following CO2injection into geological formations and resultant salt concentrations in the overlying drinking water aquifersusing arbitrary PC expansions combined with the probabilistic collocation method of [12]. Other worksinclude [13, 14] where the authors develop a screening and ranking method and a certification frameworkbased on effective trapping for geologic carbon sequestration, for selecting suitable storage sites on thebasis of health, safety, and environmental (HSE) risk resulting from CO2 or brine leakage. Similaritiesand differences between radioactive waste disposal and CO2 storage for performance assessment have beendiscussed in [15]. We also mention the paper [16] that presents a simple analytical method for the quickassessment of the CO2 storage capacity in closed and semi-closed systems to assess the expected pressurebuildup and CO2 storage capacity in such potentially pressure-constrained systems.

Spectral methods for UQ. In the present work, we rely on spectral UQ methods to build a surrogate modelfor the nonlinear function that maps the uncertain model parameters to the model observables. In particular,

2

we utilize PC expansions to build such surrogates. PC expansions, whose theory goes back to the late 30’sand 40’s [17, 18], have become an increasingly popular tool in recent years as they provide efficient meansfor performing UQ in computationally intensive mathematical models; see e.g., [19, 20, 21, 22, 23, 24, 10,25, 26, 27, 28, 29, 30] for a nonexhaustive sample of research contributions to numerical methods for UQusing PC expansions and applications of these methods to real world problems.

PC methods employ an approximation of the model variables in terms of a spectral expansion in anorthogonal polynomial basis. Once available, the PC representations can be used to efficiently approximatethe statistical properties of the model outputs. Generally, there are two approaches for computing a PCexpansion: (1) intrusive methods (see e.g. [19, 20, 21, 22, 23, 24]) and (2) non-intrusive methods (seee.g., [31, 24, 25]). Intrusive methods require a reformulation of the original uncertain partial differentialequations (PDEs) that govern the system, through a Galerkin projection onto the PC basis [32, 33]. Thisentails the need for rewriting the existing deterministic solvers. Subsequently, one has to solve a largersystem for the time/space evolution of the PC coefficients. Non-intrusive methods, on the other hand,provide a means to compute the spectral representation via a sampling of the existing deterministic solvers.In this paper, we will follow a non-intrusive approach to compute the coefficients in the PC expansion.

Our approach and contributions. Existing analyses of uncertainties in CCS using PC expansions, either relyon simplified physical models or do not rigorously establish the accuracy of the PC representations of themodel observables used for uncertainty analysis. The goal of this article is to further the understanding of theimpact of parametric uncertainties in the physical processes involved in CCS by using a more comprehensivephysical model, a rigorous numerical study of the accuracy of the computed PC representations for thequantities of interest, and a comprehensive analysis of the impact of parametric uncertainties in the physicalprocesses involved in CO2 storage, in the benchmark geological structure under study.

The fluid properties such as density, viscosity, and enthalpy of the CO2 and brine phases are expectedto change as the CO2 rises, affecting strongly CO2 arrival time to the leaky well and the leakage rate valueof the CO2. Therefore, we model these fluid properties as functions of the aquifer conditions, and use anon-isothermal two-phase two-component model to describe the flow processes of the leakage problem. Inaddition, we use nonlinear functions for the capillary pressure and the relative permeability for each phase.In Section 2, we outline the benchmark problem, and describe in detail the governing PDEs, our modelingassumptions, as well as the numerical solver used.

We rely on PC representations (see section 3 for the background material) to propagate the uncertain-ties in reservoir absolute permeability, permeability of the leakage well and reservoir porosity, and in theinjection rate on model observables of interest; see Section 4 for the description of the statistical model foruncertain parameters, and definition of the quantities of interest.

In section 5, we present a comprehensive analysis of the impact of parametric uncertainties in the phys-ical processes involved in CO2 storage, in the benchmark geological structure under study. A novel featureof the present work is a statistical analysis of the arrival time of the CO2 plume at the leaky well. This isimportant because monitoring CO2 arrival time in leaky wells and/or in observations wells is a key factorfor successful storage management to reduce risk of leakage and contamination of subsurface resources. Inaddition, we study the uncertainties in CO2 leakage through the leaky well, the maximum leakage ratio andthe time the maximum is attained, as well as the caprock pressure. A detailed computational study of theconvergence of the PC representations in distribution as well as in the sense of L2 is conducted. Moreover,using a hierarchy of quadrature rules of different resolutions, we establish that the quadrature rule used tocompute PC representations of the model observables has sufficient accuracy. Performing such convergencestudies, which is sometimes omitted in applications of spectral UQ methods in uncertainty quantification for

3

CO2 storage, is a crucial first step that establishes the accuracy and suitability of the PC surrogate for theanalysis that follows. We use the computed PC representations to understand the time dependent behavior ofthe statistical distribution of selected quantities of interest (QoIs). We also consider the statistical responseof the caprock pressure to the uncertain parameters, and devise a method for choosing an optimal injectionrate that results in minimal risk of excess pressure buildup at the leaky well.

To further understand the impact of the uncertain parameters, we provide a comprehensive analysis ofsensitivity of CO2 leakage to the uncertain parameters (see Section 5.6). This is achieved by computingthe Sobol’ indices [34, 35, 36, 37]. Traditional methods for computing the Sobol’ indices rely on com-putationally expensive sampling-based methods that require thousands of model evaluations. On the otherhand, PC expansions provide an efficient means to compute the Sobol’ sensitivity indices [38, 39, 26]; seealso [40, 41, 42] for application of PC-based sensitivity analysis to CO2 storage. We find that all the uncer-tain parameters under study have a significant impact on model variability, but that the balance of sensitivityindices changes over time. We also quantify the impact of the interactions between the uncertain parame-ters to variance. To this end, we compute second-order (joint) sensitivity indices that quantify the impactof pairwise interactions between the parameters. Moreover, to shed further light into the impact of overallinteractions among uncertain parameters to the variance, we introduce a modified variance-based sensitiv-ity measure, which we call the mixed index. This mixed index, which we describe in detail below, can becomputed at negligible computational cost, once a PC surrogate is available. We also analyze the globalsensitivity of the CO2 saturation to the uncertain inputs over the three-dimensional computational domain.

2. The mathematical model

2.1. Description of the benchmark problem

We consider the benchmark problem defined by Class et al. [6], which concerned with leakage of in-jected CO2 into the aquifer through a leaky well. This benchmark is set up based on the studies in [1, 2, 3].The focus is on a leakage scenario consisting of three hydrogeological layers—two aquifers separated by anaquitard —that are characterized by uniform thickness and homogeneous parameters. The model involvesone CO2 injection well and one leaky well. The leaky well is located at the center of the domain with theinjection well 100 m away. The domain has lateral dimensions of 1000 m × 1000 m. The sketch in Figure1 summarizes the model geometry and illustrates a 2D section of the 3D domain. The injected CO2 spreadswithin the aquifer and once it reaches the leaky well, it connects the two aquifers and rises to a shalloweraquifer. The two aquifers are each 30 m thick and the separating aquitard (caprock) is 100 m thick. Spatialheterogeneity is considered only through the different layers according to different geological media. Theformation has a permeability KA, and the leaky well, which is modelled as a porous medium, has a perme-ability KL > KA. Changes in fluid properties of CO2 are considered in this paper. Note that the CO2 andbrine fluid properties (e.g., density and viscosity) depend on the aquifer conditions, the temperature T, theCO2 pressure pc, the brine salinity ssalt = 0.1 kg NaCl per kg, and the mass fraction of CO2 in brine.

2.2. Governing equations

The physical process of CO2 injection in geologic reservoirs, including solubility trapping, is a non-isothermal two-phase two-component flow in porous media, which is governed by a system of couplednonlinear partial differential equations. In this model, the water-rich phase (brine, b) and the carbon dioxide-rich phase (CO2, c) consist of two components (water, w and CO2 component, n), as the solubility of thecomponents in the phases has to be taken into account.

4

aquifer

aquiferCO2 plume

aquitard (caprock)

leaky wellinjection well

100 m

30m

100

m30m

1000 m

Figure 1: Schematic view of the benchmark problem setup.

