Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

Petroleum Geoscience

Assessing the seismic AVA detectability of CO2 storage sites using novel time-lapse attributes Arash Jafargandomi and Andrew Curtis Petroleum Geoscience 2013, v.19; p357-374. doi: 10.1144/petgeo2012-043

Email alerting service

click here to receive free e-mail alerts when new articles cite this article

Permission request

click here to seek permission to re-use all or part of this article

Subscribe

click here to subscribe to Petroleum Geoscience or the Lyell Collection

Notes

© The Geological Society of London 2013

2012-043 2013

research-articleArticleXXX10.1144/petgeo2012-043Novel time-lapse AVA attributesA. JafarGandomi & A. Curtis

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

Assessing the seismic AVA detectability of CO2 storage sites using novel time-lapse attributes Arash Jafargandomi* and Andrew Curtis School of GeoSciences, University of Edinburgh, King’s Buildings, West Mains Road, Edinburgh EH9 3JW, UK *Corresponding author (e-mail: [email protected]) Abstract: Monitoring of stored carbon dioxide (CO2) in subsurface reservoirs is fundamental to operation and management of the storage site, and is a requirement of some national and international legislation. As a consequence, effectiveness of monitorability (the ability to observe the evolving location of subsurface CO2) for any given level of investment in monitoring technology is a significant investment uncertainty that must be assessed along other components of the storage-site selection criteria (e.g. capacity, injectivity and storage economies). We develop a workflow to assess the time-lapse seismic detectability of changes in subsurface aquifer reservoirs by analysing expected changes in seismic amplitude variation with angle (AVA) in the field. Laboratory measurements are used to calculate the seismic response of the reservoir at different saturations and pressures. We include the scattering effect of material above and below the reservoir by using a finite-difference, full-waveform modelling approach AVA analysis then assimilates local site effects into the detectability assessment. We show that performing waveform modelling which includes local geological heterogeneities above and below the reservoir interval is essential to assess the storage site monitroability. In order to quantify expected time-lapse changes in the seismic response, we introduce a new set of robust time-lapse attributes based on time–frequency decomposition. The attributes effectively separate amplitude and phase changes (time-shifts) of time-lapse seismic records, and allow us to quantify their repeatability against the background noise. Furthermore, the frequency-dependent nature of the attributes provides a quantification of the frequency–domain effects of time-lapse changes. The approach is employed to assess the detectability of supercritical CO2 in two analogue storage sites in the near-shore UK North Sea. Analysis of laboratory measurements and AVA responses indicate the contrasting monitorability of the two sites, which helps decision making about further site investigation and development. Application of the approach to hydrocarbon reservoir monitoring is straightforward. Introduction The process of capturing carbon dioxide (CO2) and injecting it into deep subsurface saline aquifers or depleted hydrocarbon reservoirs is a potentially important method to reduce atmospheric emissions of CO2 from large point-source emitters such as power stations (e.g. van der Meer 1993; Bachu et al. 1994; Law & Bachu 1996; Bachu 2000; Haszeldine 2009; Scott et al. 2012). Generally, hydrocarbon reservoirs appear to offer lowerrisk storage sites than saline aquifers owing to the existence of relatively higher-quality and more abundant pre-existing data, and to the availability of a known seal provided by geological caprock proved by having trapped hydrocarbon for millions of years. Also, CO2 has been injected into reservoirs successfully within the oil industry for many years for the purpose of enhanced oil recovery (EOR). The disadvantages of hydrocarbon reservoirs are that drilled wells and the production itself may have created leakage pathways, and they also usually have relatively small storage volume compared to saline aquifers. So, Petroleum Geoscience, Vol. 19, 2013, pp. 357–374 doi: 10.1144/petgeo2012-043 Published Online First on October, 16, 2013

despite the higher uncertainty associated with saline aquifers, due to their size they are becoming increasingly attractive targets for CO2 storage (e.g. Haszeldine 2009). Targeting saline aquifers for CO2 storage makes it essential that high-quality methods of site evaluation exist (e.g. Ringrose & Simone 2009), and the ability to detect and track post-injection changes to reduce uncertainties is essential. Assessing the monitorability of the storage site is a part of any site selection process (e.g. White et al. 2005; Vanorio et al. 2010; JafarGandomi & Curtis 2011a, b, 2012). The principle aim of monitoring is to be able to observe and track major changes in the three-dimensional (3D) distribution of CO2 during the injection phase, post-injection site management, and during post-closure stewardship. JafarGandomi & Curtis (2011a) defined a storage site to be monitorable if: (1) geophysical monitoring is possible within existing practical and financial constraints; (2) the spatial resolution is sufficient to image the spatial position of injected CO2 to within the desired level of location uncertainty; © 2013 EAGE/The Geological Society of London

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

358

Novel time-lapse AVA attributes

(3) changes in geophysically measurable signals due to CO2 injection are detectable above measurable noise; and (4) there is sufficient petrophysical resolution and uncertainty-reduction for petrophysical and fluid parameter estimates to fulfil the monitoring objectives. Neglecting practicality and cost for now, detectability of petrophysical changes in storage reservoir rocks due to fluid injection/production is the minimum requirement for a site to be monitorable: the geophysical monitoring techniques employed should be able to detect where some minimum threshold volume or saturation of CO2 has been exceeded within any specified spatial subvolume of a subsurface reservoir, or after leakage into the overburden. However, existing definitions of monitorability require that saturations of CO2 can also be estimated (e.g. JafarGandomi & Curtis 2012). In this paper, we develop a workflow to quantify the detectability of changes at potential storage sites. Seismic reflection amplitude variations with source-toreceiver offset (AVO) or equivalent with the angle between incident and reflected rays at the reflector (AVA) have been used extensively in the hydrocarbon industry to detect and identify reservoir fluids (e.g. Castagna & Smith 1994; Castagna et al. 1998), and to monitor changes in reservoir parameters (Tura & Lumley 1999; Landrø 2001; Veire et al. 2006). We develop an approach based on the AVA technique to assess the site monitorability using limited available data (specifically, sonic welllogs together with rock samples taken from analogue outcrops or drilled boreholes). Under such situations, prediction of the AVA response of the reservoir–caprock interface using the Zoeppritz equations provides valuable information about the detectability of possible changes. Further, we apply full-waveform forward modelling of common midpoint (CMP) gathers in order to include local site geology (i.e. heterogeneity both above and below the reservoir) in the AVA analysis. Although, such modelling presents a more realistic estimate of the AVA response, methods to quantify time-lapse changes by analysing different attributes of resulting waveforms are far from standardized, and, indeed, the application of technique or use of one attribute over another can lead to a degree of subjectivity (attribute-dependences) in the conclusions about time-lapse detectability. One of the main causes of such subjectivity is that changes in amplitude and changes in phase of seismic waveforms become confounded in existing methods and attributes (e.g. Ghaderi et al. 2010). To overcome this, we propose to quantify time-lapse changes using a set of robust time-lapse attributes based on the time–frequency decomposition of seismic records; this optimally separates the amplitude and phase changes in waveforms, thus removing concormitant subjectivity in waveform analysis. In the following section we introduce the monitorability assessment workflow. Then we employ the workflow to assess the monitorability of two analogue CO2 storage sites in the nearshore UK North Sea. These are not sites at which CO2 will actually be injected but, rather, have been studied in detail as analogues for potential offshore aquifer storage reservoirs (Smith et al. 2011). Finally, we discuss results of the assessment and implications for the future application of the novel time-lapse attributes introduced herein.

Detectability Assessment Workflow Overview We consider a situation where detectability of changes at the site has to be assessed based on limited pre-existing data (likely to be the case for most aquifer storage sites in the UK North Sea). The data assumed to be available from site survey and

