co 2 saturation estimates at sleipner (north sea) from ...barn/myarticles/yan2019.pdf ·...

17
Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner (North Sea) from seismic tomography and rock physics inversion Hong Yan 1,2 , Bastien Dupuy 1, Anouar Romdhane 1 and Børge Arntsen 2 1 SINTEF Industry, Trondheim, Norway, and 2 NTNU Trondheim, Norway Received October 2017, revision accepted September 2018 ABSTRACT CO 2 saturations are estimated at Sleipner using a two-step imaging workflow. The workflow combines seismic tomography (full-waveform inversion) and rock physics inversion and is applied to a two-dimensional seismic line located near the injection point at Sleipner. We use baseline data (1994 vintage, before CO 2 injection) and mon- itor data that was acquired after 12 years of CO 2 injection (2008 vintage). P-wave velocity models are generated using the Full waveform inversion technology and then, we invert selected rock physics parameters using an rock physics inversion methodol- ogy. Full waveform inversion provides high-resolution P-wave velocity models both for baseline and monitor data. The physical relations between rock physics proper- ties and acoustic wave velocities in the Utsira unconsolidated sandstone (reservoir formation) are defined using a dynamic rock physics model based on well-known Biot–Gassmann theories. For data prior to injection, rock frame properties (porosity, bulk and shear dry moduli) are estimated using rock physics inversion that allows deriving physically consistent properties with related uncertainty. We show that the uncertainty related to limited input data (only P-wave velocity) is not an issue because the mean values of parameters are correct. These rock frame properties are then used as a priori constraint in the monitor case. For monitor data, the Full waveform in- version results show nicely resolved thin layers of CO 2 –brine saturated sandstones under intra-reservoir shale layers. The CO 2 saturation estimation is carried out by plugging an effective fluid phase in the rock physics model. Calculating the effective fluid bulk modulus of the brine–CO 2 mixture (using Brie equation in our study) is shown to be the key factor to link P-wave velocity to CO 2 saturation. The inversion tests are done with several values of Brie/patchiness exponent and show that the CO 2 saturation estimates are varying between 0.30 and 0.90 depending on the rock physics model and the location in the reservoir. The uncertainty in CO 2 saturation estimation is usually lower than 0.20. When the patchiness exponent is considered as unknown, the inversion is less constrained and we end up with values of exponent varying between 5 and 20 and up to 33 in specific reservoir areas. These estimations tend to show that the CO 2 –brine mixing is between uniform and patchy mixing and variable throughout the reservoir. Key words: CO 2 storage monitoring, Full Waveform Inversion, Rock physics inver- sion, Quantitative imaging, Saturation estimation. INTRODUCTION The global warming is gaining lots of attention in the last decade with the increasing energy demand and fossil energy E-mail: [email protected] consumption. Excessive CO 2 emissions that have a substantial contribution to the global warming (Metz et al. 2005) require the implementation of immediate mitigation measures. The CO 2 Capture and Storage (CCS) technology is hence proposed as a mean to reduce the CO 2 emission into the atmosphere. 1 C 2018 European Association of Geoscientists & Engineers

Upload: others

Post on 18-Jul-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693

CO2 saturation estimates at Sleipner (North Sea) from seismictomography and rock physics inversion

Hong Yan1,2, Bastien Dupuy1∗, Anouar Romdhane1 and Børge Arntsen2

1SINTEF Industry, Trondheim, Norway, and 2NTNU Trondheim, Norway

Received October 2017, revision accepted September 2018

ABSTRACTCO2 saturations are estimated at Sleipner using a two-step imaging workflow. Theworkflow combines seismic tomography (full-waveform inversion) and rock physicsinversion and is applied to a two-dimensional seismic line located near the injectionpoint at Sleipner. We use baseline data (1994 vintage, before CO2 injection) and mon-itor data that was acquired after 12 years of CO2 injection (2008 vintage). P-wavevelocity models are generated using the Full waveform inversion technology and then,we invert selected rock physics parameters using an rock physics inversion methodol-ogy. Full waveform inversion provides high-resolution P-wave velocity models bothfor baseline and monitor data. The physical relations between rock physics proper-ties and acoustic wave velocities in the Utsira unconsolidated sandstone (reservoirformation) are defined using a dynamic rock physics model based on well-knownBiot–Gassmann theories. For data prior to injection, rock frame properties (porosity,bulk and shear dry moduli) are estimated using rock physics inversion that allowsderiving physically consistent properties with related uncertainty. We show that theuncertainty related to limited input data (only P-wave velocity) is not an issue becausethe mean values of parameters are correct. These rock frame properties are then usedas a priori constraint in the monitor case. For monitor data, the Full waveform in-version results show nicely resolved thin layers of CO2–brine saturated sandstonesunder intra-reservoir shale layers. The CO2 saturation estimation is carried out byplugging an effective fluid phase in the rock physics model. Calculating the effectivefluid bulk modulus of the brine–CO2 mixture (using Brie equation in our study) isshown to be the key factor to link P-wave velocity to CO2 saturation. The inversiontests are done with several values of Brie/patchiness exponent and show that theCO2 saturation estimates are varying between 0.30 and 0.90 depending on the rockphysics model and the location in the reservoir. The uncertainty in CO2 saturationestimation is usually lower than 0.20. When the patchiness exponent is considered asunknown, the inversion is less constrained and we end up with values of exponentvarying between 5 and 20 and up to 33 in specific reservoir areas. These estimationstend to show that the CO2–brine mixing is between uniform and patchy mixing andvariable throughout the reservoir.

Key words: CO2 storage monitoring, Full Waveform Inversion, Rock physics inver-sion, Quantitative imaging, Saturation estimation.

INTRODUCTION

The global warming is gaining lots of attention in the lastdecade with the increasing energy demand and fossil energy

∗E-mail: [email protected]

consumption. Excessive CO2 emissions that have a substantialcontribution to the global warming (Metz et al. 2005) requirethe implementation of immediate mitigation measures. TheCO2 Capture and Storage (CCS) technology is hence proposedas a mean to reduce the CO2 emission into the atmosphere.

1C© 2018 European Association of Geoscientists & Engineers

Page 2: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

2 H. Yan et al.

Large CO2 storage capacity has already been proven in deepsaline aquifers (Halland et al. 2011; Bohm et al. 2015). TheSleipner CO2 storage site is the first commercial-scale offshorestorage operation and the longest in operation in the world.Geophysical monitoring (including seismic and non-seismictechniques) is used to monitor the subsurface changes relatedto the CO2 plume migration, in order to verify conformanceand early detection of potential leakage (containment moni-toring).

The seismic monitoring programme at Sleipner includesa three-dimensional baseline survey prior to CO2 injectionand several monitor surveys every 2–3 years (Arts et al. 2008;Furre, Kiær and Eiken 2015; Furre et al. 2017). Previous stud-ies based on time-lapse analysis help to reveal the details ofthe CO2 plume development ( Eiken et al. 2000; Arts et al.

2002; Arts et al. 2004). Structural mapping and frequency de-composition have been used to estimate the upper most layerthickness (Williams and Chadwick 2012; White 2013). Artset al. (2002) give quantitative estimates of CO2 mass usingamplitude information. More recently, Furre et al. (2015) per-form a detailed analysis of the thin CO2 layers, with respectto time shift development and amplitude changes.

