Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/318506091

State-of-the-art remote characterization of shallow marine sediments: the


road to a fully integrated solution

Article in Near Surface Geophysics · September 2017


DOI: 10.3997/1873-0604.2017024

CITATIONS READS

30 2,634

6 authors, including:

Mark Vardy Maarten Vanneste


SAND Geophysics Norwegian Geotechnical Institute
86 PUBLICATIONS 1,388 CITATIONS 132 PUBLICATIONS 3,197 CITATIONS

SEE PROFILE SEE PROFILE

Timothy J. Henstock Michael Clare


University of Southampton National Oceanography Centre, Southampton
218 PUBLICATIONS 4,826 CITATIONS 133 PUBLICATIONS 2,851 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

ODP Leg 188 View project

Seychelles-India View project

All content following this page was uploaded by Mark Vardy on 18 July 2017.

The user has requested enhancement of the downloaded file.


Near Surface Geophysics, 2017, 15, 387-402 doi:10.3997/1873-0604.2017024

State-of-the-art remote characterization of shallow marine


sediments: the road to a fully integrated solution
Mark E. Vardy1*, Maarten Vanneste2, Timothy J. Henstock3, Michael A. Clare1,
Carl Fredrik Forsberg2 and Giuseppe Provenzano3
1
Marine Geosciences, National Oceanography Centre, Southampton SO14 3ZH, United Kingdom
2
Norwegian Geotechnical Institute, 0855 Oslo, Norway
3
Ocean and Earth Science, University of Southampton, Southampton SO14 3ZH, United Kingdom

ABSTRACT
Current methods for characterizing near-surface marine sediments rely on extensive coring/pene-
trometer testing and correlation to seismic facies. Little quantitative information is regularly derived
from geophysical data beyond qualitative inferences of sediment characteristics based on seismic
facies architecture. Even these fundamental seismostratigraphic interpretations can be difficult to
correlate with lithostratigraphic data due to inaccuracies in the time-to-depth conversion of geo-
physical data and potential loss and/or compression of high-porosity and under-consolidated sea-
floor material during direct sampling. To complicate matters further, when quantitative information
is derived from marine geophysical data, it often describes the sediments using terminology (e.g.,
acoustic impedance and seismic quality factor) that is impenetrable to geologists and engineers. In
contrast, for hydrocarbon prospecting, reservoir characterization using quantitative inversion of
geophysical data has developed enormously over the past 20 years or more. Impedance and ampli-
tude-versus-angle inversion techniques are now commonplace, whereas computationally expensive
waveform inversions are gaining traction, and there is a well-developed interface between these
geophysical and reservoir engineering fields via rock physics.
In this paper, we collate and review the different published inversion methods for high-resolution
geophysical data. Using several case study examples spanning a broad range of depositional envi-
ronments, we assess the current state of the art in remote characterization of shallow sediments from
a multidisciplinary viewpoint, encompassing geophysical, geological, and geotechnical angles. By
identifying the key parameters used to characterize the subsurface, a framework is developed
whereby geological, geotechnical, and geophysical characterizations of the subsurface can be
related in a less subjective manner. As part of this, we examine the sensitivity of commonly derived
acoustic properties (e.g., acoustic impedance and seismic quality factor) to more fundamentally
important soil properties (e.g., lithology, pore pressure, gas saturation, and undrained shear
strength), thereby facilitating better integration between geological, geotechnical, and geophysical
data for improved mapping of sediment properties. Ultimately, we present a number of ideas for
future research activities in this field.

