You are on page 1of 10

GEOPHYSICS, VOL. 80, NO. 6 (NOVEMBER-DECEMBER 2015); P. B167–B176, 8 FIGS., 2 TABLES.

10.1190/GEO2014-0558.1
Downloaded 10/02/15 to 132.203.227.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Case History

Ambient seismic noise tomography at Ekofisk

Sjoerd A. L. de Ridder1 and Biondo L. Biondi2

ABSTRACT sources. These sources were dominated by Scholte waves trav-


eling along the seafloor, and we sought to determine which features
We have studied an approximately 40-hr recording from the in the subsurface can be imaged by straight-ray group-velocity
Life of Field Seismic array installed over Ekofisk field to assess tomography. We located a high-velocity anomaly in the center
whether the recorded ambient seismic energy was suitable for of Ekofisk’s production-induced subsidence bowl, surrounded
passive seismic surface-wave interferometry. Passive seismic by lower velocities. This pattern seemed to result from produc-
interferometry aims to retrieve virtual seismic sources by cross- tion-induced seafloor subsidence that altered the near-surface
correlation. We have found that the noise recorded by the pressure shear strengths. A dispersion analysis showed that the Scholte-
sensors between 0.4 and 1.4 Hz consisted predominantly of wave virtual seismic source exhibited an approximate penetration
Scholte-wave microseism energy. The noise incidence directions depth to 600 m below the seafloor. These results are significant
have an almost uniform distribution over the azimuth, enabling because they demonstrated that recordings made at the ocean-
the synthesis of nearly symmetric estimated Green’s functions bottom cable array at Ekofisk field in the absence of seismic shoot-
between all stations in the array that formed virtual seismic ing can be used to image and monitor the near surface.

INTRODUCTION A theory named passive seismic interferometry uses so-called


noise recordings of the ambient seismic field to turn seismic stations
The Ekofisk field in the southern part of the Norwegian North
into virtual seismic sources by crosscorrelation. Aki (1957) first
Sea is one of the first developed oil-producing fields in the North
derives (from a modal formulation) how to retrieve the dispersion
Sea (Hermansen et al., 1997). Rapid pressure depletion in the early
of surface waves from the crosscorrelations of a circle of stations.
phase of production and weakening due to subsequent water injec-
Claerbout (1968) concludes that the 1D autocorrelation of transmis-
tion caused more than 9 m of seafloor subsidence over the Ekofisk
sion responses would yield the reflection response (using 1D reci-
field (Hermansen et al., 1997; Lyngnes et al., 2013). Since 2010,
procity theorems) and conjectures an extension to three dimensions
Ekofisk field has had a Life of Field Seismic (LoFS) 4C optical
by crosscorrelations. This was formally proven for the elastody-
sensor array of ocean-bottom cables (OBCs) buried approximately namic case by Lobkis and Weaver (2001) and Weaver and Lobkis
2 m into the sand of the seafloor (Eriksrud, 2010). The main ob- (2002) based on normal mode expansions and by Wapenaar (2003,
jective of this installation is to record production-related time-lapse 2004) and Wapenaar and Fokkema (2006) using 3D reciprocity
surveying during active seismic surveying (Folstad et al., 2010). theorems.
Because the array is permanently installed and capable of recording Baskir and Weller (1975), Scherbaum (1987), Cole (1995), and
continuously, we study the potential of imaging using the ambient Daneshvar et al. (1995) are among the first to test Claerbout’s re-
seismic noise recorded in the absence of seismic shooting. lationship between the reflection response and the autocorrelation

Manuscript received by the Editor 26 November 2014; revised manuscript received 7 May 2015; published online 23 September 2015.
1
Formerly Stanford University, Department of Geophysics, Stanford, California, USA; presently University of Edinburgh, School of GeoSciences, Edinburgh,
UK. E-mail: s.deridder@ed.ac.uk.
2
Stanford University, Department of Geophysics, Stanford, California, USA. E-mail: biondo@stanford.edu.
© 2015 Society of Exploration Geophysicists. All rights reserved.

B167
B168 De Ridder and Biondi

of the transmission response at an exploration scale. Breakthrough has yielded images of the near surface using the Scholte-wave group
success came with the application in crustal-scale seismology start- and phase velocities between 0.5 and 1.75 Hz (De Ridder and Del-
ing with Campillo and Paul (2003) and Shapiro and Campillo linger, 2011; De Ridder and Biondi, 2013; Mordret et al., 2013a,
(2004). These successes led to a resurgence of research in explora- 2013b, 2013c). The question arises whether the successes at Valhall
tion-scale attempts (Artman, 2007; Draganov, 2007; Draganov et al., can be replicated at other permanent OBC installations.
The studies at Valhall and Astero use energy from the double-
Downloaded 10/02/15 to 132.203.227.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

2009; Edme and Halliday, 2011; Ruigrok et al., 2011; Ruigrok and
Wapenaar, 2012). Notably, Stewart (2006) retrieve the Scholte frequency microseism band excited by ocean swell noise (Longuet-
waves from an ocean-bottom cable (OBC) recording of microseism Higgins, 1950). In marine environments, this energy is dominated by
noise in the Gulf of Mexico. Other applications are found in solar Scholte waves (Olofsson, 2010). Over land, this energy peak is com-
physics (Duvall et al., 1993; Rickett and Claerbout, 1999) and in posed of surface and body waves incident from coasts and shallow
laboratory acoustics (Weaver and Lobkis, 2001, 2002). seas (Ruigrok et al., 2011). Permanently deployed OBCs could be
A study of an ocean-bottom-node recording over the Astero field used to monitor the near surface for time-lapse changes using ambi-
yielded images of group velocities at frequencies between 0.18 and ent seismic noise (De Ridder et al., 2014; Mordret et al., 2014).
0.4 Hz, with correlation to known structures (Bussat and Kugler, The near surface is an imaging challenge (Butler, 2005). Knowl-
2011). Permanently installed OBC ambient noise over Valhall field edge of the near surface helps to resolve wavefield imaging challenges
at depth that are caused by near-surface transmission complexities
(Cox et al., 1999). Recently, the applied seismology community has
taken an interest in the use of surface waves for imaging of the near
surface because Scholte waves have proven to be very sensitive to
stresses in the near surface and respond to near-surface geomechan-
ical changes induced by hydrocarbon production at depth (Wills
et al., 2008; Hatchell et al., 2009). Accurate knowledge of the near
surface may also help scientists to better understand geohazards in-
volved with hydrocarbon exploration and production (Landrø and
Amundsen, 2011; Barkved, 2012).
We present a study of almost 40 hr of data recorded by the pres-
sure sensors of the LoFS at Ekofisk field. First, we analyze the noise
characteristics with special attention to the microseism noise. Then,
we crosscorrelate the microseism noise and judge the quality of the
retrieved virtual seismic Scholte-wave sources. Finally, we image
Scholte-wave group-velocities for five central frequency ranges,
from 0.4 to 1.4 Hz, by straight-ray tomography.