preliminary research include specific laboratory measurements on core plugs from the selected caprock and reservoir rock samples. At earlier stages of site investigation, when no samples from a borehole are available, samples of geological analogues, meaning accessible rock outcrops that are expected to match the properties of the subsurface caprock and reservoir, may be used to assess the geophysical response of the reservoir. Use of analogous would, of course, increase uncertainty on parameter estimates owing to mismatch between analogue and reservoir geologies, and to pressure-related differences due to the reservoir’s overburden. Figure 1 illustrates the workflow used to assess the detectability of changes at prospective sites. The laboratory experiments test the direct impact on observable geophysical parameters (e.g. P- and S-wave velocities and density) of injecting supercritical CO2 into the rock pore-space. These data are used to calibrate the petrophysical model (the relationship between P- and S-wave velocities fluid saturations at different pressures) that is required to predict geophysical parameters of the reservoir rocks under different injection scenarios. Sonic-log data are used to construct a velocity model that spans above, throughout and below the reservoir; this will be used to investigate the impact of local site geological heterogeneity on site monitorability. A primary goal of much industrial geophysical monitoring is to discriminate between different saturating fluids in porespaces. AVO/AVA analysis has proven to be effective for this purpose in the hydrocarbons industry when discrimination is typically required between oil, gas and brine. Several theoretical relationships have been proposed to predict the amplitude variation of reflected seismic waves from subsurface interfaces (Knott 1899; Zoeppritz 1919; Aki & Richards 1980; Waters 1981). Recently, there have been efforts to include overburden effects (Skopintseva & Stovas 2010) and frequency-dependence (Liu et al. 2011) in AVA analysis. Modelling of full-waveform common mid-point (CMP) gathers is routinely used in the hydrocarbon industry and includes all of these effects. Further, it may also include the effect of seismic attenuation. Synthetic CMP gathers may be generated for a 1D (vertically-varying) medium by using analytical or semi-analytical wavefield modelling methods such as so-called reflectivity modelling (e.g. Kennett & Clarke 1983). However, using a large number of layers to incorporate scattering and velocity-gradient effects renders these approaches inefficient. To calculate synthetic CMPs we use the finite-difference time-domain (FDTD) scheme developed by JafarGandomi & Takenaka (2007, 2013). The algorithm uses staggered-grid finite-difference operators in time and space. The advantages of this scheme are that it generates synthetic wavefields in the τ−p (plane-wave) domain that propagates in 3D space through a velocity model that varies in 1D, it is highly efficient due to the employed 1D approximation of the 3D Earth model, and it can incorporate any frequency-dependent attenuation model (described by 1/Q where Q is the quality factor). We generate synthetic CMP gathers in the τ−p domain by gathering synthetic traces for a range of incident angles. A petrophysical model that relates rock and pore-fluid properties to geophysically observable properties is essential in order to predict geophysical observations of the reservoir under different injection scenarios. In partially saturated rocks, the bulk modulus depends not only on the degree of saturation but also on the mesoscopic and microscopic characterisation of saturation. JafarGandomi & Curtis (2012) proposed an approach to account for the impact of mesoscopic and microscopic characteristics of saturation on seismic waves based on a combination of available rock physics models. They combine the White (1975) model, which accounts for the mesoscopic effects, with the

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

359

A. JafarGandomi & A. Curtis

(Lumley 2001). The resulting 4D anomalies may be manifest as amplitude changes and/or time-shifts (phase changes) on the seismic records. Each of these attributes carries information about different types of changes in the subsurface reservoir: while an amplitude change might indicate a variation in saturation of the reservoir, a time-shift is an important indicator of changes in the reservoir pressure and compaction, or changes in the overburden (e.g. Landrø & Stammeijer 2004; Fuck et al. 2009). In most cases, separation of amplitude and phase change are desirable but difference cross-sections and maps do not discriminate sufficiently between them. Cross-correlation is often used to estimate the time-shift between repeated arriving waves. However, such estimates are not robust for short time-windows (Ursenbach & Bancroft 2001) and their accuracy does not improve with increasing signal-to-noise ratio (Jacovitti & Scarano 1993). Here, we describe an approach based on time–frequency decomposition to separate the amplitude and phase differences between two or more vintages of time-lapse records. We follow a similar approach to that developed by Kristekova et al. (2006) for quantitative comparison of earthquake seismograms. Any signal in the time domain s(t) can be represented by its time–frequency decomposition W(t, f) calculated by applying a continuous wavelet transform (CWT) to the signal (see the Appendix). Recently, it has been shown that non-parametric time–frequency distributions, such as the spectrogram (SP) and Wigner–Ville distribution (WVD), provide better trade-off between time and frequency resolution than the CWT (Wang 2010; Liu & Fomel 2012). Here, we build on a SP representation that is the short-time quadratic integral measure of the energy distribution of s(t): W (t, f ) =





2

s (τ )h(τ − t )e −2π f τ dτ

(1)

−∞

Fig. 1. Workflow to predict the AVO monitorability of a set of reservoirs (sites A, B, C, …) if they were to undergo CO2 injection.

Pham et al. (2002) model, which accounts for microscopic effects, by adjusting the matrix bulk modulus of the latter with the frequency-dependent bulk modulus of the former. We employ the same petrophysical model in this work, which we call the hybrid White–Pham model from hereon. An assessment of the repeatability of any time-lapse survey is a requirement for successful site monitoring, which is the last stage of the detectability assessment workflow (Fig. 1). Changes between surveys in the acquisition system, and in both natural and instrumental noise cause changes in recordings regardless of whether any changes in the reservoir occurred (Lumley, 2001). Either of the individual baseline or monitor surveys may also be affected by source noises owing to the short time-intervals between successive shots, and the inaccuracy in the timing system (Landrø 2008). These effects reduce the signal-to-noise ratio of the time-lapse survey and, hence, repeatability of the seismic records. There have been many efforts to assess (e.g. Landrø 1999; Kragh & Christie 2002) and improve (Eiken et al. 2003) repeatability of seismic surveys.

Time-Lapse Monitoring The simplest way to demonstrate expected time-lapse (4D) changes on seismic records is the difference technique: the synthetic seismic data that are expected to be recorded at repeated surveys are subtracted from the baseline survey, and the resulting anomalies are presented on difference cross-sections and maps



where h(t) is a localizing window function such as a Box car function. We will also show comparisons to the alternative CWT method. Following Kristekova et al. (2006), a change in the signal amplitude is estimated by calculating the non-normalized envelope difference ( δ E ):

δ E ( t , f ) = W2 ( t , f ) − W1 ( t , f )

(2)



where W1(t, f) and W2(t, f) are the time–frequency representatives of the baseline record and the monitor (repeated) record, respectively. Changes in the signal phase are estimated by calculating the non-normalized phase difference ( δθ ):

δθ ( t , f ) = W1 ( t , f )

{ Arg W ( t , f ) − Arg W ( t, f ) } 2

1

π



(3)

where Arg . indicates the phase angle θ (in radians) of the function to which it is applied. Including the envelope W1 ( t , f ) in equation (3) imposes the shape of the baseline signal on the phase difference; this equalizes the spatial resolution of envelope and phase difference estimates in equations (2) and (3), respectively. Finally, the time–frequency envelope δE and phase differences δθ are the result of normalizing by the global maximum values for the baseline signal:

δE ( t , f ) =

δ E (t, f )

(

max W1 ( t , f )

)

(4)

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

360

Novel time-lapse AVA attributes δθ ( t , f ) =

δθ ( t , f )

(

max W1 ( t , f )

)

.



δθ =

δE then varies with the relative amplitude values, whereas δθ varies with the phase differences. It is possible to normalize δE and δθ locally with respect to W1 ( t , f ) instead of by its global maximum value but such normalization creates spiky δE and δθ distribution due to divisions by near-zero values. This effect is especially significant for signals contaminated with noise. If we wish to estimate the detectability of changes against the background noise level, δE and δθ may alternatively be normalized by the maximum noise amplitude. It is possible to estimate the amplitude and phase changes as a function either of time or frequency alone by integrating δ E and δθ over frequency or time, respectively. The time-projected counterparts of δ E ( t , f ) and δθ ( t , f ) , respectively, are: f max

δ E (t ) =

∫ δ E ( t , f ) df

f min

 f max max  ∫ W1 ( t , f ) df f  min

(6)

   



f max

δθ ( t ) =

∫ δθ ( t , f ) df

f min

 f max max  ∫ W1 ( t , f ) df f  min

   

.

(7)



Similarly the frequency-projected counterparts of δ E ( t , f ) and δθ ( t , f ) are: tmax

∫ δ E ( t , f ) dt

δE( f ) =

tmin

 tmax  max  ∫ W1 ( t , f ) dt  t   min 

(8)



tmax

δθ ( f ) =

∫ δθ ( t , f ) dt

tmin

 tmax  max  ∫ W1 ( t , f ) dt  t   min 

(9)



In equations (6)–(9), fmax, fmin, tmax and tmin indicate the upper and lower bounds of the frequency and time windows of interest. Notice that the arguments following δ E and δθ on the left of equations (4)–(9) define the domain in which each function or attribute is defined. It is also useful to describe the total average envelope and phase differences of the baseline, and monitor signals with single-valued measures: f max tmax

δE =

∫ ∫ δ E ( t , f )

f min tmin f max tmax

∫ ∫ W ( t , f ) 1

f min tmin

2

d t df (10)