The proper understanding of saturation developmentin Utsira sandstone has been widely studied by several au-thors. Ghaderi and Landrø (2009) use seismic modelling ofthin layer responses and propose a simultaneous inversionof velocity and layer thicknesses of partially saturated lay-ers with CO2. Post-stack impedance analysis (Ghosh, Senand Vedanti 2015) and time-lapse Amplitude Versus Off-set (AVO) analysis (Chadwick et al. 2010) have been ap-plied to estimate CO2 saturations. Bergmann and Chadwick(2015) propose a method using time-shift analysis to esti-mate the influence of fluid mixing law both with syntheticand real data. As an alternative, quantitative maps of veloc-ity properties derived at Sleipner using full waveform inver-sion (FWI) have been used as input to estimate saturationdistributions. Queiber and Singh (2013) and Romdhane andQuerendez (2014) apply FWI in order to derive highly re-solved P-wave velocity models. Queiber and Singh (2013)use simple rock physics relations (Gassmann (1951) modelcombined with Brie’s law (Brie et al. 1995)) to describe theCO2 and brine mixture from the P-wave velocity. Dupuy,Garambois and Virieux (2016) propose a two-step workflowwith dynamic rock physics models to tackle the existing lim-itation in frequency sensitivity and obtain reliable reservoirproperties from high-resolution velocity models. Sensitivitystudies have highlighted that extra input data in the inver-sion process, such as shear wave velocity and P-wave quality

factors are important for the estimation of rock frame mod-uli and can be helpful to better constrain in the estimation ofCO2 saturation (Bohm et al. 2015; Dupuy et al. 2016; Dupuyet al. 2017). At Sleipner, Dupuy et al. (2017) show howeverthat the estimation of CO2 saturations from P-wave velocitymodels derived from FWI is still possible.

In this work, FWI is combined with the rock physics in-version (RPI) method to invert for selected poroelastic prop-erties (rock frame moduli, porosity, CO2 saturation and Briepatchiness exponent) and to derive high-resolution maps ofCO2 saturation with uncertainty. Data from 1994 and 2008vintages are used. The rock frame properties are recovered us-ing baseline data and used as a priori parameters to estimateCO2 saturation distribution with the monitor dataset. We in-vert the CO2 saturation and other properties for the monitorcase and we constrain the inversion using a priori geologi-cal information (fluid properties, grain properties, permeabil-ity and cementation factor) and a priori information derivedfrom the baseline results (bulk modulus, shear modulus andporosity). We also analyse how the Brie exponent, describ-ing the fluids distribution (from uniform to patchy mixing), isaffecting the results.

This paper is divided into four sections. While the firstpart gives the methodology workflow, the second section pro-vides a short description of the geological context at Sleip-ner justifying the choice of some a priori parameters. Testspresented in section 3 provides expected sensitivities for es-timation of rock frame properties and CO2 saturation anddistribution. The case study at Sleipner, for both baseline andmonitor data are described in section 4.

METHODOLOGY

We use a two-step workflow combining full waveform inver-sion (FWI) and rock physics inversion (RPI) methods. FromFWI, we generate a P-wave velocity model that is used as inputof the RPI step.

FWI is performed in the frequency-space domain (Pratt,Shin and Hick 1998). The two-dimensional acoustic waveequation is solved using finite differences with the mixed gridapproach (Hustedt, Operto and Virieux 2004). The inverseproblem is solved using a pre-conditioned gradient algorithm.The gradient method in its standard form provides a rela-tionship between the perturbation model and the data resid-uals. An approximation of the Hessian operator with onlythe diagonal elements is used to save computation cost. Ba-sic data pre-processing steps (including quality control of thedata, offset selection, muting and coordinate transformation)

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 3: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

Sleipner (North Sea) from seismic tomography 3

are performed before running the inversion (Romdhane andQuerendez 2014). We derive an initial P-wave velocity modelusing root mean square velocity in time converted into inter-val velocity in depth. The misfit function and the gradient ofthe misfit at iteration n can be expressed:

ξ (m) = 12

∥∥d mod − dobs

∥∥2,

∂ξ (mn)∂m = �e

{JTδd∗} ,

(1)

where ∂ξ (mn)∂m is the gradient of the misfit at iteration n. dmod

and dobs are the modelled and observed data, respectively. �e

is the real part of a complex number. J denotes the Jacobianmatrix, thus JT is the transpose of the Jacobian matrix. δd∗

denotes the conjugate of the data residuals. The perturbationmodel is given by:

δmn = mn+1 − mn. (2)

The gradient method can be formulated as the linkbetween the perturbation model and data residuals, and isgiven as:

δmn = −αn∂ξ (mn)

∂m, (3)

where αn is the step length at iteration n. Proper scaling andregularization can be applied to ensure the computational sta-bility. The approximate Hessian Ha can be written:

Ha = �e{

JTJ∗} . (4)

If we only consider the diagonal term of the approximateHessian, the model update at iteration n becomes:

δmn = −αn(diagHan+ λI)−1 ∂ξ (mn)

∂m, (5)

where λ denotes a damping parameter used to ensure compu-tational stability.

The second step consists in solving the inverse prob-lem linking seismic properties and rock physics properties.The link between rock physics parameters and seismic at-tributes constitutes the forward problem, usually called therock physics model. The forward problem can be described ina generic manner by Biot–Gassmann theory (Gassmann 1951;Biot 1956a) as long as rock frame properties are somehow de-fined or estimated.

The Biot–Gassmann equations are valid under several as-sumptions. The Gassmann equations are valid within a lowfrequency limit and do not account for the chemical interac-tion between the grains and the fluid phases. The Biot theoryis limited to homogeneous and isotropic media, which consistsof only one type of grain. To be more realistic, Pride (2005)

extended the Biot–Gassmann equations and included a gen-eralized dynamic permeability (Johnson, Koplik and Dashen1987). The extended Biot–Gassmann theory can be applied toa multi-composite material saturated with a multi-fluid mix-ture using averaging techniques. In addition, this approach isvalid within a wide frequency range. More details on the as-sumptions and theory for this rock physics model are given inPride (2005). The extensive Biot–Gassmann equations allowus to calculate the viscoelastic attributes for an effective fluidphase even if only P-wave velocity is used.

Biot (1956b) demonstrates that there are three types ofwaves propagating in a fluid saturated porous medium. Theseare compressional wave (P-wave), shear wave (S-wave) andan additional slow compressional wave (Biot wave or slowP-wave). The slowness of P-wave is formulated by Pride(2005) as:

S2P (ω) = γ (ω)

2− 1

2

√γ 2(ω) − 4

(ρρ(ω) − ρ2

f

)HM − C2

. (6)

Thus, we can deduce the effective P-wave velocity as:

Vp(ω) = 1Re

(Sp(ω)

) . (7)

In equation (6), ρ and ρf and ρ(ω) are the bulk den-sity, the effective fluid density and the flow resistance term,respectively. Note that ω is the complex pulsation and is in-volved in frequency-dependent terms such as ρ(ω) and γ (ω).The undrained bulk modulus KU, the Biot modulus C, theshear modulus G and the fluid storage coefficient M are thefour mechanical moduli. The additional terms γ (ω) and H aredefined by:

γ (ω) = ρM + ρ(ω)H − 2ρ f C

HM − C2,

H = KU + 43

G. (8)

The four mechanical moduli are related to the homoge-nized porous solid. KU, C and M are formulated as:

KU = φKd + (1 − (1 + φ)Kd/Ks)K f

φ(1 + �),

C = (1 − Kd/Ks)K f

φ(1 + �),

M = K f

φ(1 + �), (9)

where the term � is given by:

� = 1 − φ

φ

K f

Ks

(1 − Kd

(1 − φ)Ks

), (10)

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 4: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

4 H. Yan et al.

where porosity φ is defined as the ratio between void andtotal volume in the solid matrix. Kd, Gd, Ks and Kf are re-spectively the frame bulk modulus, the frame shear modulus,the grains bulk modulus and the effective fluid bulk modulus.The undrained shear modulus G is assumed to be equal to dryshear modulus Gd.