Figure 1. Map of station locations in Ekofisk’s LoFS array (coor- AMBIENT SEISMIC FIELD RECORDED BY LoFS
dinates in ED50/universal transverse mercator[UTM] Zone 31N). AT EKOFISK FIELD
Each black dot denotes a station. The stations used to create the
spectrogram in Figure 2 and for the beam steering results in Figure 3 In this section, we analyze the characteristics of the microseism
are shown in blue. noise recorded by the pressure sensors of Eko-
fisk’s LoFS array. Ekofisk’s LoFS array uses
fiber-optic technology (Eriksrud, 2010). The
pressure sensors consist of a fiber wrapped
around a mandrel. The mandrel responds to
pressure changes lengthening the fiber, which
is measured by an interferometer (Amundsen
and Landrø, 2009). The exact response function
was not known to the authors, but the phase re-
sponse cancels by a crosscorrelation. Figure 1
contains a map with the LoFS station locations.
The inline and crossline station spacings are ap-
proximately 50 and 300 m, respectively. The ar-
ray comprises 3966 stations. For this study, we
obtained a data set spanning almost 43.5 hr, start-
ing on 24 October 2011, at Coordinated Univer-
sal Time (UTC) 00:17:00. There are three gaps in
Figure 2. Spectrogram showing spectral amplitudes versus time for the duration of the the recording lasting a combined total of approx-
entire recording used in this study. Dashed lines indicate the frequency regimes in which imately 3.5 hr.
swell noise, microseism noise, and operational noise dominate. Note that the microseism The transient nature of the ambient seismic
energy and swell-noise energy grow stronger during 24 October 2011 and remained strong
during 25 October. Blue and red curves display measurements made by a weather obser- field is observed in the spectrogram shown in
vation station at Ekofisk field (Weather Underground Inc., 2013). The blue curve shows Figure 2. The spectrogram shows how the time-
barometric pressure at sea level (in bar). The red curve shows wind speed (in kmph). frequency spectrum of the recorded seismic
Seismic noise tomography at Ekofisk B169