2

f max tmax

(5)

dtdf

∫ ∫ δθ ( t , f )

f min tmin f max tmax

∫ ∫ W ( t , f ) 1

2

2

dt df dt df

f min tmin



(11)

which may be used to assess the repeatability of the timelapse surveys as whole. From hereon, we approximate the integrals in equations (6)–(11) numerically with summation operators. The following example demonstrates the quantification of amplitude and phase changes in a synthetic signal that includes three seismic events (e1, e2 and e3 in Fig. 2), each one a Gabor wavelet with a dominant frequency of 30 Hz. We modify the three events of the original (baseline) signal to create an altered (monitoring) signal as follows: •• amplitude of e1 is reduced by 20% but its phase is unchanged; •• amplitude of e2 is unchanged but its phase is reduced by 20% •• amplitude and phase of e3 are both reduced, by 20 and 40%, respectively. For comparison we estimate amplitude and phase changes using both SP and CWT for time–frequency decomposition (Figs 2 and 3, respectively). Figure 2a shows the original and modified signals by dashed and solid lines, respectively, while Figure 2b, c shows the amplitude and phase changes of the modified signal with respect to the original signal, as estimated using the SP method. The calculated δ E ( t , f ) shows the maximum value of 20% for e1 and e3, and no value for e2, which corresponds exactly to the changes applied. The calculated δθ ( t , f ) shows the maximum values of 20 and 40% for e2 and e3, respectively, and no value for e1, which again matches exactly the applied changes. The projected amplitude and phase differences are also shown with respect to time and frequency, which present concise information about changes that occurred compared to the baseline signal. Figure 3 indicates that using CWT for time–frequency decomposition provides the same accuracy as using SP (Fig. 1), except that the time-resolution of the estimated attributes using CWT is lower than that provided by SP for lower frequencies but is higher than SP for higher frequencies. It is worth mentioning that a constant time-shift in the monitor signal causes both phase and envelope differences. We compare the SP- and CWT-based estimates of the phaseshift with that of the commonly used cross-correlation approach in Figure 4. The same window-length is used for both of the cross-correlation and SP-based estimates. The estimated phase changes δθ are converted to the corresponding time-shifts ∆t using: ∆t [ s ] =

δθ [ radians ] 2π f c



(12)

where fc is the dominant frequency of the wavelet. The true time-shift values for e2 and e3 are 3.3 and 6.6 ms, respectively. Figure 4 indicates the higher accuracy and resolution of the SP- and CWT-based estimates of the time-shift compared with that of the cross-correlation. From hereon we therefore use the SP method to calculate all time-frequency decompositions.

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

A. JafarGandomi & A. Curtis

361

Fig. 2. Estimation of amplitude and phase changes of synthetic traces using the SP method for time–frequency decomposition. (a) Original (baseline) trace (dashed line) and modified (monitor) trace (solid line). (b) Estimated δE ( t , f ) and its projections on the frequency and time axis. (c) Estimated δθ ( t , f ) and its projections on the frequency and time axis.

Application to Cassem Analogue Storage Sites We now illustrate for real sites both the performance of the new attributes and the detectability assessment method by assessing the AVA detectability of the CASSEM (CO2 Aquifer Storage Site Evaluation and Monitoring) project analogue storage sites (Smith et al. 2011). In the CASSEM project, two sites were selected as analogues of UK offshore fields. They were studied in detail to develop storage-site evaluation and monitoring methodologies (no CO2 will actually be injected at these analogue sites). The two sites are in the near-shore and on-shore UK North Sea, at the Firth of Forth region and the York–Lincolnshire region, which from hereon we refer to as the Forth site and the Lincs site, respectively. The target aquifer for the Lincs site is in the Triassic Sherwood Sandstone Group (SSG). In previous studies, this formation has been observed to have a relatively uniform thickness of 300 m. The postulated injection point for this site lies at a m. The seal is the Mercia Mudstone Group depth of 1200  (MMG), and the underlying formation is the Roxby Formation (gypsum and mudstone). The target aquifer for the Forth site is in the Kinnesswood and Knox Pulpit Formations (K&K), and the caprock is the Ballagan Formation (BGN). The thickness of the potential reservoir is about 300 m, and the main aquifer/seal levels are at an interpreted depth range of 2000–2500 m. The geological interpretation and modelling of these two sites are described in Monaghan et al. (2009). A range of laboratory experiments have been carried out to measure the ultrasonic properties of the reservoir sandstones of the CASSEM analogue storage sites, while saturating the samples

with a range of different proportions of brine v. supercritical CO2, and under a range of stress conditions (Fisher et al. 2010). These experiments are conducted on four sandstone samples: two from the Clashach Quarry, which is considered to be geologically analogous to the reservoir formation at the Forth site, and two samples from the Sherwood Sandstone Formation, which is the actual reservoir formation at the Lincs site (which outcrops some distance from the potential reservoir). Two samples from the Merica Mudstone and Ballagan Formation representing caprocks at the Lincs and Forth sites were also examined for their geophysical properties at different pressures. For the caprock samples of the Lincs site, the measurements were conducted under dry conditions and we use the Gassmann equations to correct for saturation (Smith et al. 2003). The P- and S-wave velocities are measured at 1 and 0.6 MHz frequencies, respectively. A summary of the laboratory measurements on Clashach Quarry (CL1) and Sherwood Sandstone (SSK) samples is given in Table 1, and more details about these measurements can be found in the CASSEM project report (Fisher et al. 2010). For further investigations, we used samples CL1 with a porosity of 22.6% and SSK with porosity of 20% as representative of the two aquifers. Figure 5 shows the measured P- and S-wave velocities and best-fit petrophysical models for samples CL1 (Forth site) and SSK (Lincs site) with respect to SCO2 at various differential pressures. We use the hybrid White–Pham petrophysical model (JafarGandomi & Curtis 2012) with the material parameters given in Table 2. The expected range of differential pressure in the Forth and Lincs sites in the vicinity of the injection points are approximately in the ranges 3500–4000  psi and 2500–3000 psi, respectively. Laboratory measurements indicate

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

362

Novel time-lapse AVA attributes

Fig. 3. Estimation of amplitude and phase changes of synthetic traces using the CWT for time–frequency decomposition. (a) Original (baseline), trace (dashed line) and modified (monitor) trace (solid line). (b) Estimated δE ( t , f ) and its projections on the frequency and time axis. (c) Estimated δθ ( t , f ) and its projections on the frequency and time axis.

Fig. 4. Estimated time-shifts for e2 and e3 in Figure 3a using SPand CWT-based attributes (solid lines with filled and open circles, respectively) and cross-correlation (solid line).

for both samples that, while S-wave velocity does not change significantly with SCO2 , there are more than 200 and 300 m s−1 drops in the P-wave velocities of samples CL1 and SSK, respectively, when the samples are fully saturated with CO2 when compared with brine-filled samples. An increase of pore-fluid pressure is a direct consequence of injecting CO2 into the saline aquifers. Recent studies show that this increase can be in the range of several MPa (e.g. Birkholzer et al. 2009; Vidal-Gilbert et al. 2009). For this reason, to estimate the impact of CO2 injection on the elastic parameters of

the rocks, both saturation and pore-fluid pressure increase (differential pressure decrease) have to be considered, simultaneously. Although increasing SCO2 has different impacts on the P- and S-wave velocities (decreasing the former and increasing the latter), increasing pore-fluid pressure led to a decrease of both P- and S-wave velocities. Figure 6 depicts the trajectories of the P- and S-wave velocity changes due to the combined effects of SCO2 increase and differential pressure decrease. The expected change in P-wave velocity of the reservoir rocks due to CO2 injection is greater when accounting for both SCO2 and pore-pressure increase. However, increasing SCO2 leads to on increase in S-wave velocity, which partially neutralizes the impact of the pore-pressure increase. The combined effects of SCO2 and pore-pressure increase therefore yields smaller overall changes in the S-wave velocity. Figures 5 and 6 imply that accounting for both saturation and differential pressure is necessary to obtain a realistic estimate of the elastic parameter changes expected in the reservoir. Since the frequency of the laboratory measurements (0.6–1 MHz) is much higher than the frequencies used in the field (30 Hz), we scale the measured velocities to the field frequency using the hybrid White–Pham petrophysical model. The scaling is conducted by best-fitting the hybrid model to the laboratory measurements, then reducing the input frequency to 30 Hz in the model while all the other parameters (Table 2) are fixed. The predicted P- and S-wave velocities and attenuation at 30 Hz are considered to be representative of geophysical parameters that would be estimated from a reflection seismic experiment. Table 3 provides the estimated density, P- and S-wave velocities, and P-wave attenuation (QP−1) for the reservoir rocks with SCO2 val-

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

363

A. JafarGandomi & A. Curtis Table 1. Measured P- and S-wave velocities (m s−1), VP and VS, respectively, of the potential reservoir and cap rocks of the Lincs and Forth sites (Fisher et al. 2010) Clashach Sandstone (Forth reservoir)

SCO2

Pd (MPa)

VP (m s−1)

VS (m s−1)

0.0

6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 3.5 6.9 10.4 13.8 17.3 20.7 24.2 27.6

3870 3906 3936 3961 3799 3858 3899 3942 3805 3866 3884 3915 3745 3757 3793 3817 3774 3820 3852 3852 3688 3732 3769 3808 3595 3657 3764 3787 3309 3379 3402 3424 3447 3470 3522 3548

2334 2374 2409 2426 2376 2414 2441 2449 2402 2433 2451 2464 2358 2401 2431 2438 2412 2440 2458 2470 2378 2405 2433 2447 2386 2424 2471 2492 2217 2171 2179 2166 2174 2183 2192 2212

(%)

32.8

41.6

61.1

78.5

86.6

100.0

Ballagan Formation (Forth caprock)