Local equilibrium phase exchange of the components in the phases is assumed to hold. Mass balance ofthe two components yields two partial differential equations for the components β in the phases α

φ∂t(ρmol,bxwb Sb +ρmol,cxw

c Sc)+div(ρmol,bxwb Vb +ρmol,cxw

c Vc)+div(Jwb +Jw

c ) = f w, (1)

φ∂t(ρmol,bxnbSb +ρmol,cxn

cSc)+div(ρmol,bxnbVb +ρmol,cxn

cVc)+div(Jnb +Jn

c) = f n. (2)

Here, we denote by φ the porosity, ρmol,α the molar density of phase α , Sα the α saturation, Vα the velocityphase α , Jβ

α a diffusive flux of the β component into the α phase, xβ

α the molar fraction of component β inphase α , f β a source term for the β component.

We also include the energy balance equation for thermal processes that may occur while the CO2 mi-grates through the formation. Under the assumption of local thermal equilibrium, only one energy balanceequation for the fluid-filled porous medium is necessary

φ∂ (∑α ραuαSα)

∂ t+(1−φ)

∂ρscsT∂ t

−div(λpm∇T

)+∑

α

divραhαVα= f h, α ∈ b,c, (3)

where ρα is the mass density of phase α , T is the temperature, uα is the specific internal energy, ρs and csare the density and the specific heat capacity of the porous medium, respectively, hα is the specific enthalpy,f h is the heat source term and λpm is the effective heat conductivity of the fluid-filled porous medium. Thesaturation of the α phases and the molar fractions (used to describe the composition of phases) satisfy

Sb +Sc = 1, xwb +xn

b = 1, xwc +xn

c = 1.

The relation between the phase pressures is given through the capillary pressure using the Brooks-Coreymodel [43]:

pcap(Sb) = pc− pb. (4)

The phase velocities Vα are given by the extended Darcy’s law for multiphase flow in porous media:

Vb =−Kkrb(Sb)

µb(∇pb−ρb ·g) , Vc =−K

krc(Sc)

µc(∇pc−ρc ·g) ,

where K denotes the absolute permeability tensor, krα denote the relative permeability functions, and g isthe gravity vector. Following Fick’s law, the diffusive flux of component β in phase α is given by

α =−Dβ

α ρmol,α∇xβ

α , Jwα +Jn

α = 0, (5)

5

Table 1: Fluid properties and simulation parameters.

Parameter Value/functionCO2 mass density, ρc f (T, pc) [44]Brine mass density, ρb f (T, pb,ssalt,xn

b) [45]CO2 viscosity, µc f (T, pc) [46]Brine viscosity, µb f (T,ssalt) [47]CO2 enthalpy, hc f (T, pc) [44]Brine enthalpy, hb f (T, pb,ssalt,Xn

b ) [45]Mutual solubilities, xβ

α f (T,ssalt, pc) [48]Brine salinity, ssalt 0.1 kg NaCl per kgResidual brine saturation, Srb 0.2Residual CO2 saturation,Src 0.05Relative permeability, krα Brooks and Corey [43]Capillary pressure, pcap(Sb) Brooks and Corey [43]Entry pressure, Pe 104 PaBrooks-Corey parameter, λ 2.0Leaky & injection well radius 0.15 m

where Dβ

α is the diffusion coefficient of component β in phase α .To close the system, the fluid properties of CO2 are calculated as functions of pressure and temperature.

The properties of brine additionally depend on the salinity and on the mole fraction of CO2 in brine. Detailedinformation on dependencies of the fluid properties is given in Table 1.

2.3. Simulation scenario

Since the aquitard is modeled as a layer of impermeable rock, for computational efficiency, only theaquifers and the leaky well are discretized. The boundaries between the discretized regions and the aquitardare modeled as no-flow boundaries. The initial conditions in the domain include a hydrostatic pressuredistribution that is dependent on the brine density, and a geothermal temperature distribution that dependson the geothermal gradient. The geothermal gradient is taken to be 0.03 K/m, and the initial temperature atthe bottom (at 3000 m depth) is 100

C. The aquifers are assumed to be initially saturated with brine. The

initial pressure at the bottom of the domain is taken to be 3.086 × 107 Pa. The lateral boundary conditionsare constant Dirichlet conditions and equal to the initial conditions. No-flow boundary conditions, for bothbrine and CO2, are assumed at the top and bottom of the domain.

In the benchmark setup, CO2 is being injected at a constant rate of 8.87 kg/s; this corresponds to 1600m3 per day at reservoir conditions. The total simulation time is 1500 days. All relevant parameters usedfor the simulation are given in Table 1. For more details we refer the reader to [6]. The CO2 leakage rate(denoted by Qleak in this paper), which is the output quantity of interest (QoI) of the benchmark study, isdefined as the total CO2 mass flow at midway between the top and bottom aquifers divided by the injectionrate, in percent.

We utilize the DuMuX simulator [49] to solve (1)–(5). For spatial discretization, we use the so-calledBox method [50], which is a vertex-centered finite volume method. For temporal discretization, we useimplicit Euler.

6

3. Background on spectral methods for uncertainty quantification

We begin our discussion of spectral UQ methods, by using the problem of uncertainty quantificationfor CO2 leakage, which we denote by Qleak = Qleak(t,q) where t denotes time. The vector q contains a setof parameters defining various physical properties of the system. These parameters appear as coefficients,or boundary or volume forcing terms in the governing PDEs. The elements of this vector are, in practice,known only approximatively and are hence considered uncertain. An important consideration in obtaininghigh-fidelity predictions is to quantify the impact of these parametric uncertainties on the model observablessuch as Qleak. To this end, we model the uncertain parameters as random variables that are parameterized bya vector θ of canonical random variables. Hence, we will have Qleak = Qleak(t,q(θ)). The random vector θfully characterizes the uncertain parameter vector q and, therefore, we can unambiguously use the simplernotation Qleak(t,θ) for the uncertain CO2 leakage, a convention which we follow for the other uncertainmodel variables below as well.

Below we seek to approximate the nonlinear mapping θ 7→ Qleak(t,θ) through a spectral representationof the form

Qleak(t,θ) =∞

∑k=0

ck(t)Ψk(θ),

where the Ψk’s form an orthogonal basis in an appropriate Hilbert space (discussed below), and ck(t) areexpansion coefficients. Such a spectral representation can then be used as a cheap-to-evaluate surrogatefor the parameter-to-observable map, θ 7→ Qleak(t,θ). This enables efficient methods for characterizing theuncertainties in Qleak that replace expensive PDE solves by cheap evaluations of the surrogate.

3.1. Notation and definitions

We denote by (Ω,F ,µ) a probability space, where Ω is the sample space, F is an appropriate σ -algebraon Ω, and µ is a probability measure. For a random variable θ on Ω, we write θ ∼U (a,b) to mean that θ

is uniformly distributed on the interval [a,b] and θ ∼N (0,1) to mean that θ is a standard normal randomvariable. We use the term iid for a collection of random variables to mean that they are independent andidentically distributed. The distribution function of a random variable θ on Ω is given by Fθ (x) = µ(θ ≤ x)for x ∈ R.

In the present work, we consider models with finitely many uncertain parameters. We parameterizethese uncertain parameters by a finite collection of real-valued independent random variables θ1, · · · ,θd thatare defined on Ω. We let Fθ denote the joint distribution function of the random vector θ = (θ1, · · · ,θd)

T .Since the θi are independent, Fθ(x) = ∏

di=1 Fi(xi) for x ∈ Rd , where Fi is the distribution function of the ith

coordinate.The random vector θ takes values in Rd . In fact, it is sufficient to consider the subset Θ of Rd given by

the support of the distribution function Fθ. Following common practice, we work in the image probabilityspace (Θ,B(Θ),Fθ), where B(Θ) is the Borel σ -algebra on Θ (which is a standard choice). For notationalconvenience we suppress B(Θ) below and denote the image probability space by (Θ,Fθ). We denote theexpectation of a random variable X : Θ→ R by

〈X〉=∫

Θ

X(s)Fθ(ds).