We consider that the solid material for both shale andsand layers consists of a mix of clay and quartz grains. Wecompute the grains bulk moduli Ks by averaging the bounds(Hashin and Shtrikman 1963). The grain density ρs is com-puted using an arithmetic average:

ρ = (1 − φ)ρs + φρ f . (11)

The extended Biot–Gassmann formulation given by Pride(2005) is valid over a wide range of frequency because it in-volves Johnson et al. (1987) model which gives a frequency-dependent dynamic permeability k(ω). It allows for correctingthe seismic permeability and differentiates the viscous and in-ertial predominant effects at low and high frequency domain(Dupuy et al. 2016). The dynamic permeability k(ω) is ex-pressed by Johnson et al. (1987):

k(ω) = k0√1 − 1

2 i ω

ωc− i ω

ωc

, (12)

where ω is the angular frequency. ωc is the characteristic an-gular frequency that characterizes the maximum attenuationpeak. k0 is the hydraulic permeability and i is the complexnumber. The characteristic frequency ωc can be formulatedwith the cementation factor m, the effective fluid viscosity η

and the effective fluid density ρf (Adler, Jacquin and Thovert1992) such as:

ωc = η

ρ f k0φ−m

. (13)

Then, we can define frequency-dependent flow resistanceterm ρ(ω) that is involved in the slowness equation (6) and isdefined by:

ρ(ω) = iηωk(ω)

. (14)

For the monitor step, it is assumed that the fluid phasein the pore space can be treated as a single-phase mixtureof brine and CO2, with effective properties computed usingBrie mixing law (Brie et al. 1995) for fluid bulk modulus,Teja and Rice equation (Teja and Rice 1981) for effective vis-cosity and arithmetic average (Voigt 1928) for fluid density.The sensitivity tests provided in Dupuy et al. (2016) showthat the fluid bulk modulus is the main factor affecting theP-wave velocity change. The Brie equation involves an empir-ical exponent e:

K f = (KCO2 − KW)SeW + KCO2, (15)

where SW denotes the brine saturation, KCO2 and KW denotethe CO2 and brine bulk moduli, respectively. The exponente can vary between 1 and 40 (Brie et al. 1995), which cor-responds to various mixture trends (Fig. 1). For instance, thelower bound (e = 40) stands for the so-called uniform mixing(the two fluids are mixed at the finest scale), while the up-per bound (e = 1) stands for the so-called patchy mixing. Itis worth noting that the patchy mixing is related to empiricalobservations and its definition is variable depending on the au-thors. The exponent e is proposed to be equal to 3 (Brie et al.

Figure 1 P-wave velocity variation with brine saturation for various patchiness exponents.

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 5: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

Sleipner (North Sea) from seismic tomography 5

1995) and 5 (Carcione and Picotti 2006) to match the caserelated to gas and brine mixture. The Brie exponent e is some-times called patchiness exponent (Papageorgiou, Amalokwuand Chapman 2016). The set of rock physics model equa-tions (6)–(14) constitute the forward model that allows us tocompute visco-elastic attributes (mainly P-wave velocity) fromrock physics properties (solid and fluid phases and rock frameproperties).

To solve the RPI problem, we use a semi-global op-timization method (Neighbourhood Algorithm (NA) (Sam-bridge 1999)) to minimize the discrepancy between observeddata and calculated data obtained by forward modelling. Themethod searches for the global minimum over the wholemodel space in a partially random way and the search processis guided by each iteration towards areas with a lower misfit.The scalar misfit function C(m) between the observed datadobs and the estimated data g(m) from the forward modellingcan be given as an L2 norm:

C(m) =[(dobs − g(m))T(dobs − g(m))

]. (16)

m is the model vector describing the model space. TheNA algorithm has only two control parameters: the numberof new models at each iteration and the resampling size ofVoronoı cells. The misfit of the previous iteration decides thenew sampling of Voronoı cell for the next iteration. Sambridge(1999) and Dupuy et al. (2016) provide more details on theNA algorithm and the RPI, respectively.

The RPI process give estimates of selected parameterswith the related uncertainty. The sensitivity study by Dupuyet al. (2017) shows that the estimation of CO2 saturation fromP-wave velocity input only is not well constrained when theCO2 saturation is high (it is directly related to the shape ofthe P-wave velocity versus saturation curve, see Fig. 1). Exam-ples of the misfit function shapes for the estimation of frameparameters and saturation are given in sensitivity tests. By as-suming a Gaussian distribution of estimated models, we canestimate the mean value and associated standard deviationsto assess the uncertainty.

GEOLOGICAL BA C K GR OUN D

Since 1996, about 0.9 million tons of CO2 is annually injectedinto the Utsira formation at Sleipner at a depth of 1012 m be-low the sea level through a deviated well (Arts et al. 2008).The thickness of the Utsira sandstone is about 300 m and issurrounded by the Nordland shale on the top and the Horda-land shale on the bottom. The caprock is assumed to be a goodquality seal with a low probability of leakage (Chadwick et al.

2004). The Utsira formation consists of late Miocene to earlyPliocene dominantly sandy unit and some shaly intra-reservoirhorizons (Chadwick et al. 2000). It is a highly porous, un-consolidated, permeable sandstone that shows good reservoirproperties for high injection rate through only one single well(Michael et al. 2010). The intra-reservoir shale layers are verythin and can be less than 1 m thick (Zweigel 2000).

To constrain our rock physics inversion (RPI), weneed to determine a set of realistic a priori rock physicsproperties. These parameters are defined based on the exist-ing geological studies on Sleipner field. Grain moduli Ks of theUtsira sandstone and the Nordland shale are calculated by theHashin–Shtrikman bounds method (Hashin and Shtrikman1963; Mavko, Mukerji and Dvorkin 2009). Grain density ρs

for the Utsira sandstone and the Nordland shale are computedusing volume weighted averages. The cementation parameterm (Pride 2005) is assumed to be equal to 1 for the uncon-solidated sandstone and mudstone. As discussed in sensitivitytests by Dupuy et al. (2016), the cementation factor has a neg-ligible influence on P-wave velocity. The permeability for theUtsira formation is high and ranges from 1 × 10−12 m2 to 3 ×10−12 m2 (Boait et al. 2012). We use an average permeabilityof 2 × 10−12 m2 for Utsira formation. The permeability ofNordland shale is defined from laboratory data and assumedto be equal to 1.48 × 10−17 m2. The grains and rock frameproperties of Utsira sandstone and Nordland shale are givenin Table 1.

We consider a CO2 and brine mixture after CO2 injectionand a fully brine saturated medium before CO2 injection forthe saturating fluids properties. The bulk modulus of brine isequal to 2.3 GPa at reservoir conditions (Boait et al. 2012),and the density of brine is assumed to be equal to 1030 kg/m3

(Mavko et al. 2009) at reservoir temperature and pressure.Most of the CO2 is stored in a supercritical state (Arts et al.

2008) and we use a relatively high density of 700 kg/m3 for theCO2 phase (Lindeberg 2013). Since the initial pore pressure of8 MPa is considered (Furre and Eiken 2014) and the reservoirtemperature is ranging between 27°C (top Utsira formation)and 37°C (near injection point), the range of KCO2 is estimated

Table 1 Utsira sandstone and Nordland shale properties

Grain Properties Rock Frame Properties

Ks ρs m ko

Lithology GPa kg/m3 m2

Utsira Sandstone 39.29 2663.5 1 2 × 10−12

Nordland Shale 22.6 2390 1 1.48 × 10−17

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 6: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

6 H. Yan et al.

Table 2 Brine and CO2 properties

Fluid Properties

ƞ ρf Kf

Fluid phase Pa.s kg/m3 GPa

Brine 6.9 × 10−4 1030 2.3CO2 6 × 10−5 700 0.075

to be equal to 0.02 GPa to 0.075 GPa (Ghaderi and Landrø2009). We use a value of 0.075 GPa for bulk modulus of CO2

(Lindeberg 2013) regarding in situ pressure and temperatureconditions. The viscosity of CO2 is discussed in many papersand depends on the temperature and pressure (Singh et al.2010; Gasda et al. 2012). However, the effective viscosity in-fluence on seismic velocity is minor at low frequency (Dupuyand Stovas 2016). We use 6 × 10−5 Pa.s for the CO2 viscosity.The fluids properties used in the case study are summarizedin Table 2. As CO2 is under a supercritical state and reser-voir pressure and temperature conditions are known, we havechosen to fix the CO2 properties and not estimate them. Oth-erwise, sensitivity tests have been carried out to estimate CO2

properties in addition to saturation (Dupuy et al. 2017). Thetests show that the CO2 saturation estimates are only slightlyaffected if CO2 properties are unknown.