0.0

ues of 0 and 60% at 30 and 1 MHz. Expected differential pressures at the aquifer level are approximately 13.8 and 27.6 MPa for the Lincs and Forth sites, respectively, and we assume a 3.4  MPa drop in differential pressures due to CO2 injection (MacKay et al. 2011). The expected drop in P-wave velocity of sample SSK of about 244 m s−1 at the lower frequency is greater than that of sample CL1 (c. 147 m s−1). The S-wave velocity changes of SSK and CL1 are −13 and +4 m s−1, respectively. The expected changes in density are −38 and −32 kg m−3 for the Lincs and Forth sites, respectively. Changing SCO2 has also a significant impact on QP−1. Increasing SCO2 from 0 to 60% leads to a 0.02 (1/49) and 0.01 (1/97) increase in QP−1 at 30 Hz for the Lincs and Forth sites, respectively. Table 3 implies that in practice (i.e. at field scale), greater changes in the P-wave velocity of the reservoir rocks may be observed after injecting CO2 into aquifers, compared with the values obtained in laboratory measurements. For the S-wave velocities, no significant difference between the field and laboratory measurements is expected. AVA response To distinguish the AVA characteristics of the Lincs and Forth sites we apply the classification method introduced by Rutherford & Williams (1989) that distinguishes three classes of

Sherwood Sandstone (Lincs reservoir)

SCO2

Pd (MPa)

VP (m s−1)

VS (m s−1)

0.0

6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 6.9 13.8 20.7 27.6 3.5 6.9 10.4 13.8 17.3 20.7 24.2 27.6

3766 3946 4007 4063 3838 3926 4028 4071 3731 3853 3917 3967 3660 3738 3845 3875 3549 3679 3782 3823 3704 3795 3857 3900 3436 3560 3701 3763 2562 2692 2733 2776 2797 2880 2917 2967

2101 2247 2303 2337 2144 2237 2296 2332 2133 2222 2294 2321 2158 2231 2299 2335 2122 2213 2283 2322 2227 2303 2356 2377 2155 2244 2328 2378 1501 1573 1601 1619 1637 1648 1664 1679

(%)

3.8

26.6

58.5

76.1

82.0

100.0

Merica Mudstone (Lincs caprock)

0.0

gas-sand AVA anomalies. According to their classification, an AVA anomaly represents Class 1 when the reflection coefficient of a normal-incidence P-wave is strongly positive, and the amplitude decreases with offset. A phase reversal might be expected at far offset for Class 1. In Class 2, the normal-incidence P-wave reflection coefficient is small (either positive or negative), and a large change in AVA is expected. In the case of CO2 injection, if the normal-incidence reflection coefficient is slightly positive, a phase reversal at near or moderate offsets is expected. Class 3 AVA anomalies present a large negative normal-incidence reflection coefficient. This anomaly becomes stronger with offset, which represents so-called bright spots. We calculate the AVA response of the interface between aquifer and caprock for the two storage sites at different reservoir CO2 saturations and corresponding differential pressures using the Zoeppritz equations. Figure 7 shows P-wave and converted P to S-wave AVA anomalies (PP and PS, respectively) for the Lincs and Forth sites. We convert the incident angles to ray parameters (horizontal slownesses) for consistency with later sections. Both sites clearly present Class 2 AVA anomalies; however, the gradient of the Lincs’ AVA anomaly is greater than that of the Forth’s AVA anomaly. Increasing SCO2 decreases the P-wave reflection coefficients for both sites towards negative reflection coefficients with a greater rate of decrease for the

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

364

Novel time-lapse AVA attributes

Fig. 5. Laboratory measured P- and S-wave velocities of samples SSK and CL1 (asterisks) at 6.9, 13.8, 20.7 and 27.6 MPa differential pressure and best-fit petrophysical models (solid lines) v. CO2 saturation. Table 2. Estimated representative material parameters of clay-bearing sandstone and the pore fluids it contains (adapted from Pham et al. 2002) Sand

Clay

Brine

CO2 Gas

Bulk modulus (Forth) Bulk modulus (Lincs) Shear modulus Density Average particle diameter Bulk modulus Shear modulus Density Average particle diameter Bulk modulus Density Viscosity Bulk modulus Density Viscosity Bulk modulus Density Viscosity

39 GPa 49 GPa 33 GPa 2650 km m−3 100 μm 25 GPa 9 GPa 2650 km m−3 2 μm 2.4 GPa 1030 kg m−3 1.798 cP 0.057 GPa 667 kg m−3 0.052 cP 0.01 GPa 100 kg m−3 0.02  cP

Lincs site. Also, greater separation between AVA anomalies at different CO2 saturations at the Lincs site is particularly valuable in order to detect the CO2 saturation by time-lapse monitoring. As shown in Table 3, injection of CO2 in the aquifers has little impact on the S-wave velocity of the reservoir rocks. Figure 7c, d show converted P- to S-wave AVA responses of the aquifer–caprock interfaces for both sites with different CO2 saturations. The magnitude of the converted S-wave coefficient for the Lincs site is larger than that for the Forth site. Regardless of the magnitude of the reflection coefficients, separation between the S-wave AVA anomalies for different saturations is very small. Analysis of AVA responses of aquifer–caprock interfaces using the Zoeppritz equations provides valuable information about the monitorability of changes in the aquifers. However, such analysis lacks an assessment of the impact of overall site geology (i.e. overburden/underburden effects) and frequency-dependent effects. The following subsection describes a strategy to incorporate those. Synthetic CMP gathers We now examine the detectability of changes by generating synthetic common mid-point (CMP) gathers, taking account of

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

365

A. JafarGandomi & A. Curtis

Fig. 6. Schematic demonstration of the impact of CO2 injection on (a) the P-wave velocity and (b) the S-wave velocity. The arrows show the trajectories of velocity change due to CO2-saturation increase (line A–B, dashed black), differential pressure increase (line B–C, blue in online version) and combined increase of both CO2 saturation and differential pressure (line A–C, red in online version). Table 3. Estimated changes in P- and S-wave velocities (m s−1) due to an increase of SCO2 from 0 to 60% for SSK (Lincs site) and CL1 (Forth site) sampled at 30 and 1 MHz frequency Sample CL1 (Forth)

SSK (Lincs)

Parameter (m s−1)

SCO2 =0%

VP VS ρ 1/QP VP VS ρ 1/QP

3914 2436 2284 1/60 3823 2220 2326 1/60

(30 Hz)

SCO2 =60% (30 Hz)

Change (30 Hz)

Change (1 MHz)

3766 2440 2252 1/37 3610 2206 2288 1/27

−147 4 −32 1/97 −214 −13 −38 1/49

−114 0 – – −183 −16 – –

overburden/underburden. We use the sonic-logs to construct subsurface velocity models for this purpose (Fig. 8). At the Forth site, we combine the sonic-logs obtained from two boreholes: the Firth of Forth-1 borehole, which is close to the potential injection location (Jin et al. 2010) with a total penetration depth of 2040 m; and the BGS Glenrothes borehole located approximately 15 km north of the potential injection location with a total penetration depth of 567 m. The Firth of Forth-1 borehole did not penetrate through the target caprock and aquifer formations (BGN and K&K formations, respectively). However, the BGS Glenrothes borehole penetrates both caprock and aquifer formations since, at the Glenrothes site, the formations are offset to shallow depths by a fault between the two boreholes (Brerton et al. 1988). We use the corresponding caprock and aquifer sonic-log segments from the BGS Glenrothes borehole, corrected for burial depth (Monaghan et al. 2009), at the Firth of Forth-1 borehole location. At the Lincs site, we chose the Slatfleetby-1 borehole with a total penetration depth of 2414 m near the potential injection location (Brerton et al. 1988;

Monaghan et al. 2009). Since we have no information about the S-wave velocity and the density, we use empirical relationships from Han et al. (1986) to estimate them from the sonic-logs: VS = 0.794VP − 849

ρ = 0.204VP + 1580.



(13)



(14)

In these equations, the units for P- and S-velocities are m s−1 and for density kg m−3. Attenuative implications of subsurface random heterogeneities are quite well understood among seismologists through the scattering effect (e.g. Sato & Fehler 1997). In exploration seismology, the scattering effect of these random heterogeneities is assumed to be negligible with respect to the target impedance contrasts (i.e. major layer boundaries). However, in the case of time-lapse monitoring, small changes in the seismic amplitudes are exactly the sought after signals, and such scattering effects must be taken into account. In the absence of 2D and 3D information about heterogeneities, the sonic well-log is used to quantify the expected impact of 1D random heterogeneities on the propagating seismic waves (Shiomi et al. 1997). The scattering attenuation for a seismic wave propagating through a 1D random subsurface model may be approximated by (Sato 1982): Q −1 ( k ) =

k P ( 2k ) 4

(15)

where P is the power spectral density (PSD) of the random fluctuations of velocity and k is the wavenumber. Figure 8c indicates the estimated scattering attenuation models for the Lincs and Forth sites using their P-wave velocity models (Fig. 8a, b,

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

366

Novel time-lapse AVA attributes

Fig. 7. P-wave AVA at different CO2 saturations predicted using the Zoeppritz equations for (a) the Lincs site and (b) the Forth site. Converted P- to S-wave AVA at different CO2 saturations for the Lincs and Forth sites are shown in (c) and (d), respectively.