The space of square-integrable random variables on Θ, L2(Θ,Fθ), is endowed with the inner product (·, ·)defined by (X ,Y ) =

∫Θ

X(s)Y (s)Fθ(ds) = 〈XY 〉, and the corresponding induced norm ‖ · ‖= (·, ·)1/2.

7

3.2. Polynomial chaos expansion

In the case θiiid∼N (0,1), for i = 1, . . . ,d, any X ∈ L2(Θ,Fθ) admits an expansion of the form,

X =∞

∑k=0

ckΨk, (6)

where Ψk∞0 is a complete orthogonal set consisting of d-variate Hermite polynomials, and the series

converges in L2(Θ,Fθ). The infinite series representation of X is known as the polynomial chaos (PC) orWiener–Hermite expansion of X [17, 18, 32, 33]. The Wiener-Hermite expansion is the appropriate choicein the case the model parameters are parameterized by normally distributed random variables. In the casewhere the sources of uncertainty follow other distributions, alternative parameterizations and polynomialbases are adopted [33]. For example, in the case where θi

iid∼U (−1,1) the appropriate PC basis is given bythe d-variate Legendre polynomials.

Tensor product construction of a multivariate PC basis. Let θ = (θ1, . . . ,θd), where θi are independentrandom variables that are distributed according to common choices given by standard normal, uniform, orbeta distributions. We work with a multivariate PC basis that is obtained through a tensor product of appro-priate one-dimensional bases. More precisely, if we denote by ψ j(θi)∞

j=1 the one-dimensional orthogonalpolynomial basis corresponding to θi (with the choice of basis dictated by the distribution of θi), we formthe multivariate PC basis Ψk∞

k=0 as follows:

Ψk(θ) =d

∏i=1

ψαk

i(θi), θ ∈Θ, (7)

where αk =(αk

1 ,αk2 , · · · ,αk

d

)is a multi-index, and αk

i indicates the order of the 1D polynomials in θi.For example, if θi is standard normal, then ψ

αki

is the Hermite polynomial of order αki . With this basis,

any X ∈ L2(Θ,Fθ) admits an expansion of the form: X = ∑∞k=0 ckΨk, which is known as the generalized

polynomial chaos expansion of X . In computer implementations, we will approximate X(θ) with a truncatedseries,

X(θ)≈P

∑k=0

ckΨk(θ) (8)

where P is specified based on the choice of truncation strategy. In the present work, we consider truncationsbased on the total degree of the polynomials in the series. In this case, letting p be the largest (total)polynomial degree allowed in the expansion, it is straightforward to show that P = (d + p)!/(d! p!)− 1,where as before d is the dimension of the uncertain parameter vector θ.

Note that with X expanded as in (8), using the orthogonality of the basis ΨkP0 and the convention that

Ψ0 = 1, we have immediate access to its first and second moments:

〈X〉= c0, 〈X2〉=P

∑k=0

c2k〈ψ2

k 〉,

from which we also get varX= ∑Pk=1 c2

k〈ψ2k 〉.

8

Tests of convergence. To assess accuracy of a PC expansion, one could begin by studying convergence indistribution. A practical method of doing this is by tracking the probability density function (pdf) of thePC expansion (which can be approximated efficiently by sampling the expansion and using for examplea Kernel Density Estimation (KDE) method) as the order of the expansion is increased. Moreover, to getfurther confidence in the spectral representation of a random variable X(θ), one can use the relative L2 error,Erel, between X and its truncated PC representation:

E2rel :=

∫Θ

|X(s)−P

∑k=0

XkΨk(s)|2Fθ (ds)∫Θ

|X(s)|2Fθ (ds), (9)

which can be approximated using either quadrature or sample averaging.

3.3. Non-intrusive spectral projection

Let X belong to L2(Θ,Fθ). As mentioned in the introduction, non-intrusive methods aim at computingthe PC coefficients in the finite expansion (8) via a set of deterministic evaluations of X(θ) for specific real-izations of θ. Observe that since ΨP

0 form an orthogonal system, we have: (X ,Ψk) =(

∑Pl=0 clΨl,Ψk

)=

∑Pl=0 cl(Ψl,Ψk) = ck(Ψk,Ψk), so that the coefficient ck is given by

ck =〈X Ψk〉〈Ψ2

k〉.

The moments 〈Ψ2k〉 of known orthogonal polynomials can be computed analytically, and hence, the

determination of coefficients ck amounts to the evaluation of the moments 〈XΨk〉. In the non-intrusivespectral projection (NISP) approach, these moments are approximated via quadrature:

〈XΨk〉=∫

Θ

X(s)Ψk(s)Fθ(ds)≈Nq

∑j=1

ω jX(θ( j))Ψk(θ( j)), (10)

where θ( j) ∈ Θ and ω j are the nodes and weights of an appropriate quadrature formula. Note that in thisformulation, the same set of nodes is used to compute all coefficients ck. Hence, the complexity of NISP,measured in the number of evaluations of X(θ) (i.e., the number of model solves), scales with the number ofquadrature nodes Nq. These multi-dimensional quadrature rules are constructed by full or partial tensoriza-tion of one-dimensional quadrature formulas. Therefore, the number Nq of quadrature nodes scales with thedimension of the uncertain parameter—a phenomenon commonly referred to as the curse of dimensionality.In the present work, we work with a small number of uncertain parameters, and hence a full-tensor Gaussianquadrature was found feasible. However, for higher-dimensional problems, sparse grids, or adaptive sparsegrids are more suitable [51, 33, 27].

We remark that the efficient construction of PC expansions via non-intrusive methods has resulted insignificant research activity in recent years. The efforts include adaptive pseudo-spectral projections [52, 53]as well as regression-based approaches that incorporate sparsifying penalty methods [54, 28, 29]. While thegoal of the present work is not the study of such methods, nor their extensions, we point them out as potentialsolutions for the problems with higher-dimensional parameters, where one seeks to utilize PC expansionsfor uncertainty analysis.

9

3.4. Variance-based sensitivity analysis

An important step in quantifying the impact of parametric uncertainties on the response of an uncertainsystem is that of parametric sensitivity analysis. In particular, global or variance-based sensitivity analy-sis [34, 35, 36, 37] enable the characterization of the contribution of the individual uncertain parameters ortheir interactions to the total variance of the model response. In this section we outline the concepts fromvariance-based sensitivity analysis that are used in the present work.

Consider a square-integrable random variable X(θ). The first-order (or main effect) sensitivity indicesquantify the effect of the ith coordinate θi alone on the variance of the random variable X(θ). These first-order indices, which we denote by Si, are defined as follows,

Si =varEX(θ)|θi

varX(θ) , i ∈ 1, . . . ,d. (11)

Here EX(θ)|θi denotes the conditional expectation [55] of X(θ) given θi. While the mathematical defi-nition of the first-order indices (and higher-order indices discussed below) are given in terms of conditionalexpectations, whose numerical approximations involve expensive sampling (see e.g., [36]), their computa-tion via PC expansion is straightforward and very efficient [38, 39, 26, 56].

We also point out the second-order sensitivity indices that describe joint effects. Specifically, for i, j ∈1, . . . ,d, we denote by Si j the sensitivity index that quantifies the contribution of the interaction betweenθi and θ j to the total variance. The mathematical definition of Si j is as follows,

Si j =var

E

X(θ)|θi,θ j

varX(θ) − (Si +S j). (12)

Higher-order joint sensitivity indices (for example Si jk) can be defined also, but usually are not used inapplications. However, in our numerical computations below we will discuss a sensitivity index, which wecall the mixed index, that quantifies the contribution of all interactions among uncertain parameters.

Another useful variance based sensitivity measure is the total sensitivity index [35, 37]. The total sensi-tivity index due to θi is defined by,

Ti =varX(θ)−varEX(θ)|θ[−i]

varX(θ) (13)

where θ[−i] denotes the random vector θ = (θ1, . . . ,θd) with θi removed: θ[−i] := (θ1, . . . ,θi−1,θi+1, . . . ,θd).Notice that the numerator in (13) is the total variance minus the variance of the conditional expectationEX(θ)|θ[−i]. Thus, Ti is the total contribution of θi, by itself and through its interactions with othercoordinates, to the variance.