energy changes as a function of the recording time. This spectro- et al. (2013a) observe the acoustically guided mode in Valhall OBC
gram is computed using the Fourier transformation of 2.5-min recordings, but at greater than 10 Hz.
recording windows with 50% overlap. The spectrogram was aver-
aged over a subset of 119 stations located in a radius of 750 m of
PASSIVE SEISMIC INTERFEROMETRY
UTM (514 km east, 6261 km north), depicted as blue stations in
Passive seismic interferometry refers to the creation of virtual
Downloaded 10/02/15 to 132.203.227.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 1.
Analysis of Figure 2 identifies three frequency regimes. Below seismic sources, i.e., retrieval of an estimated Green’s function
0.15 Hz, the recordings are dominated by ocean swells. These are (EGF), by crosscorrelating noise recordings made at different sta-
wind-generated gravity waves in the sea (Munk, 1950). Between 0.35 tions. Ekofisk’s LoFS array has 3966 stations that record particle
and 1.35 Hz, microseism energy dominates the recordings. Micro- velocity and pressure approximately 2 m below the seafloor. Here,
seism noise is caused by constructively interfering ocean swells that crosscorrelations between the pressure sensor recordings made at all
give rise to pressure variations on the seafloor. These pressure var- stations were computed. Let Gp̄;q̄ ðxA ; xB ; ωÞ denote the frequency-
iations excite interface waves traveling along the seafloor, Scholte domain elastodynamic Green’s function of a composite-receiver re-
and Love waves, at approximately twice the frequency of the swell cording pressure p̄ at xA due to a composite source representing a
noise (Longuet-Higgins, 1950). Notice that the microseism noise in- volume injection rate q̄ at xB . An estimate of the Green’s function
creases when the ocean swells grow stronger. Above 1.5 Hz, a variety Gp̄;q̄ ðxA ; xB ; ωÞ and its reciprocal Gp̄;q̄ ðxB ; xA ; ωÞ can be retrieved
of predominantly field-operational noise sources dominate the re- as follows (Wapenaar and Fokkema, 2006):
cordings. This study focuses on the microseism noise because
Scholte waves travel along the seafloor and provide subsurface sen- hp̄ðxA ; ωÞp̄ ðxB ; ωÞi ∝ fGp̄;q̄ ðxA ; xB ; ωÞ
sitivity to the top few hundred meters immediately below the seafloor. þ Gp̄;q̄ ðxB ; xA ; ωÞgSðωÞ; (1)
The spectrogram in Figure 2 reveals that the swell noise and
swell-noise-generated microseism energy grew stronger during where the crosscorrelated signals p̄ðxA ; ωÞ and p̄ðxB ; ωÞ denote the
24 October and remained strong on 25 October. We downloaded pressure recordings made at xA and xB (the master station), respec-
observations from a weather station at Ekofisk field (Weather tively. Complex conjugation is denoted by * and corresponds to a
Underground, Inc. 2013), and we display the recorded barometric time reversal in the time domain, and hi denotes a spatial ensemble
pressure at sea level and wind speeds in Figure 2. During 24 and average. The power spectrum of the noise source signals is denoted
25 October, we see a gradual decrease in pressure, indicating the by SðωÞ. For a direct noise crosscorrelation to yield the Green’s func-
emergence or arrival of a low-pressure system, which is indicative tion, we require the ambient seismic field to be in a state of energy
of rougher weather conditions. This observation is corroborated by equipartitioning (Lobkis and Weaver, 2001). However, that state is
the increase in wind strengths during the 24th, and wind generally not sufficient, we also require all noise sources to act incoherently
remains strong during the 25th. This correlation between weather (Snieder et al., 2010).
conditions and microseism energy levels at Ekofisk field is not guar- Under ideal circumstances, the crosscorrelation signal yields an
anteed because sea-swell-excited microseism energy could be gen- equivalence of the superposition of the causal Green’s function
erated quite far from Ekofisk field. and its anticausal reciprocal counterpart, forming a purely symmetric
Beam steering is one technique used to reveal the azimuth and signal. An estimate for the phase of the Green’s function can be found
apparent slowness of energy incident on an array (Cole, 1995; Rost by applying the Heaviside step function to the crosscorrelation signal
and Thomas, 2002). Beam steering is carried out using the same 119 either before or after symmetrizing (Wapenaar and Fokkema, 2004).
stations used for the spectrogram (the blue stations in Figure 1). The The asymmetric part of the crosscorrelation provides a measure of the
recordings are first filtered for the microseism energy between 0.55 deviation from idealized ambient seismic noise characteristics. The
and 0.65 Hz using a Hann taper in the frequency domain. A narrow- asymmetric part of the crosscorrelation signal is the difference be-
frequency band was needed because velocity dispersion causes tween the causal and anticausal Green’s functions, which are re-
blurring in the beam steering results. The data are then transformed trieved from waves traveling in opposite directions between the
to the τ‐p domain by a slant stack. The absolute value is smoothed two stations (Snieder, 2004; Bensen et al., 2007). As such, the asym-
by a triangle over 10 min, forming a smooth movie of amplitudes metry in the crosscorrelation also provides an indication of the noise
over px and py as a function of τ. Figure 3 contains 15 frames of source distribution around the station pair.
the movie formed by averaging 2.5 min at 3-hr intervals. High am-
plitudes (white) indicate the slowness and azimuth of incoming
VIRTUAL SOURCES FROM EKOFISK’S
energy, whereas low amplitudes (black) indicate the absence of en-
LoFS ARRAY
ergy. We observe a circle at absolute slowness jsj ≈ 1.9 ms∕m, cor-
responding to a velocity of jsj−1 ≈ 525 m∕s. These are the Scholte To extract the microseism noise and compress the data volume,
waves, which travel in many directions and compose the micro- the recorded pressure data were first filtered using a frequency-do-
seism noise of the ambient seismic field. Notice that the noise is main taper with a flat response for 0.4–1.3 Hz, and a Hann taper
remarkably omnidirectional. Some energy at the center of the plots, extending from 0.35 to 1.35 Hz. Filtering was done in segments
especially visible at later times, indicates that there is a small of 3 min, plus 20 s of overlap, and the data were restitched (by aver-
amount of energy arriving with a very high apparent velocity. One aging in the overlap portion) after filtering and downsampling. Seg-
possible explanation of this energy may be that it represents a small ments containing noise bursts and spikes were discarded by tapering
fraction of body-wave energy recorded by the array. Bussat and and substituting zeros. All available data were then crosscorrelated
Kugler (2011) observe an acoustically guided wave (P-wave) in 10 blocks of 4 hr. For each station pair, the crosscorrelations were
frequencies as low as 0.6 Hz, but in much deeper waters. Mordret stacked to form a virtual seismic survey with virtual sources at all
B170 De Ridder and Biondi

stations in Ekofisk’s LoFS array. Figure 4 shows an example of a persive wave mode; i.e., the wave-speed varies with frequency. A sum-
virtual seismic source. Figure 4a to 4e contains the symmetric part mary of observed phase velocities as a function of frequency and their
of the crosscorrelation signal, ℜfp̄ðxA ; ωÞp̄ ðxB ; ωÞg, whereas Fig- corresponding wavelengths is shown in Table 1. This table tells us that
ure 4f to 4j contains the antisymmetric part ℑfp̄ðxA ; ωÞp̄ ðxB ; ωÞg. the microseism energy is not aliased in the inline direction, but it be-
At later time snaps, the lower frequencies precede the arrival of the comes aliased in the crossline direction greater than 0.75 Hz.
higher frequencies due to dispersion. The Scholte waves emitted by the virtual seismic sources propa-
Downloaded 10/02/15 to 132.203.227.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