respectively). The random functions for each site are obtained by removing the background velocity, which is a smoothed version of the well-log. The estimated attenuation models imply greater seismic energy (amplitude) loss for higher wavenumbers at the Lincs site than the Forth site at the same depth. The Forth site presents greater scattering attenuation for k<0.01 (large-scale heterogeneities). The scattering effect is implicitly included in the FDTD waveform modelling by introducing fine layering (JafarGandomi & Takenaka 2013). In addition to the scattering effect, incorporating the effect of seismic wave attenuation due to fluid flow in the reservoir porespace is also essential for realistic waveform modelling. Carcione (1998) showed that viscoelastic effective rheologies can be used to represent the poroelastic effect due to wave propagation in porous media. Picotti et al. (2010) indicated that the attenuation due to wave-induced fluid flow at the mesoscopicscale (e.g. Pride et al. 2003) can be represented with the Zener rheological model. We use a Generalized Zener Body (GZB) model (consisting of five relaxation times) to incorporate the induced attenuation due to the patchiness of the CO2–brine mixture into the FDTD modelling scheme. Figure 9a, b depicts the estimated Q−1 models from the hybrid White–Pham petrophysical model for patch sizes of 10 cm, and the best-fit GZB models

for the Lincs and Forth sites, respectively. As can be seen in Figure 9, the attenuation predicted from the petrophysical model and the GZB model are in a good agreement. Note that in this case we assume an intrinsic viscoelastic attenuation of 1/60 due to friction between solid grains. We consider an off-shore time-lapse monitoring situation in which the seismic survey is conducted before and after CO2 injection. We assume that injecting CO2 into the brine-saturated aquifer creates a 50 m-thick CO2-saturated zone at the top of both aquifers, with a SCO2 value of 60%. We calculate synthetic seismograms for incident P-waves with ray parameters of between 0 and 0.14 s km−1 to create CMP gathers. This range of ray parameter is equivalent to incident angle ranges of 0o–30o and 0o–40o for the reservoir–caprock interfaces at the Lincs and Forth sites, respectively. We use the velocity models obtained from the welllogs (Fig. 8) for calculation of the pre-injection gathers and modify the top 50 m of the reservoirs following the expected changes in geophysical parameters given in Table 3 to calculate the postinjection gathers. We assume a constant intrinsic (viscoelastic) attenuation of QP−1= QS−1= 1/60 for all layers at both sites, except for the CO2-saturated zone for which attenuation models in Figure 9 are used. A zero-phase Ricker wavelet with a central frequency of 30 Hz is used as the incident downgoing wave. The positions

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

367

A. JafarGandomi & A. Curtis

Fig. 8. P-wave velocity models for (a) the Lincs site, (b) the Forth site and (c) the corresponding estimated scattering attenuation models in the wavenumber domain.

Fig. 9. Attenuation model from the best-fit petrophysical model (circles) and GZB rheological model (solid line). The latter was used for finite-difference waveform modelling for (a) the Lincs site and (b) the Forth site.

of both the centre of the initial source wave and the receiver are 100 m above the seabed. The top boundary condition of the finitedifference grid is set as a non-reflecting boundary to avoid waterlayer multiples, and hence we assume that free-surface multiples have been removed from the seismic data prior to AVA analysis. Figure 10a, c shows the post-injection synthetic CMP gathers (pre-critical angle) in the τ−p (plane-wave) domain for the Lincs and Forth sites, respectively. The amplitudes of the synthetic seismograms show the divergence of the wavefield, and hence represent data that would be recorded on hydrophones. Corresponding differences between the pre- and post-injection gathers are also

shown in Figure 10b, d, respectively. Note that the amplitude of the difference records is amplified to improve the visibility. Using pairs consisting of pre- and post-injection CMPs for each site, we calculate the time-lapse attributes δ E and δθ in both time and frequency domains using the SP-based approach for the time–frequency decomposition. Figures 11 and 12 show the calculated time-lapse attributes for the Lincs and Forth CMPs, respectively. In general, negative values of δ E ( t ) for the reservoir–caprock interface for the Lincs site (Fig. 11a) are in agreement with the estimated AVA response using the Zoeppritz equations (Fig. 7a). However, that is not the case for the Forth

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

368

Novel time-lapse AVA attributes

Fig. 10. Post-injection synthetic CMP gathers in the τ−p (plane-wave) domain calculated for (a) the Lincs site and (c) the Forth site; corresponding differences between the post- and pre-injection gathers are shown in (b) and (d), respectively. Note that the amplitude of the difference records is amplified to improve visibility.

site. The low magnitude of δ E and δθ for both sites is due to normalization by the global maximum of the baseline records, which occurs at the earlier parts of the traces. Such normalization exposes the impact of local site geology including seismic energy loss due to scattering and the attenuative effects of geological formations. The Lincs CMP shows a strong negative δ E ( t ) anomaly (−3%) at the top of the aquifer (0.7–0.75 s) followed in line by a positive anomaly (0.75–0.8 s), which indicates the base of the CO2-saturated zone (Fig. 11a). The presence of the CO2-saturated zone at the top of the aquifer also affects the travel-time and amplitude of later events in the CMP gather (e.g. Ghaderi et al. 2010). δθ ( t ) values (Fig. 11b) show very small changes for the

top reservoir event except at far-offsets, which is as expected. The strong negative δθ ( t ) anomalies between 0.92 and 1.05 s are due to the increased travel-time of the interbed multiple reflections within the high-velocity evaporite-dominant formations underlying the aquifer (Fig. 8). We select the event in the timewindow shown with the dashed lines in Figure 11a to estimate δ E ( f ) and δθ ( f ). The largest decrease in amplitude occurs at 12–22 Hz, dominantly at 13 Hz and the larger ray parameters (far offsets). At the smaller ray parameters (near offsets), the negative amplitude change is centred around both 30 and 13 Hz, the former being the dominant frequency. The interference between the reverberations within the CO2-saturated zone caused the increased amplitude in 20–25 Hz frequency range.

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

A. JafarGandomi & A. Curtis

369

Fig. 11. Time-lapse attributes estimated for the Lincs site from comparison of pre- and post-injection CMP gathers. (a) Timeprojected envelope difference δE ( t ) , (b) time-projected phase difference δθ ( t ) , (c) frequency-projected envelope difference δE ( f ) and (d) frequency-projected phase difference δθ ( f ) . Plots (c) and (d) are calculated for the CMP waveforms that occur within the time-slowness window shown with dashed lines in (a).

Corresponding estimated time-lapse attributes for the Forth site are shown in Figure 12. Generally, the values of all time-lapse attributes at the Forth site are lower than those for the Lincs site because of its deeper position, which leads to greater seismic energy loss. This implies that the monitorability of the Forth site is poorer (monitoring using AVA will be more difficult) than at the Lincs site. The maximum values of δ E ( t ) and δθ ( t ) at the Forth site are −0.6 and 0.3%, respectively. Similar to the Lincs site, the negative δ E ( t ) associated with the top reservoir (1.15– 1.25 s) at the Forth site is followed by a positive δ E ( t ) anomaly caused by reflection from the base of the CO2-saturated zone. Lack of information from below the reservoir at the Forth site makes the later parts of δ E ( t ) and δθ ( t ) (Fig. 12a, b) artificially look cleaner than those for the Lincs site (Fig. 11a, b). There is a phase-reversal at far-offsets (ray parameter 0.1 s km−1) at the reservoir–caprock interface of the Forth site, which is not present on the calculated AVA response from the Zoeppritz relationships

(Fig. 7), showing how important it is to perform waveform modelling, including surrounding rock layers, in addition to predicting reflectivity at the top reservoir only. When the maximum amplitude of the baseline trace is used in the denominator of equations (4)–(9), the magnitude of estimated δ E and δθ , both in the frequency and time domains, is an indication of the detectability of changes in the subsurface parameters given the background geology. Such measures may be used to compare monitorability of different storage sites. However, local normalization of δ E and δθ , both in the frequency and time domains, could also be applied to quantify time-lapse changes of any individual reflection event in a seismic gather (e.g. Fig. 13). Repeatability In the previous subsection, time-lapse attributes were calculated for the noise-free signals that have exposed the impact of site

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

370

Novel time-lapse AVA attributes

Fig. 12. Time-lapse attributes estimated for the Forth site from comparison of pre- and post-injection CMP gathers. (a) Timeprojected envelope difference δE ( t ) , (b) time-projected phase difference δθ ( t ) , (c) frequency-projected envelope difference δE ( f ) and (d) frequency-projected phase difference δθ ( f ) . Plots (c) and (d) are calculated for the CMP waveforms that occur within the time-slowness window shown with dashed lines in (a).

geology on the AVA detectability. However, the AVA detectability may also be affected significantly by the signal-to-noise ratio. Here, we assess the repeatability between pre- and post-injection surveys at the two sites by adding random noises to the phase and amplitude of the pre- and post-injection synthetic records:

records at the Lincs and Forth sites (Fig. 13). For comparison, we also calculate corresponding NRMS (Kragh & Christie, 2002) expressed in percentage: NRMS =

S k * =α r|S k | eiα rθ



(16)

where Sk* and Sk are the modified and the original traces, subscript k=1,2 indicates pre- or post-injection records, respectively, r is the random noise sampled from a Gaussian distribution with zero mean and unit standard deviation, and α is a scale factor representing the strength of the random noise and is defined as a fraction of the global maximum of the pre-injection record. We calculate the single-valued time-lapse attributes (equations 10 and 11) with α values between 0.01 and 0.2 for the zero-offset

200 × RMS[ S 2 − S1 ]

RMS[ S1 ] + RMS[ S 2 ]

(17)

where RMS ( x ) =

tmax

∑x tmin

2 t

/N



(18)

and N is the number of samples in the interval tmax−tmin. Figure 13 shows the calculated single-valued δ E , δθ and

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

A. JafarGandomi & A. Curtis

371

Fig. 13. Detectability of reservoir changes against background noise with strength α. (a) Estimated single-valued δE and δθ (equations 10 and 11) and NRMS (equation 16) of the pre- and post-injection records against the strength of added random noise for the Lincs site and (b) for the Forth site.