S E N S I T I V I T Y T E S T S

Sensitivity tests for baseline parameters, i.e. porosity and rockframe moduli before CO2 injection are carried out. The for-ward model is based on the extended Biot–Gassmann theorywith fully brine saturated Utsira sandstone, and is describedin the methodology part. The grain and frame properties forUtsira sandstone and brine properties are given in Tables 1and 2.

Figure 2 (bottom panel) shows a point test where we esti-mate frame bulk modulus, frame shear modulus and porosityusing a different combination of input data (P-wave, S-wavevelocity and density). Examining the size and the shape ofthe low misfit area in these sensitivity plots allows assessingthe uncertainty of each parameter estimate and how well theseestimates are constrained depending on which input data areused. The computation of S-wave velocity as well as extensivesensitivity tests are described in details in Dupuy et al. (2016).In Fig. 2, the low misfit area is wider when only P-wave veloc-ity is used than when P-wave, S-wave velocities and density areinput data. The dry shear modulus is a lot better constrainedwhen using VP, VS and ρ inputs, but porosity and dry bulk

modulus have similar uncertainty. It is also worth noting thatmean values of all low misfit models (pink models) are quiteclose to the true value, even when only P-wave velocity isused. This is a promising conclusion for the real data applica-tion where acoustic full waveform inversion is used to deriveP-wave velocity maps.

Similar sensitivity tests are given for CO2 saturation andBrie exponent estimates in Fig. 3. We observe that the CO2

saturation is correctly estimated when the saturation is low,while it is more difficult to get an accurate estimate when CO2

saturation is high. This effect is directly related to the shape ofthe VP-SCO2 curve (Fig. 1) where we have a quick drop of VP

as soon as little CO2 is present in the porous medium (fizz-gaseffect). However, for Brie exponent larger than 5 (fluid mix-ture is becoming more and more homogeneous), a change ofCO2 saturation is not affecting P-wave velocity when this sat-uration is greater than 30–50%. The inversion is consequentlymore difficult to constrain for high CO2 saturations and therelated uncertainty is larger. It is worth noting that this sen-sitivity test is carried out using VP input only. Supplementarytests show that using additional input help to mitigate thisissue of non-linearity (Dupuy et al. 2017). This first test isassuming that the fluid mixing law is known (Brie exponentequal to 5). If this patchiness exponent is inverted togetherwith CO2 saturation (bottom panel of Fig. 3), we observe atrade-off between these two parameters that is expected asonly P-wave velocity is used as input. The uncertainty of thesaturation estimates is consequently larger.

SLE IPNER T EST CASE

The two-step workflow is carried out for data before (1994vintage, baseline) and after CO2 injection (2008 vintage, mon-itor). For both vintages, seismic data from inline 1836 (north-south orientation, close to injection point) is extracted. Theapplied processing workflow aims at improving the signal tonoise ratio without affecting the amplitudes of the data. Itincludes the removal of the t2 divergence compensation, datamuting before the first arrivals and the selection of data withan offset higher than 420 m. The available data include off-sets up to 1800 m. For the initial velocity model, we assume ahorizontal sea bottom and use a stacking velocity model con-verted to interval velocity in depth and smoothed. The modelis 7 km long and has a depth of 1.35 km. It is defined on aregular 3 m squared grid to ensure sufficient accuracy whenhigh frequencies are considered. To mitigate the non-linearityof full waveform inversion, we perform the inversion fromlow to high frequencies using frequency components up to

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 7: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

Sleipner (North Sea) from seismic tomography 7

0 0.01Misfit

8

6

4

2

2 4 6 8 2 4 6 8 2 4 6 8

0.4

0.3

0.2

0.1

0.4

0.3

0.2

0.1

Dry bulk modulus (GPa) Dry shear modulus (GPa) Dry bulk modulus (GPa)

Por

osity

Por

osity

Dry

she

ar m

odul

us (G

Pa)

0 0.01Misfit

8

6

4

2

2 4 6 8 2 4 6 8 2 4 6 8

0.4

0.3

0.2

0.1

0.4

0.3

0.2

0.1

Dry bulk modulus (GPa) Dry shear modulus (GPa) Dry bulk modulus (GPa)

Por

osity

Por

osity

Dry

she

ar m

odul

us (G

Pa)

Figure 2 Estimation of dry shear modulus, dry bulk modulus and porosity from P- wave velocity (top panels) and from P-wave and S-wavevelocities and density (bottom panel). The red cross indicates the true model. The colour of each dot represents the absolute misfit value foreach computed model (between 0 and 0.01).

33 Hz to derive a highly resolved P-wave velocity model. Atarget region around the Utsira sandstone is then used as inputto perform rock physics inversion.

FWI results

Figure 4(a,b) shows the P-wave velocity maps derived byfull waveform inversion (FWI) for both baseline and moni-tor cases, respectively. Selected interpreted horizons (corre-sponding to the seabed, the top Pliocene, the intra Pliocene,the top of Utsira sand wedge, the top of Utsira and the baseof Utsira) are also plotted for reference. The top of Utsira

sand wedge is an eastward thickening package, which is sepa-rated from the main sand part by a shale layer (Zweigel et al.

2004). As observed in Romdhane and Querendez (2014) us-ing a different inline, the models derived from FWI show aclear improvement in term of resolution. Figure 4(b) showsclear indications about the lateral extent and geometry ofthe CO2 plume. Figure 4(c) shows the corresponding time-migrated section converted to depth using a simple time todepth relationship based on well data situated at around 800m distance from the two-dimensional line used. The push-down of the reflection beneath the CO2 plume is clearlyvisible.

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 8: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

8 H. Yan et al.

2= 0.80

2= 0.15

2= 0.80, = 5 2

= 0.15, = 5

10010Number of models

0.1

0.3

0.7

0.9

0 0.01Misfit

0

Brin

e sa

tura

tion

0.5

10010

0.1

0.3

0.7

0.9

0 0.01Misfit

0Number of models

Brin

e sa

tura

tion

0.5

5

10

15

20

25

30

35

0 0.01Misfit

5

10

15

20

25

30

35

0 0.01Misfit

0.6 0.8 10.40.2Brine saturation

Brie

exp

onen

t

Figure 3 Estimation of brine saturation SW (top panels) and estimation of brine saturation and patchiness exponent (bottom panels) fromP-wave velocity (Figures from Dupuy et al. 2017). The blue cross indicates the model with the lowest misfit and the red cross indicates the truemodel. Below each panel are indicated the true value for CO2 saturation and Brie exponent. The colour of each dot represents the absolutemisfit value for each computed model (between 0 and 0.01; all models with misfit higher than 0.01 are clipped).