There is very little coherent energy in the antisymmetric part, and gate along the seafloor with wavelengths approximately between
the background correlation fluctuations in the antisymmetric part 1460 m at 0.4 Hz and 340 m at 1.2 Hz (from Figure 5b). These
are on the order of the background correlation fluctuations in the long wavelengths provide sensitivity away from the seabed,
symmetric part. The beam-steering results indicate that the noise i.e., in depth. Lower frequencies will be more sensitive to deeper
incidence directions, especially when averaged over a long time, depths.
have uniform distribution over the azimuth. But the lack of (almost)
any antisymmetric part in the EGFs shows there are few (to none) STRAIGHT-RAY TOMOGRAPHY FOR
dominating and coherently acting noise sources. Thus, the micro-
GROUP-VELOCITY MAPS
seism noise at Ekofisk field is very suitable for creating Scholte-
wave virtual seismic sources by seismic interferometry. We expect Surface waves extracted using passive seismic interferometry are
that crosscorrelating and stacking of more recorded data (had it been traditionally studied to analyze their phase and group velocity (Bar-
available) would increase the signal-to-noise-ratio. min et al., 2001; Sabra et al., 2005; Shapiro et al., 2005; Gerstoft
A dispersion image is formed by taking the amplitude in the Ra- et al., 2006; Yao et al., 2006). These studies approximate surface
don domain. We sorted all the EGFs as a function of absolute offset waves as waves propagating in two dimensions through maps of
(Figure 5a) and then performed a slant stack to the τ‐p domain. apparent velocity (Wielandt, 1993). Here, we will invert the
Next, the data are transformed from the τ‐p domain to the ω‐p do- Scholte-wave seismic sources retrieved in the previous section
main by Fourier transformation. The amplitudes are balanced over by straight-ray group-velocity tomography (Barmin et al., 2001).
frequencies for clarity to produce Figure 5b. To this purpose, we picked group traveltimes of the surface waves
In Figure 5a, we observe that the crosscorrelations are dominated by for a series of frequencies, by picking the peak of the envelope of
a dispersive wave arrival interpreted to be an interface wave, specifi- the symmetrized EGFs after band-pass filtering (Levshin et al.,
cally, Scholte waves (Scholte, 1942a, 1942b). The fundamental 1989; Bensen et al., 2007). Our band-passes had a flat response over
Scholte-wave mode is the only mode visible in Figure 5b. It is a dis- an 0.2-Hz interval and extended as Hann tapers over an additional

Figure 3. Results from beam steering of data filtered between 0.55 and 0.65 Hz, selecting microseism noise. Each image contains a beam-
steering result from data 3 hr apart and indicates the slowness and azimuth of incoming Scholte-wave energy averaged over 10 min.
Seismic noise tomography at Ekofisk B171

0.2 Hz on either side. These are wider band-passes than the more ments (Shapiro and Singh, 1999). We anticipate this to be small
conventionally used Gaussian tapers, resulting in measurements addi- given the relatively constant spectrum within the microseism band
tionally averaged over frequency but more stable. We use the ratio (Figure 2).
between the maximum of the envelopewithinanestimated linear move- We pose a regularized straight-ray inverse problem; see, for ex-
out window to the average of the envelope outside the window as an ample, Barmin et al. (2001). First, we subtract the contribution of
the average slowness from the data. The average slowness m0 is
Downloaded 10/02/15 to 132.203.227.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

signal-to-noise-ratio quality factor. The spectrum of the ambient


noise and the neglected station response function may bias the measure- computed using the traveltime picks t ¼ ðt1 ; t2 ; : : : ; tN ÞT :

a) b) c) d) e)

f) g) h) i) j)

Figure 4. (a-e) Snap shots for symmetric and (f-j) antisymmetric parts of the EGFs. Correlation lags are (a and f) 0 s, (b and g) 4 s, (c and h) 8 s,
(d and i) 12 s, and (e and j) 16 s.

a) b)

Figure 5. (a) Offset gather for crosscorrelations between all station pairs with midpoints within UTM (513–514 km east, 6269–6270 km
north). (b) Dispersion image generated by transforming the gather in (a) to the ω‐p domain and balancing the amplitude over frequencies.
B172 De Ridder and Biondi

1X N
ti and by 140 (north–south) grid cells, 100 by 100 m wide. The grid
m0 ¼ : (2) cell size is below the resolution of the wavelength. Approximately
N i¼1 Δxi
half the model space is covered by the array, so we only have ap-
proximately 6300 model parameters to constrain. This means that
The contribution of the average slowness to the traveltime measure- the inverse problem is overdetermined and we can discard a signifi-
ments is subtracted from the traveltimes, yielding traveltime resid-
Downloaded 10/02/15 to 132.203.227.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

cant portion of our data.