NRMS with respect to the noise strength (α) for the Lincs and Forth sites. The larger than δ E and δθ values of NRMS is due to the cumulative effects of envelope and phase differences on the NRMS estimates. Variation of all three parameters of δ E , δθ and NRMS with respect to the noise strength α indicates a greater gradient for the Forth site than the Lincs site, which implies greater repeatability of time-lapse measurements for the Lincs site.

Discussion The risk of potential leakage associated with storage sites may be reduced through comprehensive monitoring. Assessing monitorability of a site contributes to the risk plan that is necessary for any storage site to be legal under current European Union legislation. Site detectability assessments based on the analysis of available data and estimated attributes indicate that the two sites considered here demonstrate distinctly different AVA detectabilities: the Lincs site shows greater changes in reservoir rock parameters due to supercritical CO2 injection, greater values of the new time-lapse attributes and, therefore, better detectability than the Forth site. Site-specific characteristics such as greater burial depth and greater velocity fluctuations (Fig. 8) in the overburden, as well as lower sensitivity of P- and S-wave velocities of the reservoir rock to CO2 saturation (Table 2), cause lower detectability of the Forth site. Although there are some agreements between the estimated AVA responses using the Zoeppritz equations (see the subsection on ‘AVA response’) and the synthetic CMPs (see the subsection on ‘Synthetic CMP gathers’) for the Lincs site, for both sites there are also some expected discrepancies between the results of these two approaches due to the incorporation of local site geology in the latter method. For example, while the Lincs AVA response predicted by the Zoeppritz equations shows a polarity reversal (Fig. 7a), by contrast there is no indication of such a reversal in the modelled wavefield in Figure 11a. The opposite contrast is observed for the Forth site. Such discrepancies highlight the importance of having knowledge of local site geology in the overburden/underburden when upscaling the AVA results obtained from the Zoeppritz equations. Ideally, the monitorability of each specific storage site would be assessed using a 3D volume, including the volumes above and below the reservoir. However, the quality of such assessments is subject to the reliability of detailed geological and geophysical models, which are usually not available at the earlier

stages of site selection. In addition, moving towards 2D and 3D analyses requires significantly increased computational resources and cost. Given these pros and cons, it seems reasonable that a method to assess site detectability based on 1D (depth-dependent) data, and the compilation of information from neighbouring wells and outcrop analogue such as the method described here, provides sufficient information for the initial monitorability assessment of sites from a short-list of potential storage sites. It is likely that porosity of the reservoir rocks varies after injecting CO2 into brine-saturated aquifers due to both chemical and mechanical effects. This in turn will affect the estimated Pand S-wave velocities. Local mechanical impacts on porosity at the pore-scale are included by introducing pressure into the analysis. Including chemical diagenetic effects, which might be considerable in the long term, would require that the petrophysical model included the impact of chemical reactions (e.g. Vanorio et al. 2008, 2010; Agersborg et al. 2011). The new attributes proposed in equations (4)–(9) are robust (e.g. Fig. 2) and more informative (frequency-dependent) than the commonly used cross-correlation or Taylor expansion methods (e.g. Zabihi Naeini et al. 2009) to estimate time-lapse changes on seismic records. These attributes might also be used to overcome the difficulties in detection of reservoir pore-pressure (e.g. Kvam & Landrø 2005), and estimation of thickness and velocity changes of injected CO2 layers (e.g. Ghaderi & Landrø 2009). Frequency-dependent attenuation (Chapman et al. 2006) is usually ignored during AVA analysis. However, in future, the new time-lapse attributes may offer a significant step towards quantitative assessment of fluid saturation (e.g. CO2, natural gas or oil) in reservoirs using low-frequency seismic methods. Figure 14 shows the impact of fluid-induced attenuation on the estimated attributes at different frequencies and times. In this figure we compare the estimated δ E ( t ) and δ E ( f ) of the zerooffset records at the Lincs and Forth sites with those estimated for viscoelastic conditions in the reservoir using a constant QP=QS=60 (i.e. neglecting fluid-flow effects on the seismic waves). In this case we normalize the envelope differences and the phase differences with the local maximum of the reflection events within a 4T wide time-window (where T is the dominant period of the incident wave) centred at the top-reservoir reflection. Figure 14 indicates that for both sites the incorporation of fluid-flow-induced attenuation decreases the expected amplitude changes. This is because the increased attenuation in the CO2saturated zone reduces the velocity and, hence, the impedance

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

372

Novel time-lapse AVA attributes

Fig. 14. Estimated amplitude changes of zero-offset records at the Lincs (a) and (c), and Forth (b) and (d) sites, for perfectly elastic (dashed line) and attenuative (solidcrossed line) conditions in the CO2-flooded zones. The top row shows δE ( t ) and the bottom row δE ( f ) . Bold intervals in (a) and (b) show the interval containing the reservoir.

contrasts at the reservoir–caprock interface. The greater change in δ E (equations 10 and 11) for the Forth site is due to the higher velocity of the reservoir rocks, which leads to a greater velocity reduction due to flow-induced attenuation. Ideally, timelapse attributes may be inverted to estimate saturation and thickness of the CO2-flooded zone. What is clear is that attenuation is certainly frequency-dependent and the measurement of this by using attribute δ E ( f ) may aid this inversion. Although, we use AVA data to assess the site monitorability, the same general approach may be applied to a variety of existing seismic and non-seismic geophysical methods (e.g. JafarGandomi & Curtis 2011a, 2012) and emerging geophysical methods that may be applicable to monitoring CO2 storage reservoirs. Such methods may use more exotic data types or may be used in more heterogeneous reservoir/overburden structures. For example, Zhou et al. (2010) used coda wave interferometry in a down-hole vertical seismic profile (VSP) setting to monitor possible leakage from CO2 stores. Khatiwada et al. (2008) showed that the same interferometry approach may be used to monitor CO2 stored in geologically complicated environments such as layered basalts. In each case, a workflow similar to that in Figure 1 can be created, replacing AVA by the relevant data type.

Conclusion We propose an approach to assess the AVA detectability of subsurface reservoirs as a key component of the overall site monitorability based on the results of laboratory measurements and the use of seismic amplitude variation with offset/angle (AVO/AVA) data. We developed the approach within the framework of assessing the detectability of injected supercritical

SCO2 , where CO2 is to be stored in saline aquifer reservoirs, as a part of the storage-site selection process. We show that, although laboratory measurements on the reservoir and caprock samples under different saturations and pressures provide valuable information, further site-specific data such as well-logs are necessary in order to represent local site geology and to have a reliable estimate of site monitorability. We introduce a new set of robust time-lapse attributes based on time–frequency decomposition to quantify the changes in the seismic records due to the changes in the reservoir fluids. We apply the proposed approach to assess the detectability of the CASSEM project analogue CO2 storage sites in the near-shore UK North Sea in the Firth of Forth region (Forth site) and the York– Lincolnshire region (Lincs site). This result highlights the importance of performing waveform modelling that includes local geological heterogeneities above and below the reservoir interval rather than using only petrophysical relations to predict reflection amplitudes from the top of the reservoir. The latter approach can give quite misleading results. Overall, expected time-lapse changes using the new time-lapse attributes for CO2-injection scenarios indicate more reliable timelapse detectability and repeatability for the Lincs site than for the Forth site. The research related to this paper has been carried out within the CASSEM project, which is a project supported by the Technology Strategy Board. The authors wish to acknowledge the support of the TSB and the EPSRC, and the project industry partners: AMEC, Marathon, Schlumberger, Scottish Power, and Scottish and Southern Energy, and the academic partners: British Geological Survey, Heriot-Watt University, University of Edinburgh, and the University of Manchester. We also thank Q. Fisher and his colleagues at the University of Leeds for providing the laboratory measurements.

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

A. JafarGandomi & A. Curtis Appendix Any signal in the time domain s(t) can be represented by its time–frequency representation W(t, f), which is calculated by a continuous wavelet transform (CWT): W (t, f ) =



∫ s(t )ψ

* a ,b

(t )dt

(A1) where ψ a ,b (t ) is a scaled (by a = ω0 / 2π f ) and translated (by t = b ) version of some function such as a Morlet wavelet that has zero amplitude at negative frequencies: −∞

ψ (t ) =

 t2  1 exp ( −iω0t ) exp  −  . 2π  2

(A2)

Here t and f represent time and frequency, respectively, the asterisk indicates complex conjugation, and the parameter ω0 > 5 allows a trade-off between time and frequency resolution (e.g. Vetterli & Kovacevic, 1995).