Close-up of the target region at the depth of Utsira reser-voir where the CO2 is expected to accumulate are shown inFig. 4(d,e) with interpreted horizons corresponding to the topof Utsira sand wedge, the top of Utsira and to the base ofUtsira from top to bottom. For the baseline case, two thicklow-velocity layers can be observed. These layers are probablyshowing high porosity unconsolidated sandstones where theCO2 is expected to migrate after injection. For the monitorcase, we observe thin low velocity layers that can be associ-ated with the accumulation of CO2. The velocity pushdowneffect at the base of the Utsira can also be observed. The betterresolution of the shale-sand interlayers is probably related tothe higher contrast between the slow-velocity CO2 plume and

the surrounded high-velocity brine-saturated reservoir, shaleoverburden and interbedded layers. It is worth noting that thevelocities of the overburden layers (just above Utsira reservoir)are slightly lower in the monitor case, while no effect relatedto CO2 anomaly should be observed. In addition, velocity dif-ferences observed below the reservoir are not meaningful ofany physical change because CO2 migration is driven by buoy-ancy effect. These uncertainties in velocity can be partly due tosome smearing effects (Queiber and Singh 2013; Romdhane,Querendez and Ravaut 2014) that can lead to overestimationof velocities in the CO2 plume (i.e. underestimation of veloc-ity change) observed in Utsira sandstone, and to the limitedavailable offset range (maximum of 1800 m).

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 9: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

Sleipner (North Sea) from seismic tomography 9

0

400

800

1200

Dep

th (

m)

00060054000300510Distance (m)

1500

1750

2000

2250

2500

P-w

ave

velo

city

(m

/s)

0

400

800

1200

Dep

th (

m)

00060054000300510Distance (m)

1500

1750

2000

2250

2500

P-w

ave

velo

city

(m

/s)

0

400

800

1200

Dep

th (

m)

1500 3000 4500 6000Distance (m)

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

1650

1900

2150

2400

P-w

ave

velo

city

(m

/s)

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

1650

1900

2150

2400

P-w

ave

velo

city

(m

/s)

(a)

(b)

(c)

(d)

(e)

Figure 4 (a) P-wave velocity model of baseline vintage (1994) derived by FWI, (b) P-wave velocity model of monitor vintage (2008) derived byFWI. (c) Time migrated section converted to depth of monitor vintage (2008). (d, e) Close-ups around the reservoir where CO2 is injected for(d) baseline and (e) monitor models. The injection point is located at 1012 m deep. The FWI has been carried out with limited offset (1800 m)with frequencies inverted up to 33 Hz. The black lines (and blue lines in (c)) correspond to selected interpreted horizons. For the full sections(a, b, c), the black lines are indicating the seabed, the top Pliocene, the intra Pliocene, the top of Utsira sand wedge, the top of Utsira and thebase of Utsira from top to bottom. For the close-up sections (d, e), the black lines are indicating the top of Utsira sand wedge, the top of Utsiraand the base of Utsira from top to bottom. The base of the Utsira taking into account the pushdown effect is added to (c).

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 10: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

10 H. Yan et al.

Baseline dataset: Estimation of rock frame properties

We apply the rock physics inversion (RPI) on baseline dataconsidering that the porous medium is saturated with brine.Prior geological information is helpful to constrain the es-timation of rock frame properties (Dupuy et al. 2016).The RPI is applied to the full waveform inversion de-rived P-wave velocity model to estimate rock frame mod-uli (Kd, Gd) and porosity φ for the reservoir where theCO2 plume is expected to accumulate. The inversion rangesare given as: 0.1 GPa < Kd < 10 GPa, 0.01 GPa <

Gd < 10 GPa, 0 < φ < 0.45. The inversion frequency is30 Hz.

The estimated bulk frame modulus, shear frame modulusand porosity and associated uncertainties are given in Figs 5

and 6. The bulk and shear frame moduli within the two low-velocity sand layers vary from 1 to 3 GPa with ±1 to 1.5 GPauncertainty and from 1 to 1.5 GPa with ±0.9 to 1.2 GPa un-certainties, respectively. The surrounding layers (interpretedas being thin shale interbedded layers and the overburden lay-ers) show higher values for both moduli (3 to 6 GPa with ±1.5to 2.5 GPa for the bulk modulus and 2 to 2.5 GPa with ±1.15to 1.5 GPa for the shear modulus). The porosity varies from0.30 to 0.35 with ±0.05 to 0.075 uncertainty in the sands,while it shows lower values in the surrounding layers (0.20 to0.25 with ±0.075 to 0.10 uncertainty). It is worth noting thatwe can observe the imprint of the a priori parameters (graindensity, grain bulk moduli and permeability) on the porositymap (Fig. 5). It is also noticed that the high porosity layers areslightly shifted compared to the interpreted layers. This can be

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

123456

x109 Dry

bul

k m

odul

us (

Pa)

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

0.8

1.3

1.8

2.3

x109

Dry

she

ar m

odul

us (

Pa)

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

0.15

0.2

0.25

0.3

Por

osity

Figure 5 From top to bottom: frame bulk modulus Kd, frame shear modulus Gd and porosity φ. The black lines are figuring the interpretedhorizons, i.e. the top of Utsira sand wedge, the top of Utsira and the base of Utsira from top to bottom.

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 11: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

Sleipner (North Sea) from seismic tomography 11

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

123456

x109

Dry

bul

k m

odul

us u

ncer

tain

ty (

Pa)

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

0.8

1.3

1.8

2.3

x109

Dry

she

ar m

odul

us u

ncer

tain

ty (

Pa)

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

0.15

0.2

0.25

0.3

Por

osity

unc

erta

inty

Figure 6 From top to bottom: inversion uncertainty of frame bulk modulus Kd, frame shear modulus Gd and porosity φ. The black lines arefiguring the interpreted horizons, i.e. the top of Utsira sand wedge, the top of Utsira and the base of Utsira from top to bottom.

caused by a partially erroneous time to depth conversion ofinterpreted horizons and related to the smearing effect ob-served in full waveform inversion (FWI) results. However,knowing that expected resolution of FWI is around half awavelength that corresponds to approximately 30 m for afrequency of 33 Hz and a velocity of 2000 m/s, this depthmismatch of 10 m maximum is in the range of uncertainty.That is why we chose to not do any stretch matching of thesehorizons.

Monitor dataset: Estimation of CO2 saturation andpatchiness exponent

After the estimation of rock frame properties, we focus onthe 2008 vintage dataset for the target region. Figure 7 showsthe maps of CO2 saturation using Brie rock physics model

with two exponent values (e = 1 and e = 5), and when e

is inverted together with saturation. Figure 8 shows the as-sociated uncertainties of CO2 saturation for the three cases.Figure 9 shows the estimation of patchiness exponent and itsassociated uncertainty for the third case (joint inversion ofCO2 saturation and Brie exponent). The estimation of CO2

saturation using e = 1 (so-called patchy mixing) shows clearCO2–brine-saturated regions, mainly around 880–970 m deepwith saturations reaching 0.90 ± 0.15–0.17 uncertainty. Theestimations of CO2 saturation within deeper sandstone layersare lower than 0.50 with ± 0.12 to 0.17 uncertainty.