uals Δt ¼ ðΔt1 ; Δt2 ; : : : ; ΔtN ÞT : To omit the poorer quality data, we select traveltime picks with an
signal-to-noise-ratio of more than two. To avoid station couples for
Δt ¼ t − m0 Δx; (3)
which the causal and anticausal Green’s functions are not well-
where Δx ¼ ðΔx1 ; Δx2 ; : : : ; ΔxN ÞT is a vector containing the off- separated, we only selected traveltime picks for interstation offsets
sets Δx for each specific traveltime pick. The traveltime residuals exceeding 1500 m (Bensen et al., 2007). This is smaller than the
are predicted from a perturbation slowness model Δm by a straight- ideal three wavelengths, especially for the lowest of frequencies.
ray tomography operator F as But the crosscorrelation quality and limited size of the array do
not allow for a larger exclusion zone. We took two measures to dis-
Δt ¼ FΔm: (4) card outlier data. First, we discard the traveltime perturbations falling
under the 5th and greater than the 95th percentiles. Second, we solve
This operator is simply a slowness integration kernel, in which each each tomography twice. The first time with a regularization strength ϵ
row contains the distances through each model cell of a straight line three orders of magnitude larger than the regularization strength that
between a specific source-receiver couple. A conjugate-direction we finally determined to be appropriate. Then, we discard the trav-
algorithm is used to find the minimum of the following misfit eltime picks with misfits greater than the 95th percentile, and we
function: solve the inverse problem again.
This whole procedure is repeated for five frequency ranges:
JðΔm; ϵÞ ¼ kFΔm − Δtk22 þ ϵk∇2 Δmk22 ; (5) 0.4–0.6, 0.6–0.8, 0.8–1.0, 1.0–1.2, and 1.2–1.4 Hz. In Table 2,
the number of selected traveltime picks and their corresponding
2 2 2
where kk2 denotes an L norm. We use the ∇ operator as a regu- average velocities m−1 0 are shown for each frequency range.
larization to force a smooth model (a model styling objective), also The upper frequency range, 1.2–1.4 Hz, is pushing it a little over
known as second-order Tikhonov regularization (Aster et al., 2005). the range for which we have well-formed Scholte waves, deter-
After inversion, the slowness map can be recovered from the per- mined from the dispersion analysis in the previous section. This
turbation map by adding the average slowness m ¼ Δm þ m0 . is also reflected by the significant decrease in accepted picks for
There are several strategies to determine regularization strength. this frequency range.
Usually, the regularization strength ϵ is picked as a value at which Figure 6 contains the five maps found by solving the inverse
we find an optimal trade-off between the data-fitting objective and problem. We observe a high-velocity anomaly in the center of the
the model styling objectives (Aster et al., 2005). Here, we found that array, surrounded by a lower velocity region. The high-velocity
a small regularization strength resulted in a strong acquisition im- anomaly coincides with the center of the seafloor subsidence bowl.
print appearing in the inverted model. We choose a regularization Under the southern end of the array, where the magnitude of the
strength that minimizes this acquisition footprint, and our strategy is gradient of the seafloor is smaller, we find higher velocities again.
further explained below in the resolution analysis. Ekofisk’s LoFS has a dense concentration of stations resulting in
The LoFS array has 3966 stations providing for almost 7.9 mil- a very good ray coverage and measurements throughout the domain
lion unique raypaths. The model space is formed by 90 (east–west) of imaging (Figure 7a). Figure 6c is the ray coverage for all trav-
eltime picks used for the group-velocity tomog-
raphy between 0.8 and 1.0 Hz.
One measure for resolution is a checkerboard
Table 1. Measurements of phase velocity and wavelength as a function of
frequency by picking maxima in Figure 5b. test (Lévěque et al., 1993). For a known model,
usually a checker pattern, computing the perfect
traveltime data and running the inversion give an
Frequency (Hz) 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 indication of which anomalies could potentially
Phase velocity (m∕s) 584.1 565.6 525.2 490.2 466.4 448.0 432.5 416.7 407.2 be recovered given the ray coverage. We attempt
to recover the checkerboard grid pattern shown
Wavelength (m) 1460 1131 875.3 700.3 583.0 497.8 432.5 378.8 339.3
in Figure 7b given the raypath coverage and
the regularization strength for the group-velocity
tomography between 0.8 and 1.0 Hz. The differ-
Table 2. Number of traveltime picks used per inversion and their corresponding ence between causal and anticausal traveltime
average velocity. picks, a proxy for the noise in the data, was
added to the computed traveltime in the true
model. The edges of the recovered checkerboard
Frequency range (Hz) 0.4–0.6 0.6–0.8 0.8–1.0 1.0–1.2 1.2–1.4 grid cells are smoothed by the regularization
Number of traveltime picks 1212778 5528579 5434445 3579255 552305 (Figure 7c). The squares are not always well re-
Average group velocity (m∕s) 378.8 349.4 328.0 311.3 301.2 solved at the edges of the array due to biased azi-
muthal coverage of the traveltime picks.
Seismic noise tomography at Ekofisk B173

The optimization was run by a conjugate directions solver. Soft- ing experiments on four recordings made during October, Decem-
ware with this aim typically require a forward and adjoint operator ber, January, and February at Valhall field (Mordret et al., 2013a; De
(Claerbout and Fomel, 2014), and our implementation was no ex- Ridder, 2014). Thus, we suspect that this may be representative of
ception. The forward operator projects the model space to the data microseism-energy conditions in the North Sea during fall and win-
space, and the adjoint operator projects the data space to the model ter, and potentially year around.
space. This allows us to see the effect of the chosen regularization Omnidirectional microseismic noise translates into omnidirec-
Downloaded 10/02/15 to 132.203.227.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

parameter by projecting the final data misfit with the adjoint of tional virtual seismic sources (Snieder, 2004). The resulting
the tomographic operator (Figure 7d) (Claerbout and Fomel, virtual seismic sources at Ekofisk field, where no balancing
2014). Most of the energy in the residual is an imprint of the acquis- was applied before stacking crosscorrelations from different time
ition geometry. There is also some energy that corresponds to windows, are almost perfectly symmetric. There is almost no
geologically reasonable features. However, decreasing the regulariza- observable antisymmetric part to the crosscorrelation stacks,
tion strength would make the acquisition imprint appear in the and the background correlation fluctuations in the antisymmetric
inverted model. part of the crosscorrelations are of the same strength as in the sym-
metric part. These background fluctuations will diminish when
DISCUSSION crosscorrelating and stacking more ambient seismic recordings
(Bensen et al., 2007), provided they are made under similar
We find that to the first order, microseism noise is equally inci- conditions.
dent from all directions during the duration of the recording (Fig- Virtual seismic sources are dominated by a single dispersive
ure 3). This omnidirectional microseism noise is very favorable for wave mode, corresponding to the fundamental Scholte-wave mode.
seismic interferometry (Weaver and Lobkis, 2002; Wapenaar and Scholte waves are well-formed between 0.4 and 1.2 Hz, and they
Fokkema, 2006). Similar observations are made using beam-steer- have wavelengths as short as 340 m at the upper end of the fre-

a) b) c)

d) e)

Figure 6. Straight-ray tomography maps for group velocities with center frequencies of (a) 0.4–0.6 Hz, (b) 0.6–0.8 Hz, (c) 0.8–1.0 Hz, (d) 1.0–
1.2 Hz, and (e) 1.2–1.4 Hz. Blue lines indicate 2-m contour levels of the bathymetry.
B174 De Ridder and Biondi

quency band and as long as 1460 m at the lower end. Based on of interest to image and monitor the near surface over a producing
an approximate penetration depth of 40% of the wavelength (Low- reservoir. Studies at Valhall have shown that Scholte waves retrieved
rie, 2007), they should exhibit depth sensitivity to approximately from noise correlations are sensitive to small time-lapse changes (De
600 m below the seafloor. Which means that these waves can be Ridder et al., 2014; Mordret et al., 2014).
Downloaded 10/02/15 to 132.203.227.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 7. Resolution indicators for the Scholte- a) b)


wave group-velocity images at the central fre-
quency range 0.8–1.0 Hz. (a) Cumulative ray length
through each cell, (b) model checkerboard grid, and
(c) retrieved checkerboard grid using the raypath
coverage at (a). (d) Adjoint of the tomographic op-
erator applied to final data misfit.

c) d)

a) b) c)