4. Uncertain parameters and quantities of interest

In the present study, we study the effect of uncertainties in reservoir porosity φ , reservoir absolutepermeability KA and permeability of the leakage well KL on the model response. To support decision-making based on the approach presented here, we also consider one design parameter, the CO2 injectionrate QCO2 . This will help to study the influence of the injection rate on the CO2 leakage rate. As reflectedin Table 2 all uncertain parameters are modeled as log-normal. The distributions were adopted based on thesetup in [8, 10].

10

Table 2: Distributions of the uncertain parameters.

Parameter Distributionlog-porosity N (−1.8971,0.22)log-absolute permeability N (−30.002,1.22)log-leaky well permeability N (−27.631,0.3679)log-injection rate N (2.1827,0.22)

In the analysis below, we focus on the following model observables that characterize the flow: (a) theCO2 leakage through the leaky well as a function of time, (b) arrival time of the CO2 plume at the leaky well,(c) the maximum leakage ratio, and (d) the corresponding time; these quantities are denoted, respectively,by Qleak, tarrival, Qmax

leak , and tmaxleak. Note that here the CO2 leakage rate is defined as in the benchmark studyas the CO2 mass flux, in percent, at midway between top bottom aquifer divided by the injection rate; also,Qmax

leak(θ) = maxt Qleak(t,θ). We also aim to understand the effect of model uncertainties on the spatiallydistributed pressure and saturation as functions of time.

5. Analysis of uncertainties in CO2 storage

In our computations, we used NISP based on a fully tensorized Gauss-Hermite quadrature to computethe spectral expansion of the model output in the PC basis. To enable a systematic analysis of convergence ofthe PC expansions, we used a hierarchy of quadrature grids. Namely, we constructed full tensor quadratureformulas using two, three, four, and five nodes in each stochastic dimension, resulting in non-nested quadra-ture grids of n4

q nodes, with nq ∈ 2,3,4,5. The required model evaluations were run on a 20-core IntelXeon E5-2680 v2 (2.80GHz) workstation. The computational time for a single simulation run was about 18hours using five cores. The highest resolution grid of 54 = 625 nodes supports a fourth-order PC expansion,which was found to provide sufficient accuracy for the statistical tests needed in our computations.

5.1. Analyzing uncertain response of CO2 leakage

In Figure 2 (left), we plot the realizations of Qleak as a function of time. These realizations are obtainedby 625 model solves with the parameter values set according to the 625 nodes of the Gauss-Hermite quadra-ture. To understand the solution behavior better, in Figure 2 (right) we report the sample mean of the CO2realization with the averaging done over the realizations computed at the quadrature points. Note that in thatfigure, we have used a log scale for the horizontal axis to provide a clearer picture of the dynamics of CO2at the early times.

The results reported in Figure 2 merely provide an initial screening. While a small Monte Carlo sample(in the order of the number of the chosen quadrature nodes) might be used for such an initial screening, themodel evaluations at the quadrature nodes enable construction of PC representations for the observables inthe expensive-to-simulate numerical model under study. It allows for a complete and reliable characteriza-tion of statistical properties of the model observables. In particular, the PC representations can be cheaplysampled, as many times as needed, in statistical studies.

To obtain a PC representation of CO2 leakage, we first project the log of CO2 leakage in a PC basis,

logQleak(t,θ)≈P

∑k=0

ck(t)Ψk(θ).

11

time [days]50 500 1000 1500

CO

2 le

aka

ge

[%

]

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1 15 30 60 180 365 15000

0.05

0.1

0.12

time [days]

averag

eCO

2leakage[%

]

Figure 2: Evoluation of CO2 leakage rate. Left: the curves depict the 625 realizations corresponding to the quadrature points inthe uncertain parameter space. Right: sample mean of the CO2 leakage rate over time (averaged over realizations computed atquadrature nodes).

The response surface for Qleak can then be constructed using

Qleak(t,θ)≈ exp

(P

∑k=0

ck(t)Ψk(θ)

). (14)

This log-projection, in particular, ensures the positivity of Qleak. Figure 3 shows instantaneous distributionsof CO2 leakage rate. These distributions are obtained by sampling the PC-based surrogate model (14) atselected times. As seen in the plots, the distributions seem to level off as the PC order is increased to p = 4suggesting that a fourth-order expansion is sufficient.

To further illustrate the evolution of the distribution of CO2 leakage over time, we show in Figure 4p-percentiles of the distribution at different times for p ∈ 5,25,50,75,95. These plots are generated bysampling the PC representation of CO2 leakage with a Monte Carlo sample size of 106. These results furtherillustrate the skewed distribution of CO2 leakage and its spread.

To get a more complete picture of the response of the model to parametric uncertainties, we use spectralrepresentations to approximate the (uncertain) arrival time of the CO2 plume at the leaky well (which isdefined as the time at which the leakage value is greater than 3.0×10−3%), the maximum leakage ratio andthe corresponding time; see Figure 5.

The use of nonlinear relative permeability-saturation relation, as done in this work, leads to a laterarrival time compared to the case of linear relative permeability [3, 6]. The reason behind this is that usingthe nonlinear relation, the sum of the relative permeability values of the brine and CO2 phases is less thanunity for most saturations; this leads to an overall reduced mobility of the flow and thus to a reduced leakageat early times with later arrival and lower peak as the leakage rate does not rise as high as in the case oflinear relative permeability [3, 6]. This later arrival time is seen in our numerical results, for example bylooking at the expected value of the arrival time, easily obtained using the PC representation of tarrival, thatis approximately 52 days. This should be contrasted with the arrival reported in [6], where averaging thearrival times computed using different numerical solvers that use the linear relative permeability-saturationrelation is about 9 days.

There are further effects contributing to the later arrival time of CO2 at the leaky well, like increasedinfluence of the viscous forces in the system due to the lower relative permeabilities, compared to buoyancydue to density differences, which makes the shape of the plume become more cylindrical [8]; see Figure6 (top) that illustrates the saturation of CO2 after 120 days, obtained for one realization model. Figure 6

12

0 5 · 10−2 0.10

100

200

300

CO2 leakage [%]

pdf

distribution of CO2 leakage after 60 days

p = 1p = 2p = 3p = 4

0 0.1 0.2 0.250

5

10

15

20

25

CO2 leakage [%]

distribution of CO2 leakage after 100 days

p = 1p = 2p = 3p = 4

0 0.1 0.2 0.250

5

10

15

20

25

CO2 leakage [%]

pdf

distribution of CO2 leakage after 500 days

p = 1p = 2p = 3p = 4

0 0.1 0.2 0.250

5

10

15

20

25

CO2 leakage [%]

distribution of CO2 leakage after 1500 days

p = 1p = 2p = 3p = 4

Figure 3: Distribution of CO2 leakage at selected times. p denotes the highest polynomial degree in the truncated expansion. Ineach case, the PC expansion was sampled 106 times to generate the distribution curve.

corresponds to a vertical slice through the middle of the domain. In that figure (bottom image), we alsoshow a typical realization of the pressure field along the same vertical slice.

As a result of the increased overall resistance to the flow, the leakage rate rises smoothly when the CO2reaches the well and then approaches steady-state. This behavior can be attributed to the lateral boundaryconditions that influence the pressure in the domain. In Figure 7, we report the time evolution of the meancaprock pressure, where we see a reduced pressure over time. This is the reason for the leakage rate to startdecreasing after the peak of the arrival of the CO2 flux at the leaky well [3].