The estimation of CO2 saturation using exponent e =5 also shows one high CO2–brine-saturated region (870–970m deep). CO2 saturation in this layer varies from 0.30 to0.70 with ±0.15 to 0.17 uncertainty. Especially, the high-est uncertainty tends to be associated with the highest CO2

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 12: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

12 H. Yan et al.

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

0

0.2

0.4

0.6

0.8

CO

2 sa

tura

tion

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

0

0.2

0.4

0.6

0.8

CO

2 sa

tura

tion

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

0

0.2

0.4

0.6

0.8

CO

2 sa

tura

tion

Figure 7 From top to bottom: CO2 saturation estimated by RPI using patchiness exponent e = 1, e = 5 and exponent inverted (given in Fig. 9).

saturated regions. For example, at 910–940 m deep, the CO2

saturation is going up to 0.70 with ±0.17 uncertainty. Wecan also see this distinct high CO2 saturated region at thedepth of approximately 900 m when we invert for the Brie(or patchiness) exponent and CO2 saturation simultaneously.In this case, the CO2 saturation is reaching 0.50 in this mainlayer with an uncertainty around 0.22–0.25. The estimationsof CO2 saturations within the deeper sandstone layers varyin the range of 0.20–0.30 with ±0.13 to 0.17 uncertainty.It is worth noting that uncertainties related to CO2 satura-tion estimates are higher in this third case, up to 0.25 in-stead of 0.17 in the two first cases. The patchiness exponent e

within these CO2–brine-saturated layers is ranging from 4 to33 with an uncertainty around 5–11. Again, the higher uncer-tainty in CO2 saturation tends to be associated with the highCO2 saturated regions, while lower uncertainty and highervalue in the estimation of the patchiness exponent tend to be

associated with the high CO2 saturated regions. For example,the estimations of patchiness exponent vary in the range of31–36 with ± 5 uncertainty at depth of 910–940 m where theCO2 saturation is reaching 0.50. High CO2 saturation anduniform mixing tend to happen at the same locations.

It is worth noting that there are large CO2 saturationsabove the top Utsira in Fig. 7, especially for the case whereBrie exponent is equal to 1 (top panel). This feature in over-burden should not be interpreted as leaking CO2. It is linkedto the low velocity patch (green-blue colours) observed inFig. 4(e). It is important to note that this high saturation valueis only visible for the patchy saturation case (e = 1) and thatthe results of CO2 saturation estimates are affected by uncer-tainties both related to the first (FWI) and second inversionsteps (RPI). The full waveform inversion (FWI) step is depen-dent on data quality that is different for the baseline (1994vintage) and monitor (2008 vintage) datasets, making any

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 13: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

Sleipner (North Sea) from seismic tomography 13

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

0

0.2

0.4

0.6

0.8

CO

2 sa

tura

tion

unce

rtai

nty

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

0

0.2

0.4

0.6

0.8

CO

2 sa

tura

tion

unce

rtai

nty

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

0

0.2

0.4

0.6

0.8

CO

2 sa

tura

tion

unce

rtai

nty

Figure 8 From top to bottom: inversion uncertainty of CO2 saturation estimated by RPI using patchiness exponent e = 1, e = 5 and exponentinverted (given in Fig. 9).

quantification of 4D changes difficult. For this reason, weopted for a parallel time-lapse FWI strategy. The baselinedata is used in our case to derive constraints on rock frameproperties in the reservoir. Smearing effect in FWI resultscan also be a source of ’ghost’ changes above and below thereservoir as noticed in Fig. 4. In addition, the overburdenshales have low porosity (<20%) and complex mineralogy,so additional caution must be considered when lookingat CO2 saturations in the overburden because of greateruncertainties in a priori constraints. Finally, Biot–Gassmanntheory combined with the effective fluid phase rock physicsmodel have been shown to be relevant for fluid substitutionin brine–CO2 saturated sands but its applicability for shalesis less straightforward (Queiber and Singh 2013; Bergmannand Chadwick 2015; Dupuy et al. 2017; Falcon-Suarez et al.

2018). High CO2 saturation estimates in the overburden are

consequently not reliable and no conclusion of any leakingCO2 should be derived from this result.

Figure 10 shows the one-dimensional profiles of CO2 sat-uration and patchiness exponent for the different test cases.The profile is extracted at an offset of 3240 m going throughthe high CO2 saturation zone. The main CO2–brine saturatedlayers appear between 900 and 950 m deep with CO2 satu-ration ranging from 0.60 to 0.90. It is interesting that bothpatchy mixing (e = 1 or e = 5 depending on authors) and uni-form mixing (e = 40) models show CO2 saturations reaching0.90. The CO2 saturation increases when the patchiness expo-nent decreases, which is directly related to the shape of the ve-locity versus saturation curves (Fig. 1). It is worth noting thatthe results for exponent equal to 1 are enveloping the otherresults showing more high-resolution variations. When we in-vert patchiness exponent and CO2 saturation at the same time,

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 14: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

14 H. Yan et al.

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

4914192429

Brie

exp

onen

t

750

850

950

1050

1150

Dep

th (

m)

1500 1900 2300 2700 3100 3500Distance (m)

4914192429

Brie

exp

onen

t unc

erta

inty

Figure 9 From top to bottom: estimation of patchiness exponent and related uncertainty. The patchiness has been inverted together with CO2

saturation (given in Fig. 7, bottom panel).

Dep

th (

m)

750

800

850

900

950

1000

1050

1100

1150

1200

CO2 saturation (%)

0 50 100

e=1e=5e=40e inverted

750

800

850

900

950

1000

1050

1100

1150

1200

CO2 saturation (%)

0 50 100

Mean value, e=5Mean+1stdMean-1std

750

800

850

900

950

1000

1050

1100

1150

1200

Brie exponent0 20 40

Mean value eMean+1stdMean-1std

Figure 10 From left to right: 1D profiles of CO2 saturation estimation using different patchiness exponents, 1D profiles of CO2 saturationestimation and related uncertainty for e = 5 and 1D profiles of patchiness exponent and related uncertainty. The 1D profiles are extracted for anoffset equal to 3240 m. The dotted lines correspond to the mean values and the dashed lines stand for the mean value ± one standard deviation.

we observe more oscillations due to the under-determined in-version system. The uncertainty for the estimation of CO2

saturation varies in the range of ±0.12–0.17, when we areusing a known Brie exponent. When we invert the CO2 sat-uration and the patchiness exponent together, the associateduncertainty of CO2 saturation goes up to ±0.25.

D I S C U S S I O N

In our study, velocity models from both the baseline and mon-itor datasets and these velocity models are used as input toinfer rock physics properties. The inversion of the monitordataset provides clear indications about the geometry and

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 15: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

Sleipner (North Sea) from seismic tomography 15

lateral extent of the CO2 plume with the lowest P-wave ve-locities being in the uppermost layers. We opted for a paralleltime-lapse strategy where models from velocity analysis areused as initial models for the inversion. Alternative strategies(Asnaashari et al. 2015), including the use of a sequentialtime-lapse inversion and the introduction of a priori informa-tion (for example about the overburden), can contribute tothe improvement of the derived models.

The values of CO2 saturation obtained for inline 1836are higher than those obtained in Dupuy et al. (2017). Theinline data used in our study is located very close to the injec-tion point, while the one-dimensional section shown in Dupuyet al. (2017) is located approximately 500 m away from theinjection point. For comparison, Queiber and Singh (2013)get values up to 0.90 using 2006 vintage, while Golding et al.

(2011) suggest that saturation is probably higher than 0.30.Recent work by Ghosh et al. (2015) gives estimations between0.20 and 0.80 depending on the mixture type (patchiness ex-ponent). In previous tests using the same methodology shownby Dupuy et al. (2017), rock frame properties are derivedfrom well-log data acquired at the injection well. The effectof a priori properties (i.e., rock frame, grains and fluid pa-rameters) on CO2 saturation estimates can strongly affect theresults, and these a priori parameters should be handled withcare. Estimating them from baseline data can be a better so-lution. It is anyway worth keeping in mind that uncertaintyrelated to baseline estimates (bulk and shear moduli as wellas porosity) is not negligible and is a key factor in the CO2

saturation quantification. Even if our sensitivity tests showthat the mean values of these baseline properties estimatedfrom P-wave velocity only are close to the true value, uncer-tainty is not negligible and other sensitivity tests show that theCO2 saturation can be changed by 0.10 to 0.20 if the baselineproperties are different.