Figure 8. Bathymetry at the Ekofisk field. The blue lines indicate 2-m contour levels of the bathymetry, and the LoFS array station locations
are indicated with yellow dots. (a) Bathymetry in meters with respect to sea level. (b) Magnitude of the gradient of the bathymetry. (c) Laplacian
evaluated on the bathymetry.
Seismic noise tomography at Ekofisk B175

Group-velocity images generally show a high-velocity anomaly AS, and Petoro AS) for access to Ekofisk’s LoFS data and for per-
coinciding with the center of the seafloor subsidence. At the south- mission to publish these results. Many thanks go to O. Knoth, A.
ern end of the array, there is a high-velocity region again. We con- Tura, A. Bertrand, R. Kazinnik, and L. Vedvik of ConocoPhillips
sidered whether the high anomaly could be caused by the flat for helpful discussions and suggestions. We thank S. Levin for his
seafloor approximation in the tomography. But a subsidence of expertise on SEG-D formats and for help with reading the data and
9 m over an approximately 2-km horizontal distance (Figure 8a) meticulously decoding the header fields. This work was carried out
Downloaded 10/02/15 to 132.203.227.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

disproves that a velocity anomaly of more than 20 m∕s cannot as part of a Ph.D. program at Stanford University, and S. de Ridder
be caused by the flat-earth approximation in the tomography. This thanks the sponsors of the Stanford Exploration Project and the
pattern holds over multiple central frequency ranges, which can be Global Climate and Energy Project at Stanford for financial support.
an indication that it is not reflective of lithology but of stress pat- We acknowledge the two anonymous reviewers and the editors for
terns that carry across lithologic boundaries. constructive comments and critique that substantially improved the
In the group-velocity maps (Figure 6e), one can see that the low quality of this manuscript.
velocities form a ring surrounding the seafloor subsidence (Figure 8).
We generally find higher Scholte-wave group velocities within the
subsidence bowl, and lower Scholte-wave group velocities at the REFERENCES
edge of the subsidence bowl. Within the subsidence bowl, the Lap- Aki, K., 1957, Space and time spectra of stationary stochastic waves, with
lacian of the seafloor shape is positive whereas along the edges of special reference to microtremors: Bulletin of the Earthquake Research
the array it is negative (Figure 8c). These areas correspond to where Institute, 35, 415–456.
Amundsen, L., and M. Landrø, 2009, Measuring seismic with light: Geo-
the subsidence causes relative compression and relative decompres- Expro, 6, no. 2.
sion of the stress state. This affects the near-surface shear strengths Artman, B., 2007, Passive seismic imaging: Ph.D. thesis, Stanford Univer-
and thus the Scholte wave velocity. Similar anomalies have been sity.
Aster, R. C., B. Borchers, and C. H. Thurber, 2005, Parameter estimation
found at Valhall field from critically reflected P-waves (Hatchell and inverse problems: Academic Press.
et al., 2009). Weakening of the shear strength in the radial directions Barkved, O. I., 2012, Seismic surveillance for reservoir delivery from a prac-
titioner’s point of view: EAGE Publications bv.
with respect to the center of the subsidence bowl also causes strong Barmin, M. P., M. H. Ritzwoller, and A. L. Levshin, 2001, A fast and reliable
azimuthal anisotropy in surface-wave phase velocities over Ekofisk method for surface wave tomography: Pure and Applied Geophysics, 158,
field (De Ridder et al., 2015). 1351–1375, doi: 10.1007/PL00001225.
Baskir, E., and C. Weller, 1975, Sourceless reflection seismic exploration:
Crosscorrelating longer time recordings should increase the signal- Geophysics, 40, 158–159.
to-noise-ratio, especially at the upper end of the frequency range. Dif- Bensen, G. D., M. H. Ritzwoller, M. P. Barmin, A. L. Levshin, F. Lin, M. P.
ferent noise conditions, potentially under heavy weather, can increase Moschetti, N. M. Shapiro, and Y. Yang, 2007, Processing seismic ambient
noise data to obtain reliable broad-band surface wave dispersion measure-
the signal in the lower and upper ends of the microseism frequency ments: Geophysical Journal International, 169, 1239–1260, doi: 10.1111/j
range. Depending on the exact processing and inversion schemes, .1365-246X.2007.03374.x.
group-velocity images from ambient seismic noise are remarkably Bussat, S., and S. Kugler, 2011, Offshore ambient-noise surface-wave
tomography above 0.1 Hz and its applications: The Leading Edge, 30,
stable from recordings as short as a day (De Ridder and Bi- 514–524, doi: 10.1190/1.3589107.
ondi, 2013). Butler, D. K., ed., 2005, Near-surface geophysics: SEG.
Campillo, M., and A. Paul, 2003, Long-range correlations in the diffuse seis-
mic coda: Science, 299, 547–549, doi: 10.1126/science.1078551.
Claerbout, J., and S. Fomel, 2014, Geophysical image estimation by exam-
CONCLUSIONS ple: Lulu.com.
Claerbout, J. F., 1968, Synthesis of a layered medium from its acoustic trans-
We have shown that the microseism noise in the ambient seismic mission response: Geophysics, 33, 264–269, doi: 10.1190/1.1439927.
field between 0.4 and 1.4 Hz, as recorded by Ekofisk’s LoFS array, Cole, S. P., 1995, Passive seismic and drill-bit experiments using 2-D arrays:
Ph.D. thesis, Stanford University.
is suitable for retrieval of Scholte waves by seismic interferometry. Cox, M. J. G., E. F. Scherrer, and R. Chen, 1999, Static corrections for seis-
The microseism energy at Ekofisk field in this recording is particu- mic reflection surveys: SEG.
Daneshvar, M. R., C. S. Clay, and M. K. Savage, 1995, Passive seismic im-
larly uniformly distributed over the azimuth and thus is ideal for aging using microearthquakes: Geophysics, 60, 1178–1186, doi: 10.1190/
seismic interferometry. Dispersive virtual seismic sources emitting 1.1443846.
Scholte waves between 0.4 and 1.4 Hz are retrieved by passive seis- De Ridder, S., and J. Dellinger, 2011, Ambient seismic noise eikonal tomog-
raphy for near-surface imaging at Valhall: The Leading Edge, 30, 506–
mic interferometry. The interface waves should exhibit a depth sen- 512, doi: 10.1190/1.3589108.
sitivity to approximately 600 m below the seafloor. The Scholte- De Ridder, S. A. L., 2014, Passive seismic surface-wave interferometry for
wave group-velocity tomography locates a high-velocity anomaly reservoir-scale imaging: Ph.D. thesis, Stanford University.
De Ridder, S. A. L., and B. L. Biondi, 2013, Daily reservoir-scale subsurface
in the center of the array, surrounded by a lower velocity region. monitoring using ambient seismic noise: Geophysical Research Letters,
The high-velocity anomaly coincides with the center of the 40, 2969–2974, doi: 10.1002/grl.50594.
De Ridder, S. A. L., B. L. Biondi, and R. G. Clapp, 2014, Time-lapse seismic
seafloor subsidence bowl. The ring of lower velocities corresponds noise correlation tomography at Valhall: Geophysical Research Letters,
with high magnitudes of the bathymetry gradient. Under the 41, 6116–6122, doi: 10.1002/2014GL061156.
southern end of the array, we find higher velocities again. This De Ridder, S. A. L., B. L. Biondi, and D. Nichols, 2015, Elliptical-aniso-
tropic eikonal phase velocity tomography: Geophysical Research Letters,
behavior may reflect overburden stress states caused by decades 42, 758–764, doi: 10.1002/2014GL062805.
of production and reservoir depletion. Draganov, D., 2007, Seismic and electromagnetic interferometry — Retriev-
ing the earth’s reflection response using crosscorrelation: Ph.D. thesis,
Delft University of Technology.
Draganov, D., X. Campman, J. Thorbecke, A. Verdel, and K. Wapenaar,
ACKNOWLEDGMENTS 2009, Reflection images from ambient seismic noise: Geophysics, 74,
no. 5, A63–A67, doi: 10.1190/1.3193529.
Duvall, T. L. Jr., S. M. Jefferies, J. W. Harvey, and M. A. Pomerantz, 1993,
We thank ConocoPhillips Skandinavia AS and the PL018 Part- Time-distance helioseismology: Nature, 362, 430–432, doi: 10.1038/
nership (Total E&P Norge AS, ENI Norge AS, Statoil Petroleum 362430a0.
B176 De Ridder and Biondi