5.2. Accuracy of the PC representation for CO2 leakage

The accuracy of our spectral representations, so far, have been examined by looking at the convergenceof the pdfs, indicating convergence in distribution. To further examine the accuracy of the spectral repre-sentation of the QoIs, we approximate relative L2 errors defined in (9). To avoid computing these errorsover the same 625 quadrature nodes used to construct the PC representations, we approximated the relativeL2 errors by quadrature, in the lower resolution grid with four points in each stochastic dimension. Sincethe quadrature grids are not nested, this approach provides a reliable cross validation of the computed PCrepresentations. Figure 8 (left) shows the relative L2 error for the scalar quantities of interest, tarrival, Qmax

leak ,tmaxleak, versus the PC order, and hence the number of quadrature nodes required for computing the PC ex-pansion, increases. Figure 8 (middle) shows the time evolution of the relative L2 errors for t > 120, at whichtime over 98% of realizations of the model output indicate arrival of CO2 plume to the leaky well.

From Figure 8 (left), we note that with a fourth-order PC expansion, the errors are at around 1% for log

13

60 200 500 1000 1200 15000

0.025

0.05

0.1

0.15

0.2

time [days]

CO

2leakage[%

]

5th & 95th percentiles25th & 75th percentiles

median

Figure 4: Percentiles of the distribution of CO2 leakage over time.

0 20 40 60 80 100 120 1400

1 · 10−2

2 · 10−2

3 · 10−2

arrival time [days]

p = 1p = 2p = 3p = 4

0 0.1 0.2 0.3 0.4 0.50

5

10

15

leakage value [%]

p = 1p = 2p = 3p = 4

−500 0 500 1,000 1,500 2,000 2,500 3,0000

1 · 10−3

2 · 10−3

3 · 10−3

4 · 10−3

time of max leakage [days]

p = 1p = 2p = 3p = 4

Figure 5: Distribution of the arrival time of the CO2 plume at the leaky well, the maximum leakage ratio and the correspondingtime of the maximum leakage ratio. p in the legend denotes the highest polynomial degree in the truncated expansion. In each case,the PC expansion was sampled 106 times to generate the distribution curve.

of tarrival and Qmaxleak , and around 2% for log tmaxleak; the errors are acceptable even for a third-order expansion.

Moreover, from Figure 8 (middle), we note that after the initial transient regime, the error for the fourth-order PC expansion for log-Qleak is about 1%. As before, we also note that a third-order PC expansionprovides a good balance between accuracy and computational cost.

The relative L2 errors reported correspond to the PC expansion for the log of the quantities of interest.Wealso examined the L2 error of the computed quantities of interest: with a fourth-order PC expansion, the es-timated relative L2 errors for tarrival, Qmax

leak , and tmaxleak were about 4%, 2%, and 15%, respectively. Moreover,the relative L2 error for the CO2 leakage over time was no more than around 8.5% for t > 120.

The idea of projecting the log of a quantity in a PC basis and approximating it by evaluating the expo-nential of the PC expansion was found to be a useful tool in simulating the distribution of the quantities ofinterest in the present study—it was observed to improve convergence in distribution as well as preservingpositivity of quantities of interest. However, we found that projecting the CO2 leakage directly into PCbasis provides acceptable accuracy also (see Figure 8 (right)), and is convenient to use for global sensitivityanalysis, presented later in this section.

We see that in this problem, the error in PC representation stabilizes over time, and a fixed (low) order

14

Figure 6: A typical realization of CO2 saturation (top), and a typical realization of pressure (bottom) at 120 days.

0 500 1,000 1,500 2,000306

308

310

312

314

time [days]

meancaprock

pressure

[bars]

Figure 7: Mean caprock pressure as a function of time.

PC expansion is suitable over the simulation time as the system tends to an equilibrium. While this phe-nomenon holds in many applications, we point out that there are important situations where a straightforwardapplication of PC methodology is not suitable and one needs to resort to techniques such as precondition-ing [25], asynchronuous integration [24], or techniques such as ones proposed in [57], when constructingspectral representations over time. A challenge that could occur in some problems (not observed in thepresent study) is a form of parametric stiffness that entails excitation of higher-order PC coefficients overtime, entailing the need for increasing the PC order as time increases; see e.g., the discussion in [25, 58].

We also provide a visual comparison between the “true” CO2 leakage, computed with the numericalmodel, and its approximation given by the PC model in Figure 9. We selected four realizations of the CO2leakage, from among the 256 realizations, which we used as validation data in the error study above. Foreach realization we evaluate the PC model at the same point. The plots show a best case (Figure 9(a)),where the PC model agrees well with the numerical model, and a worst case (Figure 9(b)), and two otherrealizations (bottom row). We also report the quadrature weight corresponding to each point in the title ofeach plot. Note that the worst case scenario corresponds to the smallest quadrature weight from among theones reported. To provide physical insight, we also report the parameter values corresponding to each plotin Table 3. Note that the realization in Figure 9(c) exhibits a very different physical response compared tothe other cases. This is due to the fact that this realization corresponds to a very low absolute permeabilityin the formation compared to leaky well permeability, which entails large levels of CO2 leakage.

15

1 2 3 4

10−2

10−1

PC order

relative

L2error

tarrivalQmax

leak

tmaxleak

120 500 1000 1500

10−2

10−1

100

time [days]

relative

L2error

p = 1p = 2p = 3p = 4

120 500 1000 1500

0.1

.08

.05

.03

time [days]

relative

L2error

Figure 8: Left: relative L2 error for the log- tarrival, Qmaxleak , tmaxleak, as PC order increases; middle: time evolution of relative L2 error

for the log-Qleak; right: relative L2 error of 4th order PC expansion for Qleak over time.

Table 3: Parameter values corresponding to the realizations in Figure 9, rounded to two decimal places. Here φ is porosity, KA isabsolute permeability, and KL is leaky well permeability.

Image φ [-] KA [mD] KL [mD] Injection rate [kg/s](a) 1.29×10−1 230.52 2391.78 7.65(b) 9.40×10−2 1558.16 2391.78 14.15(c) 2.39×10−1 5.75 1331.34 7.65(d) 2.39×10−1 38.85 429.29 14.15

We emphasize that whereas a single evaluation of the numerical model took 18 hours (using five cores),as mentioned above, evaluating a PC expansion has trivial computational cost (less than a second). Ourresults highlight the efficiency of PC expansions in building reasonably accurate and cheap-to-evaluate sur-rogate models.

5.3. Studying the impact of quadratureOur choice of the resolution of the quadrature rule is mainly guided by the requirement that the quadra-

ture formula should preserve the discrete orthogonality of the PC basis. That is, we require that (withinmachine precision)

(Ψi,Ψ j)≈Nq

∑k=1

Ψi(θk)Ψ j(θk)wk = δi j (Ψi,Ψi) .

This has guided the number of quadrature nodes we have placed in each stochastic dimension. However,for a given QoI X , the computation of PC modes requires evaluating (10) whose accuracy will be affectedby regularity of X . Hence, to gain confidence in our computations, we need to examine the effect of theresolution of the quadrature on the PC representation of the QoIs. In Figure 10, we study this by lookingat the distribution of the maximum leakage, which is a key QoI, when sampling its PC expansion of ordertwo and three, computed using quadrature formulas with increasing resolutions. These results indicate thatthe choice of the quadrature is appropriate to compute the PC expansions, for maximum leakage. Similarbehavior was observed with other quantities of interest.

5.4. Estimating probability of excess CO2 leakageAn important consideration in modeling CO2 leakage in reservoirs is understanding the likelihood of

excess CO2 leakage. In particular, we consider the probability,

Pexcess leakage(t) := prob(Qleak(t,θ)> Lmax),

16

120 500 1000 1500

0.02

0.04

0.06

time [days]

CO

2leakage[%

]

w = 4.296432× 10−3

True modelPC model

120 500 1000 1500

0.02

0.04

0.06

time [days]

CO

2leakage[%

]

w = 4.429316× 10−6

True modelPC model

(a) (b)

120 500 1000 15000

0.1

0.2

0.3

0.4

time [days]

CO

2leakage[%

]

w = 4.3402778× 10−4

True modelPC model

120 500 1000 1500

0.02

0.04

0.06

time [days]

CO

2leakage[%

]

w = 4.384571× 10−5

True modelPC model

(c) (d)

Figure 9: Four realizations of the leakage computed using the numerical model (red) and the corresponding realization using thePC model (black).

over time. Notice that computing such a probability is in general a computationally expensive task. However,using the cheap-to-evaluate PC representation of Qleak enables estimation of such probabilities at negligiblecomputational cost. Below we use Lmax = 0.123% and Lmax = 0.227%, which correspond, respectively, tothe maximum leakage ratio obtained in the benchmark study of CO2 leakage through an abandoned well[6] using a simplifying assumption to reduce the complexity of the equations to that obtained by the morephysically detailed equations used also in our study. Figure 11 (left) depicts the time-dependent behavior ofPexcess leakage(t). These results indicate that given our assumed statistical model for the uncertain parameters,the probability of Qleak exceeding 0.227% remains below 5%, but probability of Qleak exceeding 0.123%reaches values of more than 10%.