Regarding the mixture distribution, Sen, Ghosh andVedanti (2016) use a capillary pressure equilibrium theory andshow that the CO2 distribution at Sleipner is more uniformfor high CO2 saturation and more patchy for low CO2 satu-ration. We found similar trends in our results. For example,the patchiness exponent is higher for the highly CO2 saturatedlayers, which is indicating a more uniform mixing while thepatchiness exponent is smaller for the less CO2 saturated lay-ers. Ghosh et al. (2015) also found that the CO2 phase is notfully patchy or uniformly saturated in the reservoir. Zhang,Lu and Zhu (2014) and Chadwick et al. (2010) propose anapproach using history-matching simulations and 4D seismicdata. They get higher values of CO2 saturation with some val-ues up to 0.98. In our study, the highly saturated layers show

saturations varying between 0.60 and 0.90 with an uncer-tainty of 0.12–0.20. Lumley et al. (2008) suggest that it is diffi-cult to quantify the CO2 saturation when it is larger than 0.30.In our study, we also observe that the estimated uncertainty ishigher for higher CO2 saturation (Figs 1 and 3). Relevant un-certainty assessment and quantification still need to be carriedout. Proper uncertainty propagation from seismic imaging (us-ing the method suggested in Eliasson and Romdhane (2017)for example) to rock physics inversion can be taken into ac-count using a fully Bayesian formulation. More constraints tothe inversion process could be added to mitigate the trade-offbetween saturation and patchiness exponent (standing for therock physics model itself). For example, using additional in-puts from seismic inversion (attenuation, shear wave velocity,AVO attributes) or from other geophysical techniques (grav-ity, CSEM) can lead to reduce the non-uniqueness of the in-version (Gao, Abubakar and Habashy 2012; Liang, Abubakarand Habashy 2016; Subagjo et al. 2018).

CONCLUSIONS

We derive a quantitative spatial distribution of CO2 satura-tion through the combination of full waveform inversion androck physics inversion. The two-step workflow is applied toa selected inline from both baseline and monitor surveys atSleipner. We first estimate rock frame properties based onbaseline data that are then used as an input to the monitorvintage study. We find that CO2 saturations can reach 0.60–0.90 ± 0.12–0.25 near the injection point when the patchinessexponent is assumed to be known. The uncertainty of the CO2

saturation estimation is relatively low for most cases but tendsto increase for higher CO2 saturations. We show the possibil-ity to estimate the CO2–brine mixture law by estimating thepatchiness exponent and CO2 saturation simultaneously. Thepatchiness exponent is globally varying from 5 to 33, whichindicates a distribution of CO2 and brine between fully patchy(e = 1) and fully uniform (e = 40) mixings. Uniform mixingprevails when CO2 saturation is very high.

ACKNOWLEDGEMENTS

This publication has been produced with support from theNCCS Centre, performed under the Norwegian researchprogram Centres for Environment-friendly Energy Research(FME). The authors acknowledge the following partnersfor their contributions: Aker Solutions, ANSALDO En-ergia, CoorsTek Membrane Sciences, Gassco, KROHNE,Larvik Shipping, Norcem, Norwegian Oil and Gas, Quad

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 16: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

16 H. Yan et al.

Geometrics, Shell, Equinor, TOTAL and the Research Councilof Norway (257579/E20). We thank also the uniCQue KPNproject (233716/E20). We also thank Malcolm Sambridgefor providing NA software and Equinor for sharing Sleipnerdataset.

REFERENCES

Adler P.M., Jacquin C.G. and Thovert J.F. 1992. The formation factorof reconstructed porous media. Water resources research 28, 1571–1576.

Arts R., Chadwick A., Eiken O., Thibeau S. and Nooner S. 2008. Tenyears’ experience of monitoring CO2 injection in the Utsira Sandat Sleipner, offshore Norway. First break 26, 65–72.

Arts R., Eiken O., Chadwick A., Zweigel P., Van der Meer L. andZinszner B. 2004. Monitoring of CO2 injected at Sleipner usingtime-lapse seismic data. Energy 29, 1383–1392.

Arts R., Elsayed R., Van Der Meer L., Eiken O., Ostmo S., Chad-wick A., Kirby G. and Zinszner B. 2002. Estimation of themass of injected CO2 at Sleipner using time-lapse seismic data.64th EAGE Conference & Exhibition, Florence, Italy, ExpandedAbstracts.

Asnaashari A., Brossier R., Garambois S., Audebert F., Thore P. andVirieux J. 2015. Time-lapse seismic imaging using regularized full-waveform inversion with a prior model: which strategy? Geophys-ical prospecting 63, 78–98.

Bergmann P. and Chadwick A. 2015. Volumetric bounds on subsur-face fluid substitution using 4D seismic time shifts with an applica-tion at Sleipner, North Sea. Geophysics 80, B153–B165.

Biot M.A. 1956a. Theory of propagation of elastic waves in a fluid-saturated porous solid. I. Low-frequency range. The Journal of theacoustical society of America 28, 168–178.

Biot M.A. 1956b. Theory of propagation of elastic waves in a fluid-saturated porous solid. II. Higher frequency range. The journal ofthe acoustical society of America 28, 179–191.

Boait F., White N., Bickle M., Chadwick R., Neufeld J. and HuppertH. 2012. Spatial and temporal evolution of injected CO2 at theSleipner field, North Sea. Journal of geophysical research: solidearth 117, B03309–B03309.

Bohm G., Carcione J.M., Gei D., Picotti S. and Michelini A. 2015.Cross-well seismic and electromagnetic tomography for CO2 de-tection and monitoring in a saline aquifer. Journal of petroleumscience and engineering 133, 245–257.

Brie A., Pampuri F., Marsala A. and Meazza O. 1995. Shear sonicinterpretation in gas-bearing sands. Conference shear sonic inter-pretation in gas-bearing sands. SPE Annual Technical Conferenceand Exhibition, Dallas, TX, October 1995. Society of petroleumengineers.

Carcione J.M. and Picotti S. 2006. P-wave seismic attenuation byslow-wave diffusion: effects of inhomogeneous rock properties.Geophysics 71, O1–O8.

Chadwick A., Williams G., Delepine N., Clochard V., Labat K., Stur-ton S., et al. 2010. Quantitative analysis of time-lapse seismic mon-itoring data at the Sleipner CO2 storage operation. The leadingedge, 29, 170–177.

Chadwick R., Holloway S., Kirby G., Gregersen U. and JohannessenP. 2000. The Utsira Sand, Central North Sea–an assessment ofits potential for regional CO2 disposal. Proceedings of the 5thInternational Conference on Greenhouse Gas Control Technolo-gies (GHGT-5), Cairns, Australia. 2000, pp. 349–354. CSIROPublishing.

Chadwick R., Zweigel P., Gregersen U., Kirby G., Holloway S. andJohannessen P. 2004. Geological reservoir characterization of aCO2 storage site: the Utsira Sand, Sleipner, northern North Sea.Energy 29, 1371–1381.

Dupuy B., Garambois S. and Virieux J. 2016. Estimation of rockphysics properties from seismic attributes—Part 1: strategy andsensitivity analysis. Geophysics 81, M35–M53.

Dupuy B., Romdhane A., Eliasson P., Querendez E., Yan H., Tor-res V.A., et al. 2017. Quantitative seismic characterization ofCO2 at the Sleipner storage site, North Sea. Interpretation 5,SS23–SS42.

Dupuy B. and Stovas A. 2016. Effect of anelastic patchy saturatedsand layers on the reflection and transmission responses of aperiodically layered medium. Geophysical prospecting 64, 299–319.

Eiken O., Brevik I., Arts R., Lindeberg E. and Fagervik K. 2000.Seismic monitoring of CO2 injected into a marine acquifer. SEGtechnical program, expanded abstracts 2000 1623–1626.

Eliasson P. and Romdhane A. 2017. Uncertainty quantificationin waveform-based imaging methods-a Sleipner CO2 monitoringstudy. Energy procedia 114, 3905–3915.