Edme, P., and D. Halliday, 2011, Extracting reflectivity response from point- Ruigrok, E., X. Campman, and K. Wapenaar, 2011, Extraction of P-wave
receiver ambient noise: 81st Annual International Meeting, SEG, Ex- reflections from microseisms: Comptes Rendus Geoscience, 343, 512–
panded Abstracts, 1602–1607. 525, doi: 10.1016/j.crte.2011.02.006.
Eriksrud, M., 2010, Towards the optical seismic era in reservoir monitoring: Ruigrok, E., and K. Wapenaar, 2012, Global-phase seismic interferometry
First Break, 28, 105–111. unveils P-wave reflectivity below the Himalayas and Tibet: Geophysical
Folstad, P. G., L. Amundsen, and M. Landrø, 2010, Monitoring of the Eko- Research Letters, 39, L11303.
fisk Field: GeoExpro, 7, 72–74. Sabra, K. G., P. Gerstoft, P. Roux, W. A. Kuperman, and M. C. Fehler, 2005,
Gerstoft, P., K. G. Sabra, P. Roux, W. A. Kuperman, and M. C. Fehler, 2006, Surface wave tomography from microseisms in Southern California: Geo-
Downloaded 10/02/15 to 132.203.227.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Green’s functions extraction and surface-wave tomography from micro- physical Research Letters, 32, L14311, doi: 10.1029/2005GL023155.
seisms in southern California: Geophysics, 71, no. 4, SI23–SI31, doi: 10 Scherbaum, F., 1987, Levinson inversion of earthquake geometry SH-trans-
.1190/1.2210607. mission seismograms in the presence of noise: Geophysical Prospecting,
Hatchell, P. J., P. B. Wills, and C. Didraga, 2009, Production induced effects 35, 787–802, doi: 10.1111/j.1365-2478.1987.tb02258.x.
on near-surface wave velocities at Valhall: 71st Annual International Scholte, J. G., 1942a, On the Stoneley wave equation I: Proceedings of the
Conference and Exhibition, EAGE, Extended Abstracts, T016. Koninklijke Nederlandse Akademie van Wetenschappen, 45, 20–25.
Hermansen, H., L. K. Thomas, J. E. Sylte, and B. T. Aasbøe, 1997, Twenty Scholte, J. G., 1942b, On the Stoneley wave equation II: Proceedings of the
five years of Ekofisk reservoir management: SPE Annual Technical Con- Koninklijke Nederlandse Akademie van Wetenschappen, 45, 159–164.
ference and Exhibition, 38927. Shapiro, N. M., and M. Campillo, 2004, Emergence of broadband Rayleigh
Landrø, M., and L. Amundsen, 2011, Using water layer normal modes to waves from correlations of the ambient seismic noise: Geophysical Re-
detect shallow gas and CO2 leakage: EAGE Workshop on Permanent search Letters, 31, L07614, doi: 10.1029/2004GL019491.
Reservoir Monitoring (PRM), Extended Abstract, 10341. Shapiro, N. M., M. Campillo, L. Stehly, and M. H. Ritzwoller, 2005, High-
Lévčque, J.-J., L. Rivera, and G. Wittlinger, 1993, On the use of the checker- resolution surface-wave tomography from ambient seismic noise: Sci-
board test to assess the resolution of tomographic inversions: Geophysical ence, 307, 1615–1618, doi: 10.1126/science.1108339.
Journal International, 115, 313–318, doi: 10.1111/j.1365-246X.1993 Shapiro, N. M., and S. K. Singh, 1999, A systematic error in estimating
.tb05605.x. surface-wave group-velocity dispersion curves and a procedure for its cor-
Levshin, A. L., T. B. Yanovskaya, A. V. Lander, B. Bukchin, M. P. Barmin, rection: Bulletin of the Seismological Society of America, 89, 1138–1142.
L. I. Ratnikova, and E. N. Its, 1989, Seismic surface waves in a laterally Snieder, R., 2004, Extracting the Green’s function from the correlation of
inhomogeneous earth: Kluwer. coda waves: A derivation based on stationary phase: Physics Review E,
Lobkis, O. I., and R. L. Weaver, 2001, On the emergence of the Green’s 69, 046610, doi: 10.1103/PhysRevE.69.046610.
function in the correlations of a diffuse field: Journal of the Acoustical Snieder, R., Y. Fan, E. Slob, and K. Wapenaar, 2010, Equipartitioning is not
Society of America, 110, 3011–3017, doi: 10.1121/1.1417528. sufficient for Green’s function extraction: Earthquake Science, 23, 403–
Longuet-Higgins, M. S., 1950, A theory of the origin of microseisms: Philo- 415, doi: 10.1007/s11589-010-0739-1.