5.5. Choosing optimum design to reduce risk of failureHere we consider the caprock pressure at a point near the leaky well, and denote this quantity by

pcaprock(t,θ). Let us consider the quantity,

p∗caprock(θ) = pcaprock(t = 1500,θ).

We define the failure probability as that of p∗caprock exceeding a critical caprock pressure equal to 330bar. An optimal injection rate is the largest injection rate for which the failure probability remains below 5percent. To define this quantity mathematically, we denote

Πfail(QCO2) := Prob(p∗caprock(θ1,θ2,θ3;θ4(QCO2))> 330), with θ4(QCO2) = (QCO2− QCO2)/σQCO2,

17

0 0.1 0.2 0.3 0.4 0.50

5

10

15

leakage value [%]

nq = 3, p = 2nq = 4, p = 2nq = 5, p = 2

0 0.1 0.2 0.3 0.4 0.50

5

10

15

leakage value [%]

nq = 4, p = 3nq = 5, p = 3

Figure 10: The effect of quadrature on distribution of maximum leakage as approximated by PC expansions of order 2 (left) andorder 3 (right). Here nq denotes the number of Gauss–Hermite quadrature points in each stochastic dimension. Note that in eachcase, the total number of quadrature nodes is n4

q.

0 500 1,000 1,5000

0.025

0.05

0.08

0.1

0.15

time [days]

prob

abili

ty

Prob(Qleak(t) > 0.123%)

Prob(Qleak(t) > 0.227%)

1.8 2 2.2 2.4 2.6

2 · 10−2

4 · 10−2

6 · 10−2

8 · 10−2

0.1

log injection rate r

failure

probab

ility

p∗ = 0.05

r=

2.15

3

Figure 11: Left: Probability of excess leakage, as defined by leakage exceeding 0.123% (blue) and 0.227% (black), over time.Right:Curve showing Πfail(r) as a function of r. We notice that r≈ 2.153 gives the largest log-injection rate (which translates to aninjection rate of approximately 8.61 kg/s) that results in probability of failure being less than 5%.

where QCO2 and σQCO2are the mean and standard deviation of QCO2 chosen according to Table 2. Note

that to compute Πfail(QCO2) at a given rate QCO2 , we use the PC surrogate for p∗caprock, fix θ4 at θ4(QCO2)and, considering p∗caprock as a function of θ1, θ2, θ3, compute the probability p∗caprock > 330 bar using MonteCarlo Sampling, which can be done very efficiently using the PC surrogate. Then, we define the optimuminjection rate Q∗CO2

according to

Q∗CO2= max

QCO2

QCO2 : Πfail(QCO2)< 0.05.

Figure 11 (right) illustrates the choice of the injection rate based on a critical caprock pressure equal to 330bar after 1500 days. The results in Figure 11 (right) indicate that the maximum injection rate where thecaprock pressure does not exceed the limit of 330 bar is approximately 8.61 kg/s.

18

5.6. Global sensitivity analysis

In this section, we analyze the importance of each of the uncertain parameters to the uncertainties in theCO2 leakage. To this end, we perform a variance-based sensitivity analysis, where we find how much eachuncertain input parameter contributes to the total variance in CO2 leakage.

Figure 12 (left) depicts the time-dependent behavior of total sensitivity indices for CO2 leakage. Forclarity, we denote the total sensitivity indices for the random inputs by Tφ , TKA , TKL , and TQCO2

, correspondingto total sensitivity index for porosity, reservoir permeability, leaky well permeability and CO2 injection rate.The results in Figure 12 (left) indicate that at early time, the porosity and the CO2 injection rate have asignificant influence on the total variance. However, as the flow reaches the leaky well, the variance of CO2becomes dominated by the uncertainties in KA and KL.

We also report the first-order indices Sφ , SKA , SKL , and SQCO2in Figure 12 (right). While the first-

order indices show a similar trend as the total sensitivity indices, comparing their values with total indices,especially at early times, suggests that interactions between parameters have a significant contribution tovariability in CO2 leakage. To further understand this, we compute the second-order indices, which quantifythe contribution of the interaction between random parameters to the total variance; the results are reportedin Figure 13 (left). We note that at early times the interactions between the uncertain parameters have anoticeable contribution to the total variance, but as the time passes the second-order interactions mostlyvanish (except the interaction between KA and KL) and the first-order indices are almost equal to the totalindices reported in Figure 12 (left). That is, as the CO2 plume reaches the leaky well the response of thesystem becomes nearly additive in the uncertain parameters.

50 100 500 1000 15000

0.2

0.4

0.6

0.8

1

time [days]

fraction

ofvarian

ce

Tφ TKATKL

TQCO2

50 100 500 1000 15000

0.2

0.4

0.6

0.8

1

time [days]

fraction

ofvarian

ce

Sφ SKASKL

SQCO2

Figure 12: Variance based sensitivity analysis for CO2 leakage, Qleak. Left: total sensitivity indices over time. Right: first-ordersensitivity indices over time.

To quantify the total contribution of the interactions between uncertain parameters to the variance ofmodel output (CO2 leakage in the present case), we consider the following mixed sensitivity index,

Tmix :=variance due to interactions between parameters

total variance.

This mixed index can be defined in terms of conditional expectations (cf. Section 3.4), and is general difficultto approximate. However, using the PC representation of the model response (CO2 leakage here), we caneasily derive the following expression for Tmix. Using the multi-index construction of the multivariate PC

19

50 100 500 15000

0.1

0.2

0.3

0.4Sφ,KA

50 100 500 15000

0.1

0.2

0.3

0.4Sφ,KL

50 100 500 15000

0.1

0.2

0.3

0.4Sφ,QCO2

50 100 500 15000

0.1

0.2

0.3

0.4SKA,KL

50 100 500 15000

0.1

0.2

0.3

0.4SKA,QCO2

50 100 500 15000

0.1

0.2

0.3

0.4SKL,QCO2

50 100 500 1000 1500

0.1

0.3

0.5

0.7

1

time [days]

fractionofvariance

Tmix

Figure 13: Variance based sensitivity analysis for CO2 leakage, Qleak. Left: second-order sensitivity indices over time. Right: themixed sensitivity index Tmix over time.

basis in (7), we can define,

Tmix ≈∑k∈K c2

k ‖Ψk‖2L2(Ω)

∑Pk=1 c2

k ‖Ψk‖2L2(Ω)

,

where the index set K is defined by

K = k ∈ 1, . . . ,P : ‖αk‖0 > 1.

Note that here we have used the multi-index notation, used in construction of the PC basis, and denotedby ‖ · ‖0 the `0-“norm”. That is, for a vector x, ‖x‖0 is the number of nonzero elements of x. The indexTmix quantifies the contribution to the variance due to all interactions (second- and higher-order) between theuncertain inputs. Using the fourth-order PC expansion we have computed for CO2 leakage, we approximatethe mixed index for this QoI; see Figure 13 (right). The results reported in Figure 13 (right) show that,as also seen from the study of first- and second-order indices, there is significant contributions to modelvariability coming from interactions between the parameters at early times. These mixed-effect interactionslevel off at around 10% as the CO2 plume reaches the leaky well.