References Agersborg, R., Johansen, T.A., Mavko, G. & Vanorio, T. 2011. Modeling of elasticity effects of sandstone compaction using coated inclusions. Geophysics, 76, E69–E79. Aki, K. & Richards, P.G. 1980. Quantitative Seismology: Theory and Methods. W.H. Freeman, San Francisco, CA. Bachu, S. 2000. Sequestration of CO2 in geological media: Criteria and approach for site selection in response to climate change. Energy Conversion & Management, 41, 953–970. Bachu, S., Gunter, W.D. & Perkins, E.H. 1994. Aquifer disposal of CO2: Hydrodynamic and mineral trapping. Energy Conversion & Management, 35, 269–279. Birkholzer, J.T., Zhou, Q. & Tsang, C.F. 2009. Large-scale impact of CO2 storage in deep saline aquifers: A sensitivity study on pressure response in stratified systems. International Journal of Greenhouse Gas Control, 3, 181–194. Brerton, R., Browne, M.A.E., Cripps, A.C., Gebski, J.S., Bird, M., Halley, D.N. & McMillan, A.A. 1988. Glenrothes Borehole: Geological Well Completion Report, WJ/GE/88/2. Investigation of the Geothermal Potential of the UK. British Geological Survey, Keyworth, Nottingham. Carcione, J.M. 1998. Viscoelastic effective rheologies for modelling wave propagation in porous media. Geophysical Prospecting, 46, 249–270. Castagna, J.P. & Smith, S.W. 1994. Comparison of AVO indicators: A modeling study. Geophysics, 59, 1849–1855. Castagna, J.P., Swan, H.W. & Foster, D.J. 1998. Framework for AVO gradient and intercept interpretation. Geophysics, 63, 948–956. Chapman, M., Liu, E. & Li, X. 2006. The influence of fluid-saurated dispersion and attenuation on AVO analysis. Geophysical Journal International, 167, 89–105. Eiken, O., Haugen, G.U., Schonewille, M. & Duijndam, A. 2003. A proven method for acquiring highly repeatable towed streamer seismic data. Geophysics, 68, 1303–1309. Fisher, Q.J., Martin, J., Grattoni, C., Angus, D. & Guise, P. 2010. Ultrasonic Velocity and Electromagnetic Property Analysis of Sandstone Samples with Varying Brine and Supercritical CO2 Saturations. British Geological Survey, CASSEM Project Report. British Geological Survey, Keyworth, Nottingham. Fuck, R.F., Bakulin, A. & Tsvankin, I. 2009. Theory of traveltime shifts around compacting reservoirs: 3D solutions for heterogeneous anisotropic media. Geophysics, 74, D25–D36. Ghaderi, A. & Landrø, M. 2009. Estimation of thickness and velocity changes of injected carbon dioxide layers from prestack time-lapse seismic data. Geophysics, 74, O17–O28. Han, D., Nur, A. & Morgan, D. 1986. Effects of porosity and clay content on wave velocities in sandstones. Geophysics, 51, 2093–2107. Haszeldine, S. 2009. Carbon capture and storage: How green can black be? Science, 325, 1647–1651. Jacovitti, G. & Scarano, G. 1993. Discrete time techniques for time delay estimation. IEEE Transactions on Signal Processing, 41, 525–533. JafarGandomi, A. & Curtis, A. 2011a. Detectability of petrophysical properties of subsurface CO2-saturated aquifer reservoirs using surface geophysical methods. The Leading Edge, 30, 838–845.

373

JafarGandomi, A. & Curtis, A. 2011b. How do we know where CO2 is? Developments in geophysical monitoring. In: Smith, M., Campbell, D., Mackay, E. & Polson, D. (eds) CO2 Aquifer Storage Site Evaluation and Monitoring. Heriot-Watt University, Edinburgh, 103–119. JafarGandomi, A. & Curtis, A. 2012. Assessing the monitorability of CO2 saturation in subsurface aquifers. International Journal of Greenhouse Gas Control, 7, 244–260, http://dx.doi.org/10.1016/j.ijggc.2011.10.015 http://dx.doi.org/. JafarGandomi, A. & Takenaka, H. 2007. Efficient FDTD algorithm for plane-wave simulation for vertically heterogeneous attenuative media. Geophysics, 72, H43–H53. JafarGandomi, A. & Takenaka, H. 2013. FDTD3C - A FORTRAN program to model multi-component seismic waves for vertically heterogeneous attenuative media. Computers & Geosciences, 51, 314–323, http://dx.doi. org/10.1016/j.cageo.2012.07.022. Jin, M., Pickup, J., Mackay, E., Todd, A., Monaghan, A. & Naylor, M. 2010. Static and dynamic estimates of CO2 storage capacity in two saline formations in the UK. Paper SPE 131609, presented at the SPE EUROPEC/EAGE Annual Conference and Exhibition, 14–17 June 2010, Barcelona, Spain. Kennett, B.L.N. & Clarke, T.J. 1983. Seismic waves in a stratified half-spaceIV P-SV wave decoupling and surface wave dispersion. Geophysical Journal International, 72, 633–645. Khatiwada, M., van Wijk, K., Clement, W.P. & Haney, M. 2008. Numerical modeling of time-lapse monitoring of CO2 sequestration in a layered basalt reservoir. In: 78th Society of Exploration Geophysicists International Exposition and Annual Meeting, SEG Annual Meeting, November 9–14, 2008, Las Vegas, Nevada. Geophysical Expanded Abstracts. Society of Exploration Geophysicists, Tulsa, OK, 3189–3193. Knott, C.G. 1899. Reflexion and refraction of elastic waves with seismological applications. Philosophical Magazine, 48, 64–97. Kragh, E. & Christie, P. 2002. Seismic repeatability, normalized and rms, and predictability. The Leading Edge, 21, 640–647. Kristekova, M., Kristek, J., Moczo, P. & Day, S.M. 2006. Misfit criteria for quantitative comparison of seismograms. Bulletin of the Seismological Society of America, 96, 1836–1850. Kvam, Ø. & Landrø, M. 2005. Pore-pressure detection sensitivities tested with time-lapse seismic data. Geophysics, 70, O39–O50. Landrø, M. 1999. Repeatability issues of 3-D VSP data. Geophysics, 64, 1673–1679. Landrø, M. 2001. Discrimination between pressure and fluid saturation changes from time-lapse seismic data. Geophysics, 66, 836–844. Landrø, M. 2008. The effect of noise generated by previous shots on seismic reflection data. Geophysics, 73, Q9–Q17. Landrø, M. & Stammeijer, J. 2004. Quantitative estimation of compaction and velocity changes using 4D impedance and traveltime changes. Geophysics, 69, 949–957. Law, D.H.-S. & Bachu, S. 1996. Hydrogeological and numerical analysis of CO2 disposal in deep aquifers in the Alberta sedimentary basin. Energy Conversion & Management, 37, 1167–1174. Liu, L., Cao, S. & Wang, L. 2011. Poroelastic analysis of frequency-dependent amplitude-versus-offset variations. Geophysics, 76, C31–C40. Liu, Y. & Fomel, S. 2012. Seismic data analysis using local time-frequency decomposition. Geophysical Prospecting, 61, 516–525, http://dx.doi. org/10.1111/j.1365-2478.2012.1062.x. Lumley, D. 2001. Time-lapse seismic reservoir monitoring. Geophysics, 66, 50–53. MacKay, E., Hamilton, S. et al. 2011. Fate of CO2 in the store. Rock mechanics, geochemistry, and fluid flow. In: Smith, M., Campbell, D., Mackay, E. & Polson, D. (eds) CO2 Aquifer Storage Site Evaluation and Monitoring. Heriot-Watt University, Edinburgh, 69–101. Monaghan, A.A., McInroy, D.B., Browne, M.A.E. & Napier, B.R. 2009. CASSEM Work Package One – Forth Geological Modelling. BGS Commissioned Report, CR/08/151. British Geological Survey, Edinburgh. Pham, N.H., Carcione, J.M., Helle, H.B. & Ursin, B. 2002. Wave velocities and attenuation of shaley sandstones as a function of pore pressure and partial saturation. Geophysical Prospecting, 50, 615–627. Picotti, S., Carcione, J.M., Rubino, J.G., Santos, J.E. & Cavallini, F. 2010. A viscoelastic representation of wave attenuation in porous media. Computers & Geosciences, 36, 44–53. Pride, S.R., Berryman, J.G. & Harris, J.M. 2003. Seismic attenuation due to wave-induced flow. Journal of Geophysical Research, 109, B01201, http://dx.doi.org/10.1029/2003JB002639. Ringrose, R. & Simone, A. 2009. Site selection. In: Cooper, C. (ed.) A Technical Basis For Carbon Dioxide Storage. CO2 Capture Project. Chris Fowler International, London, 11–27.