Falcon-Suarez I., Papageorgiou G., Chadwick A., North L., Best A.I.and Chapman M. 2018. CO2-brine flow-through on an UtsiraSand core sample: experimental and modelling. Implications forthe Sleipner storage field. International journal of greenhouse gascontrol 68, 236–246.

Furre A.-K., Eiken O., Alnes H., Vevatne J.N. and Fredrik A. 2017.20 years of monitoring CO2-injection at Sleipner. Energy procedia114, 3916–3926.

Furre A.K. and Eiken O. 2014. Dual sensor streamer technology usedin Sleipner CO2 injection monitoring. Geophysical prospecting 62,1075–1088.

Furre A.-K., Kiær A. and Eiken O. 2015. CO2-induced seismic timeshifts at Sleipner. Interpretation 3, SS23–SS35.

Gao G., Abubakar A. and Habashy T.M. 2012. Joint petrophysi-cal inversion of electromagnetic and full-waveform seismic data.Geophysics 77, WA3–WA18.

Gasda S., Nilsen H. and Dahle H. Year. Upscaled models for CO2

migration in geological formations with structural heterogeneity.13th European Conference on the Mathematics of Oil Recovery,Expanded Abstracts.

Gassmann F. 1951. Uber die Elastizitat poroser Medien. Vierteljahrss-chrift der Naturforschenden Gesellschaft in Zurich, 96, 1–23.

Ghaderi A. and Landrø M. 2009. Estimation of thickness and velocitychanges of injected carbon dioxide layers from prestack time-lapseseismic data. Geophysics 74, O17–O28.

Ghosh R., Sen M.K. and Vedanti N. 2015. Quantitative interpretationof CO2 plume from Sleipner (North Sea), using post-stack inversionand rock physics modeling. International journal of greenhouse gascontrol 32, 147–158.

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17

Page 17: CO 2 saturation estimates at Sleipner (North Sea) from ...barn/Myarticles/Yan2019.pdf · Geophysical Prospecting, 2018 doi: 10.1111/1365-2478.12693 CO 2 saturation estimates at Sleipner

Sleipner (North Sea) from seismic tomography 17

Golding M.J., Neufeld J.A., Hesse M.A. and Huppert H.E. 2011.Two-phase gravity currents in porous media. Journal of fluid me-chanics 678, 248–270.

Halland E.K., Gjeldvik I., Johansen W., Magnus C., Meling I., Ped-ersen S., et al. 2011. CO2 storage atlas Norwegian North Sea.Norwegian petroleum directorate.

Hashin Z. and Shtrikman S. 1963. A variational approach to thetheory of the elastic behaviour of multiphase materials. Journal ofthe mechanics and physics of solids 11, 127–140.

Hustedt B., Operto S. and Virieux J. 2004. Mixed-grid and staggered-grid finite-difference methods for frequency-domain acoustic wavemodelling. Geophysical journal international 157, 1269–1296.

Johnson D.L., Koplik J. and Dashen R. 1987. Theory of dynamic per-meability and tortuosity in fluid-saturated porous media. Journalof fluid mechanics 176, 379–402.

Liang L., Abubakar A. and Habashy T.M. 2016. Reservoir propertymapping and monitoring from joint inversion of time-lapse seis-mic, electromagnetic, and production data. Geophysics 81, ID73–ID84.

Lindeberg E. 2013. Calculation of thermodynamic properties of CO2,H2O and their mixtures also including salt with the Excel macro“CO2 Thermodynamics”. SINTEF report.

Lumley D., Adams D., Wright R., Markus D. and Cole S. 2008.Seismic monitoring of CO2 geo-sequestration: realistic capabilitiesand limitations. SEG technical program expanded abstracts 2841–2845.

Mavko G., Mukerji T. and Dvorkin J. 2009. The Rock Physics Hand-book: Tools for Seismic Analysis of Porous Media. Cambridge Uni-versity Press.

Metz B., Davidson O., De Coninck H., Loos M. and Meyer L. 2005.IPCC special report on carbon dioxide capture and storage. In-tergovernmental Panel on Climate Change, Geneva (Switzerland).Working Group III.

Michael K., Golab A., Shulakova V., Ennis-King J., Allinson G.,Sharma S., et al. 2010. Geological storage of CO2 in salineaquifers—a review of the experience from existing storage opera-tions. International journal of greenhouse gas control 4, 659–667.

Papageorgiou G., Amalokwu K. and Chapman M. 2016. Theoreticalderivation of a Brie-like fluid mixing law. Geophysical prospecting64, 1048–1053.

Pratt R.G., Shin C. and Hick G. 1998. Gauss–Newton and full New-ton methods in frequency–space seismic waveform inversion. Geo-physical journal international 133, 341–362.

Pride S. 2005. Relationships between Seismic and Hydrological Prop-erties. In: Hydrogeophysics: Water Science and Technology Li-brary, (eds Y. Rubin and S.S. Hubbard), pp. 253–284, Springer.

Queiber M. and Singh S.C. 2013. Full waveform inversion in thetime lapse mode applied to CO2 storage at Sleipner. Geophysicalprospecting 61, 537–555.

Romdhane A. and Querendez E. 2014. CO2 characterization at theSleipner field with full waveform inversion: application to syntheticand real data. Energy procedia 63, 4358–4365.

Romdhane A., Querendez E. and Ravaut C. 2014. CO2 thin-layerdetection at the Sleipner field with full waveform inversion: appli-cation to synthetic and real data. Energy procedia 51, 281–288.

Sambridge M. 1999. Geophysical inversion with a neighbourhoodalgorithmoII. Appraising the ensemble. Geophysical journal inter-national 138, 727–746.

Sen A., Ghosh R. and Vedanti N. 2016. Saturation estimation from4D seismic data from Sleipner Field by a capillary pressure-basedrock-physics model. SEG technical program, expanded abstracts3314–3321.

Singh V.P., Cavanagh A., Hansen H., Nazarian B., Iding M. andRingrose P.S. 2010. Reservoir modeling of CO2 plume behaviorcalibrated against monitoring data from Sleipner, Norway. SPE An-nual Technical Conference and Exhibition, Florence, Italy, Septem-ber 2010. Society of Petroleum Engineers.

Subagjo I., Dupuy B., Park J., Romdhane A., Querendez E. and StovasA. 2018. Joint rock physics inversion of seismic and electromag-netic data for CO2 monitoring at Sleipner. 24th European Meetingof Environmental and Engineering Geophysics, Porto, Portugal,Expanded Abstracts.

Teja A. and Rice P. 1981. Generalized corresponding states methodfor the viscosities of liquid mixtures. Industrial & engineeringchemistry fundamentals 20, 77–81.

Voigt W. 1928. Lehrbuch der Kristallphysik. Berlin: Leipzig.White D. 2013. Toward quantitative CO2 storage estimates from

time-lapse 3D seismic travel times: an example from the IEA GHGWeyburn–Midale CO2 monitoring and storage project. Interna-tional journal of greenhouse gas control 16, S95–S102.

Williams G. and Chadwick A. 2012. Quantitative seismic analysis ofa thin layer of CO2 in the Sleipner injection plume. Geophysics 77,R245–R256.

Zhang G., Lu P. and Zhu C. 2014. Model predictions via historymatching of CO2 plume migration at the Sleipner Project, Norwe-gian North Sea. Energy procedia 63, 3000–3011.

Zweigel P. 2000. Reservoir geology of the storage unit in the SleipnerCO2-injection case. SINTEF.

Zweigel P., Arts R., Lothe A.E. and Lindeberg E.B. 2004. Reservoirgeology of the Utsira Formation at the first industrial-scale un-derground CO2 storage site (Sleipner area, North Sea). Geologicalsociety publications 233, 165–180.

C© 2018 European Association of Geoscientists & Engineers, Geophysical Prospecting, 1–17