Finally, we compute the sensitivity indices over the three-dimensional computational domain. In par-ticular, we consider the sensitivity of CO2 saturation to the uncertain inputs. Figure 14 shows the spatialdistribution of the total sensitivity indices, Tφ and TKA for CO2 saturation at t = 70 days. A vertical slicethrough the middle of the domain indicates that the regions where porosity has a significant contributionto variance travel with the fronts of the CO2 plume, whereas the reservoir permeability maintains a nearlyconstant dominant effect on variance within the regions with high CO2 saturation.

6. Conclusions and Summary

A non-intrusive spectral projection approach was implemented to propagate and quantify parametricuncertainties for CO2 storage in geological formations using a common 3D leakage benchmark problem ofinjected CO2 into overlying formations through a leaky well. A non-isothermal two-phase two-componentflow system with equilibrium phase exchange is used. Moreover, we use nonlinear functions for the capillarypressure and the relative permeability for each phase.

In our numerical results, we find that the use of a nonlinear relative permeability-saturation relationin our mathematical model leads to an overall reduced mobility of the flow. This behavior is seen in the

20

Figure 14: Space propagation of sensitivities after 70 days. Here we consider the sensitivity of saturation to porosity φ (top) andreservoir permeability (bottom).

statistical distribution of the CO2 arrival time to the leaky well. In particular, tracking the time evolution ofthe distribution of the CO2 leakage, we see that the leakage rate starts decreasing after the peak of the arrivalof the CO2 flux at the leaky well. This is in contrast with the cases where one uses simplified assumptionssuch as linear relative permeabilities, which decreases the influence of the viscous forces in the system andresults in early arrival times of the CO2 plume at the leaky well [6].

We find, given our assumed statistical distributions for the random inputs, that the risk of CO2 leakagein excess of 0.123% could exceed 10% within the first two years of the simulation. However, this risk fallswell below 5%, when we consider a leakage threshold of 0.227%. In our computation of optimum injectionrate, we find that, given our assumed statistical distributions for the random inputs, an injection rate of 8.61kg/s still ensures low risk of failure (defined as excess pressure buildup at the leaky well).

In our sensitivity analysis, we find that the balance of sensitivities changes as a function of time, wherethe CO2 injection rate and porosity exhibit significant contribution to variance at early times, and becomeless important as the CO2 plume reaches the leaky well. On the other hand, the reservoir permeabilityand leaky well permeability become dominant contributors to the variance of CO2 leakage at later times.We also find that at early times the interactions among the different uncertain parameters has significantcontribution to variance, but as the CO2 plume reaches the leaky well, the bulk effect of interactions betweenthe parameters to the variance is due to the permeability of both the reservoir and the leaky well. The studyof sensitivity of saturation to uncertain inputs in the three-dimensional domain reveals that the regions whereporosity has a significant contribution to the variance travel with the fronts of the CO2 plume; however, thereservoir permeability maintains a nearly constant dominant effect on variance within the regions with highCO2 saturation.

Acknowledgement

Research reported in this publication was supported by the King Abdullah University of Science andTechnology (KAUST) under the Academic Excellency Alliance (AEA) UT Austin-KAUST project ”Un-certainty quantification for predictive modeling of the dissolution of porous and fractured media”. Com-putational resources for the simulations presented in this publication have been made available by KAUSTResearch Computing and KAUST SuperComputing Lab. Bilal Saad is grateful for the support by the SaudiArabia Basic Industries Corporation (SABIC). Bilal Saad, Serge Prudhomme, and Omar Knio are alsoparticipants of the KAUST SRI Center for Uncertainty Quantification in Computational Science and Engi-neering.

21

References

References

[1] J. N. Nordbotten, M. A. Celia, S. Bachu, Analytical solutions for leakage rates through abandonedwells, Water Resour. Res. 40 (4) (2004) W04204.

[2] J. N. Nordbotten, M. A. Celia, S. Bachu, Injection and storage of CO2 in deep saline aquifers: analyti-cal solution for CO2 plume evolution during injection, Transport in Porous Media 3 (2005) 339–360.

[3] J. N. Nordbotten, M. A. Celia, S. Bachu, H. Dahle, Semi-analytical solution for CO2 leakage throughan abandonedn well, Environmental science and technology 2 (2005) 602–611.

[4] H. Nilsen, P. Herrera, M. Ashraf, I. Ligaarden, M. Iding, C. Hermanrud, K. A. Lie, J. N. Nordbotten,H. Dahle, E. Keilegavlen, Field-case simulation of CO2 plume migration using vertical-equilibriummodels, Energy Procedia 4 (2011) 3801–3808.

[5] A. Ebigbo, H. Class, R. Helmig, β leakage through an abandoned well: problem-oriented benchmarks,Computational Geosciences 11 (2) (2007) 103–115.

[6] H. Class, A. Ebigbo, R. Helmig, H. Dahle, J. N. Nordbotten, M. A. Celia, P. Audigane, M. Darcis,J. Ennis-King, Y. Fan, B. Flemisch, S. Gasda, M. Jin, S. Krug, D. Labregere, A. Naderi, R. J. Pawar,A. Sbai, G. T. Sunil, L. Trenty, L. Wei, A benchmark-study on problems related to CO2 storage ingeologic formations, Computational Geosciences 13 (2009) 451–467.

[7] A. Hansson, M. Bryngelsson, Expert opinions on carbon dioxide capture and storage a framing ofuncertainties and possibilities, Energy Policy 37 (2009) 2273–2282.

[8] A. Kopp, H. Class, R. Helmig, Investigations on CO2 storage capacity in saline aquifers - part 1:Dimensional analysis of flow processes and reservoir characteristics, Int. J. of Green-house Gas Control3 (2009) 263–276.

[9] A. Y. Sun, M. Zeidouni, J. P. Nicot, Z. Lu, D. Zhang, Assessing leakage detectability at geologic CO2sequestration sites using the probabilistic collocation method, Advances in Water Resources 56 (2013)49–60.

[10] S. Oladyshkin, H. Class, R. Helmig, W. Nowak, An integrative approach to robust design and proba-bilistic risk assessment for CO2 storage in geological formations, Computational Geosciences 15 (3)(2011) 565–577.

[11] L. Walter, P. J. Binning, S. Oladyshkin, B. Flemisch, H. Class, Brine migration resulting from CO2injection into saline aquifers- an approach to risk estimation including various levels of uncertainty,International Journal of Greenhouse Gas Control 9 (2012) 495–506.

[12] S. Oladyshkin, H. Class, R. Helmig, W. Nowak, A concept for data-driven uncertainty quantificationand its application to carbon dixide storage in geological formations, Advances in Water Resources 34(2011) 1508–1518.

[13] C. Oldenburg, Screening and ranking framework for geologic CO2 storage site selection on the basisof health, safety, and environmental risk, Environmental Geology 54 (2008) 1687–1694.

22

[14] C. Oldenburg, S. Bryant, J.-P. Nicot, Certification framework based on effective trapping for geo-logic carbon sequestration, Environmental GeologyInternational Journal of Greenhouse Gas Control 4(2009) 444–457.

[15] P. R. Maul, R. Metcalfe, J. Pearce, D. Savage, J. M. West, Performance assessments for the geologi-cal storage of carbon dioxide: learning from the radioactive waste disposal experience, InternationalJournal of Greenhouse Gas Control 4 (1) (2007) 444–455.

[16] Q. Zhou, J. T. Birkholzer, C. F. Tsang, J. Rutqvist, A method for quick assessment of CO2 storagecapacity in closed and semi-closed saline formations, International Journal of Greenhouse Gas Control2 (2008) 626–639.

[17] N. Wiener, The Homogeneous Chaos, Amer. J. Math. 60 (1938) 897–936.

[18] R. H. Cameron, W. T. Martin, The orthogonal development of non-linear functionals in series offourier-hermite functionals, Ann. Math. 48 (1947) 385–392.

[19] D. Xiu, D. Lucor, C. Su, G. Karniadakis, Stochastic modeling of flow structure interactions usinggeneralized polynomial chaos, J. Fluids Engrg 124 (2002) 51–59.