Downloaded from http://pg.lyellcollection.org/ by guest on December 28, 2013

374

Novel time-lapse AVA attributes

Rutherford, S.R. & Williams, R.H. 1989. Amplitude-versus-offset variations in gas sands. Geophysics, 54, 680–688. Sato, H. 1982. Amplitude attenuation of impulsive waves in random media based on travel time corrected mean wave formalism. Journal of the Acoustical Society of America, 71, 559. Sato, H. & Fehler, M. 1997. Seismic Wave Propagation and Scattering in the Heterogeneous Earth. Springer-Verlag, New York. Scott, V., Gilfillan, S., Markusson, N., Chalmers, H. & Haszeldine, R.S. 2012. Last chance for carbon capture and storage. Nature Climate Change, 3,105–3,111. Shiomi, K., Sato, H. & Ohtake, M. 1997. Broad-band power-law spectra of well-log data in Japan. Geophysical Journal International, 130, 57–64. Skopintseva, L. & Stovas, A. 2010. Overburden dependent AVA inversion. Geophysics, 75, C15–C23. Smith, M., Campbell, D., Mackay, E. & Polson, D. (eds). 2011. CO2 Aquifer Storage Site Evaluation and Monitoring. Heriot-Watt University, Edinburgh. Smith, T.M., Sondergeld, C.H. & Rai, C.S. 2003. Gassmann fluid substitutions: A tutorial. Geophysics, 68, 430–440. Tura, A. & Lumley, D.E. 1999. Estimating pressure and saturation changes from time-lapse AVO data. In: 69th SEG Annual International Meeting, Offshore Technology Conference, 1–4 May 2000, Houston, Texas. Geophysical Expanded Abstracts. Society of Exploration Geophysicists, Tulsa, OK, 1655–1658. Ursenbach, C.P. & Bancroft, J.C. 2001. Playing with fire: Noise alignment in trim and residual statics. In: 71st SEG Annual International Meeting, September 9–14, 2001, San Antonio, Texas. Geophysical Expanded Abstracts. Society of Exploration Geophysicists, Tulsa, 1973–1976. van der Meer, L.G.H. 1993. The conditions limiting CO2 storage in aquifers. Energy Conversion & Management, 34, 959–966.

Vanorio, T., Scotellaro, C. & Mavko, G. 2008. The effect of chemical and physical processes on acoustic properties of carbonate rocks. The Leading Edge, 27, 1040–1048, http://dx.doi.org/10.1190/1.2967558. Vanorio, T., Mavko, G. & Vialle, S. & Spratt, K. 2010. The rock physics basis for 4D seismic monitoring of CO2 fate: Are we there yet? The Leading Edge, 29, 156–162. Veire, H.H., Borgos, H.G. & Landrø, M. 2006. Stochastic inversion of pressure and saturation changes from time-lapse AVO data. Geophysics, 71, C81–C92. Vetterli, M. & Kovacevic, J. 1995. Wavelets and Subband Coding. PrenticeHall, Englewood Cliffs, NJ. Vidal-Gilbert, S., Nauroy, J.F. & Brosse, E. 2009. 3D geomechanical modelling for CO2 geological storage in the Dogger carbonates of the Paris basin. International Journal of Greenhouse Gas Control, 3, 288–299. Wang, Y. 2010. Multichannel matching pursuit for seismic trace decomposition. Geophysics, 75, V61–V66. Waters, K.H. 1981. Reflection Seismology: A Tool for Energy Resource Exploration. Wiley, New York. White, C.M., Smith, D.H., Jones, K.L. et al. 2005. Sequestration of Carbon Dioxide in Coal with Enhanced Coalbed Methane Recovery – A Review. Energy Fuels, 19, 659–724. White, J.E. 1975. Computed seismic speeds and attenuation in rocks with partial gas saturation. Geophysics, 40, 224–232. Zabihi Naeini, E. Hoeber, H. Poole, G. & Siahkoohi, H. R. 2009. Simultaneous multivintage time-shift estimation. Geophysics, 74, V109–V121. Zhou, R., Huang, L., Rutledge, J.T., Fehler, M., Daley, T.M. & Majer, E.L. 2010. Coda-wave interferometry analysis of time-lapse VSP data for monitoring geological carbon sequestration. International Journal of Greenhouse Gas Control, 4, 679–686. Zoeppritz, K. 1919. Erdbebenwellen VIIIB, On the Reflection and Propagation of Seismic Waves. Göttinger Nachrichten, I, 66–84.

Received 20 June 2012; revised typescript accepted 16 June 2013.

time-lapse attributes storage sites using novel 2 ...

Oct 16, 2013 - to receive free e-mail alerts when new articles cite this article here click ... Downloaded from ... the three-dimensional (3D) distribution of CO2 during the injec- ..... Novel time-lapse AVA attributes. 360 δθ δθ. t f. t f. W t f. ,. , max.

3MB Sizes 1 Downloads 193 Views

Recommend Documents

Creating covariates using cohort attributes - GitHub
Mar 28, 2016 - 3.1 Creating the cohort attributes and attributes definitions . ... covariate builders that create covariates on the fly and can be re-used across ...

Attributes-Attitudes.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item.

A Novel Storage Scheme for Parallel Turbo Decoder
We do this by restricting the whole number of colors seen by a SISO processor when it decode the two component codes. If p χ can be restricted, the total tri-state buffer consumption will decrease. The resultant “reordered first fit” algorithm i

A novel material for hydrogen storage
A novel material for hydrogen storage. E. Titus*, Gil Cabral, J.C. Madaleno, M.C. Coelho, T. Shokuhfar and J.Gracio. Department of Mechanical Engineering, ...

Criteria interactions across visual attributes
temporally limited visual entity. On the other hand, it connects .... be judged; for the C# cases, the post-cue remained black during the whole block; ..... entities. Obviously, the relationship between the sensory and decisional aspects of distribut

Google​​Finance​​Attributes
"​marketcap​"​​-​​The​​market​​capitalization​​of​​the​​stock. ... "​shares​"​​-​​The​​number​​of​​outstanding​​shares.

A novel material for hydrogen storage
Hydrogen storage, Composites ... A novel hydrogen storage medium is highly .... Callejas MA. Chem Commun 40,. 1450, 2001. *Corresponding author: email:.

A Novel Scheme for Remote Data Storage - Dual Encryption - IJRIT
Abstract:- In recent years, cloud computing has become a major part of IT industry. It is envisioned as a next generation in It. every organizations and industries ...

A reordered first fit algorithm based novel storage ... - Springer Link
context, we call edges resulting from the 1st phase, thus drawn below the numbers, ..... Networks, SAMSUNG Electronics Co., 1999. [6] Crozier S, Guinand P. ... Turbo Decoding [C]// 9th International Conference on. Electronics, Circuits and ...

consolidating fragmented identity: attributes ...
Information Systems Institute, Faculty of Business and Economics, University ... Modern organizations become distributed and maintain multiple identity repositories. ... technology architects are asked to define end-to-end business processes ...

Moneyspots: Extraneous Attributes and the ...
the first subperiod, each agent has access to a linear production tech- nology that ... App. Cч, the main results of this paper generalize to coarser moneyspots, i.e., ...... economies because of legal restrictions that prevent financial companies.

A Novel Scheme for Remote Data Storage - Dual Encryption - IJRIT
stored in the cloud. By using the corresponding private key, the embedded data and the key can be extracted successfully from the cloud. This scheme ensures ...

Coordination of V2G and Distributed Wind Power Using the Storage ...
Coordination of V2G and Distributed Wind Power Using the Storage-like Aggregate PEV Model.pdf. Coordination of V2G and Distributed Wind Power Using the ...

Performance Testing of Object-Based Block Storage Using ... - Media15
extending beyond business files to en- compass photographs, audio, video, and logs. Along with the associated transition to big data, it has been projected that.

Privacy Preserving Public Auditing for Secure Cloud Storage Using TPA
the task of allowing a third party auditor (TPA), on behalf of the cloud client, to verify the integrity of the dynamic data stored in the cloud. To securely introduce an ...

4 Rational Function Attributes Test review.pdf
4 Rational Function Attributes Test review.pdf. 4 Rational Function Attributes Test review.pdf. Open. Extract. Open with. Sign In. Main menu.

storage protein-2 as a dependable biochemical index ...
Differential Expression of SP2, their ranking and pupation rate in the multivoltine .... Kyushu University 1968, 23: 205-214. 5. ... Los Altos, California 1993, pp.31-.

Emailing Roleplays 2 - Using English
Roleplay email exchanges from below, saying exactly what you would write in your email. Continue each ... My name is… and I work for… I was given your ...

Agile manufacturing: The drivers, concepts and attributes
customer choice and expectation, competitive pri- .... response to customer demand (customer needs and .... support to include the provision of access to en-.

Qualification Positive Attributes of Independence & Performance ...
Qualification Positive Attributes of Independence & Performance Evaluation-2016.02.13.pdf. Qualification Positive Attributes of Independence & Performance ...