sophical Transactions of the Royal Society of London A, 243, 1–35, doi: Stewart, P., 2006, Interferometric imaging of ocean bottom noise: 76th An-
10.1098/rsta.1950.0012. nual International Meeting, SEG, Expanded Abstracts, 1555–1559.
Lowrie, W., 2007, Fundamentals of geophysics 2nd ed.: Cambridge Univer- Wapenaar, K., 2003, Synthesis of an inhomogeneous medium from its acous-
sity Press. tic transmission response: Geophysics, 68, 1756–1759, doi: 10.1190/1
Lyngnes, B., H. Landa, K. Ringen, and N. Haller, 2013, Life of Field Seis- .1620649.
mic at Ekofisk — Utilizing 4D seismic for evaluating well target: 75th Wapenaar, K., 2004, Retrieving the elastodynamic Green’s function of an
Conference and Exhibition, EAGE, Extended Abstracts, We 12 09. arbitrary inhomogeneous medium by cross correlation: Physics Review
Mordret, A., M. Landès, N. M. Shapiro, S. C. Singh, P. Roux, and O. I. Letters, 93, 254301, doi: 10.2528/PIER09041004.
Barkved, 2013a, Near-surface study at the Valhall oil field from ambient Wapenaar, K., and J. Fokkema, 2004, Reciprocity theorems for diffusion,
noise surface wave tomography: Geophysical Journal International, 193, flow, and waves: Journal of Applied Mechanics, 71, 145–150, doi: 10
1627–1643, doi: 10.1093/gji/ggt061. .1115/1.1636792.
Mordret, A., N. M. Shapiro, and S. Singh, 2014, Seismic noise-based time- Wapenaar, K., and J. Fokkema, 2006, Green’s function representations for
lapse monitoring of the Valhall overburden: Geophysical Research Let- seismic interferometry: Geophysics, 71, no. 4, SI33–SI46, doi: 10.1190/1
ters, 41, 4945–4952, doi: 10.1002/2014GL060602. .2213955.
Mordret, A., N. Shapiro, S. Singh, P. Roux, and O. I. Barkved, 2013b, Helm- Weather Underground, Inc., 2013, The weather channel companies: http://
holtz tomography of ambient noise surface wave data to estimate Scholte www.wunderground.com/, accessed 15 October 2013.
wave phase velocity at Valhall Life of the Field: Geophysics, 78, no. 2, Weaver, R. L., and O. I. Lobkis, 2001, Ultrasonics without a source: Thermal
WA99–WA109, doi: 10.1190/geo2012-0303.1. fluctuation correlations at MHz frequencies: Physics Review Letters, 87,
Mordret, A., N. Shapiro, S. Singh, P. Roux, J.-P. Montagner, and O. I. 134301, doi: 10.1103/PhysRevLett.87.134301.
Barkved, 2013c, Azimuthal anisotropy at Valhall: The Helmholtz equa- Weaver, R. L., and O. I. Lobkis, 2002, On the emergence of the Green’s func-
tion approach: Geophysical Research Letters, 40, 2636–2641, doi: 10 tion in the correlations of a diffuse field: Pulse-echo using thermal phonons:
.1002/grl.50447. Ultrasonics, 40, 435–439, doi: 10.1016/S0041-624X(02)00156-7.
Munk, W. H., 1950, Origin and generation of waves: Council of Wave Wielandt, E., 1993, Propagation and structural interpretation of non-plane
Research: Proceedings of the First Conference on Coastal Engineering, waves: Geophysical Journal International, 113, 45–53, doi: 10.1111/j
1–4. .1365-246X.1993.tb02527.x.
Olofsson, B., 2010, Marine ambient seismic noise in the frequency range 1– Wills, P. B., P. J. Hatchell, and S. J. Bourne, 2008, Time-lapse measurements
10 Hz: The Leading Edge, 29, 418–435, doi: 10.1190/1.3378306. of shallow horizontal wave velocity over a compacting field: 70th Interna-
Rickett, J., and J. Claerbout, 1999, Acoustic daylight imaging via spectral tional Conference and Exhibition, EAGE, Extended Abstracts, G039.
factorization: Helioseismology and reservoir monitoring: The Leading Yao, H., R. D. Van Der Hilst, and M. V. De Hoop, 2006, Surface-wave array
Edge, 18, 957–960, doi: 10.1190/1.1438420. tomography in SE Tibet from ambient seismic noise and two-station
Rost, S., and C. Thomas, 2002, Array seismology: Methods and applica- analysis — I. Phase velocity maps: Geophysical Journal International,
tions: Reviews of Geophysics, 40, 1008, doi: 10.1029/2000RG000100. 166, 732–744, doi: 10.1111/j.1365-246X.2006.03028.x.

You might also like