[20] D. Xiu, G. Karniadakis, The Wiener-Askey Polynomial Chaos for stochastic differential equations,SIAM J. Sci. Comput. 24 (2002) 619–644.

[21] D. Xiu, G. Karniadakis, Modeling uncertainty in flow simulations via generalized Polynomial Chaos,J. Comput. Phys. 187 (2003) 137–167.

[22] O. Le Maıtre, O. Knio, H. Najm, R. Ghanem, Uncertainty propagation using Wiener-Haar expansions,J. Comput. Phyics 197 (1) (2004) 28–57.

[23] O. Le Maıtre, H. Najm, R. Ghanem, O. Knio, Multi-resolution analysis of Wiener-type uncertaintypropagation schemes, J. Comput. Phys. 197 (2) (2004) 502–531.

[24] O. Le Maıtre, L. Mathelin, O. Knio, M. Hussaini, Asynchronous time integration for polynomial chaosexpansion of uncertain periodic dynamics, Discrete and Continuous Dynamical Systems 28 (1) (2010)199–226.

[25] A. Alexanderian, O. L. Maıtre, H. Najm, M. Iskandarani, O. Knio, Multiscale stochastic precondition-ers in non-intrusive spectral projection, Journal of Scientific Computing 50 (2012) 306–340.

[26] A. Alexanderian, J. Winokur, I. Sraj, A. Srinivasan, M. Iskandarani, W. C. Thacker, O. M. Knio,Global sensitivity analysis in an ocean general circulation model: a sparse spectral projection approach,Computational Geosciences 16 (3) (2012) 757–778.

[27] J. Winokur, P. Conrad, I. Sraj, O. M. Knio, A. Srinivasan, W. C. Thacker, Y. Marzouk, M. Iskandarani,A priori testing of sparse adaptive polynomial chaos expansions using an ocean general circulationmodel database, Comput. Geosci. 17 (6) (2013) 899–911.

[28] L. Yan, L. Guo, D. Xiu, Stochastic collocation algorithms using `1-minimization, Int. J. Uncertain.Quantif. 2 (3) (2012) 279–293.

23

[29] J. Peng, J. Hampton, A. Doostan, A weighted `1-minimization approach for sparse polynomial chaosexpansions, Journal of Computational Physics 267 (2014) 92–111.

[30] D. Bigoni, A. P. Engsig-Karup, C. Eskilsson, Efficient uncertainty quantification of a fully nonlinearand dispersive water wave model with random inputs, Journal of Engineering Mathematics 101 (1)(2016) 87–113.

[31] I. Babuska, F. Nobile, R. Tempone, A stochastic collocation method for elliptic partial differentialequations with random input data, SIAM J. Numer. Anal. 45 (3) (2007) 1005–1034.

[32] R. Ghanem, P. Spanos, Stochastic Finite Elements: A Spectral Approach, Dover, 2002, 2nd edition.

[33] O. Le Maıtre, O. Knio, Spectral Methods for Uncertainty Quantification With Applications to Compu-tational Fluid Dynamics, Scientific Computation, Springer, 2010.

[34] I. Sobol’, Estimation of the sensitivity of nonlinear mathematical models, Matematicheskoe Mod-elirovanie 2 (1) (1990) 112–118.

[35] T. Homma, A. Saltelli, Importance measures in global sensitivity analysis of nonlinear models, Relia-bility Engineering & System Safety 52 (1) (1996) 1–17.

[36] I. Sobol’, Global sensitivity indices for nonlinear mathematical models and their monte carlo estimates,Mathematics and Computers in Simulation 55 (1-3) (2001) 271–280, the Second IMACS Seminar onMonte Carlo Methods.

[37] A. Saltelli, Sensitivity analysis for importance assessment, Risk Analysis 22 (3) (2002) 579–590.

[38] T. Crestaux, O. L. Maitre, J.-M. Martinez, Polynomial chaos expansion for sensitivity analysis, Relia-bility Engineering & System Safety 94 (7) (2009) 1161 – 1172, special Issue on Sensitivity Analysis.

[39] B. Sudret, Global sensitivity analysis using polynomial chaos expansions, Reliability Engineering &System Safety 93 (7) (2008) 964 – 979.

[40] M. Ashraf, S. Oladyshkin, W. Nowak, Geological storage of CO2: Application, feasibility and effi-ciency of global sensitivity analysis and risk assessment using the arbitrary polynomial chaos, Interna-tional Journal of Greenhouse Gas Control 19 (2013) 704–719.

[41] A. Namhata, S. Oladyshkin, R. M. Dilmore, L. Zhang, D. V. Nakles, Probabilistic assessment of abovezone pressure predictions at a geologic carbon storage site, Scientific Reports 6.

[42] Y. Zhang, Y. Liu, G. Pau, S. Oladyshkin, S. Finsterle, Evaluation of multiple reduced-order models toenhance confidence in global sensitivity analyses, International Journal of Greenhouse Gas Control 49(2016) 217–226.

[43] A. N. Brooks, A. T. Corey, Hydraulic properties of porous media, Hydrol. Pap. Fort Collins, ColoradoState University.

[44] R. Span, W. Wagner, A new equation of state for carbon dioxide covering the fluid region from thetriple-point temperature to 1100 K at pressures up to 800 MPa, J. Phys. Chem. Ref. Data 25 (6) (1996)1509–1596.

24

[45] Release on the IAPWS industrial formulation 1997 for the thermodynamic properties of water andsteam, http://www.iapws.org/relguide/IF97-Rev.html, accessed: 2017-08-07.

[46] A. Fenghour, W. Wakeham, V. Vesovic, The viscosity of carbon dioxide, J. Phys. Chem. Ref. Data27 (1) (1998) 31–44.

[47] M. Batzle, Z. Wang, Seismic properties of pore fluids, Geophysics 57 (1992) 1396–1408.

[48] N. Spycher, K. Pruess, CO2-H2O mixtures in the geological sequestration of CO2. ii. partitioning inchloride brines at 12-100c and up to 600 bar, Geochim. Cosmochim. Acta 69 (13) (2005) 3309–3320.

[49] B. Flemisch, M. Darcis, K. E. nad B. Faigle, A. Lauser, K. Mosthaf, S. Muthing, P. Nuske, A. Tatomir,M. Wolff, R. Helmig, DUMUX: DUNE for Multi-Phase, Component, Scale, Physics, flow andtransport in porous media, Advances in Water Resources 34 (9) (2011) 1102–1112.

[50] R. Huber, R. Helmig, Node-centered finite-volume discretization for the numerical simulation of mul-tiphase flow in heterogenous porous media, Comput. Geosci. (4) (2000) 141–164.

[51] G. Blatman, B. Sudret, Adaptive sparse polynomial chaos expansion based on least angle regression,Journal of Computational Physics 230 (6) (2011) 2345–2367.

[52] P. Conrad, Y. Marzouk, Adaptive Smolyak pseudospectral approximations, SIAM J. Sci. Comput.35 (6) (2013) A2643–A2670.

[53] C. Bryant, S. Prudhomme, T. Wildey, Error decomposition and adaptivity for response surface approx-imations from PDEs with parametric uncertainty, SIAM/ASA Journal on Uncertainty Quantification3 (1) (2015) 1020–1045.

[54] J. Hampton, A. Doostan, Compressive sampling of polynomial chaos expansions: convergence analy-sis and sampling strategies, J. Comput. Phys. 280 (2015) 363–386.

[55] D. Williams, Probability with martingales, Cambridge Mathematical Textbooks, Cambridge UniversityPress, Cambridge, 1991.

[56] A. Alexanderian, On spectral methods for variance based sensitivity analysis, Probability Surveys 10(2013) 51–68.

[57] M. Gerritsma, J.-B. Van der Steen, P. Vos, G. Karniadakis, Time-dependent generalized polynomialchaos, Journal of Computational Physics 229 (22) (2010) 8333–8363.

[58] A. Alexanderian, F. Rizzi, M. Rathinam, O. P. Le Maıtre, O. M. Knio, Preconditioned Bayesian re-gression for stochastic chemical kinetics., J. Sci. Comput. 58 (3) (2014) 592–626.

25