INTRODUCTION published literature demonstrably derives more advanced proper-


Current standard practice for characterizing marine sediments in ties, such as gas saturation (e.g., Morgan et al. 2012; Morgan,
the top-hole and foundation zones (i.e., top tens to 200 m), in Vanneste and Vardy 2014; Toth et al. 2014; Cevatoglu et al.
both academia and industry, relies on extensive coring/cone pen- 2015; Toth, Spiess and Kiell 2015; Vardy et al. 2015), that are of
etrometer testing and stratigraphic correlation to seismic facies broader interest to geologists and geotechnical engineers.
(e.g., Stoker et al. 2009; Evans 2011; Power et al. 2011; Vanneste This limited quantitative characterization and disparate termi-
et al. 2012). The derivation of quantitative information directly nology often leads to problems integrating geophysical data with
from geophysical data is rare and often limited to acoustic prop- the results from directly sampled geological and geotechnical
erties primarily of interest to geophysicists (e.g., Pinson et al. data. Purely qualitative integration based on seismo- and
2008; Vardy et al. 2012; Vardy 2015). Only a small subset of the lithostratigraphic correlation is often difficult because, while the
geological/geotechnical data are recorded in depth below sea-
*
mev@noc.ac.uk floor (SF), the geophysical data are acquired in two-way travel

© 2017 European Association of Geoscientists & Engineers 387

19693-NSG17 August BOOK.indb 387 12/07/17 09:56


388 M.E. Vardy et al.

time, and therefore, the geophysical data require time-to-depth geotechnical parameters is assessed using several case study
conversion, which depends on accurate knowledge of the seismic examples that span a broad range of depositional environments
velocity field. Time-to-depth conversion is challenging, as (fjord basin, open slope, glacially altered, and deep water). This
velocities in the shallow subsurface can vary rapidly, both later- aims to ascertain the limitations on what useful information can
ally and with depth. This is particularly true in geologically be derived from the geophysical data. Based on the results from
complex environments such as those in previously glaciated these case studies, we identify a set of key challenges to be
areas (e.g., Pinson et al. 2013). Where subsurface velocities have addressed and suggest some potential pathways forward towards
been derived (or are simple enough to approximate), geological a novel integrated approach for shallow marine sediment charac-
and geophysical data have been successfully integrated to gain terization or geotechnical site characterization.
insights into a variety of processes, including sediment deforma-
tion in peri- and paraglacial environments (e.g., Stoker and CHARACTERIZATION OF SEDIMENTS
Bradwell 2009) and submarine geohazards (e.g., Vardy et al. One of the major issues with taking a coherent integrated
2012; Vanneste et al. 2014). Furthermore, for imaging the stra- approach to the characterization of shallow subsurface marine
tigraphy of shallow sediments, one often relies on short-offset sediments is the breadth and multidisciplinary nature of methods
single-channel seismic recording systems, which do not allow used to describe the make-up of marine sediments. Geologists,
the derivation of seismic velocities in the subsurface. In contrast, sedimentologists, geotechnical engineers, and geophysicists
the use of multichannel streamers would alleviate this problem have all contributed various, often complementary, methods for
on the condition that the channel spacing is sufficiently small (c., characterizing marine sediments, spanning everything from
1 m for a typical high-resolution Boomer or Sparker source) and micro-computed tomography (CT) imaging of sedimentary grain
the active length of the streamer is not shorter than the depth to textures to basin-scale geophysical imaging of facies structure.
the deepest target. Historically, these different disciplines have often worked inde-
Effectively integrating geophysical with geotechnical data is pendently, with little to no collaboration due to their differing
altogether more difficult. This is made more complex by the con- objectives and have therefore developed fundamentally different
trasting strain regimes. Geophysical data are at relatively low methods based on fundamentally different philosophies and with
strains, whereas geotechnical data are at large strain (Clayton conflicting conventions. As of late, the focus has moved towards
2011). In addition, intrusive methods cause mechanical deforma- a multidisciplinary integration in which the complementary
tion, which can be very significant for high-porosity near-surface aspects of the various disciplines form the key to progress (e.g.,
sediments during geotechnical sampling procedures (e.g., Evans 2011; Vanneste et al. 2012).
Clayton, Hight and Hopper 1992; Lunne 2012; Priest et al. 2015). In the following sections and Table 1, we attempt to briefly
This deformation inherently alters the nature of the sampled summarise/outline the basic principles behind the characterization
material, introducing uncertainties on key mechanical parameters of shallow subsurface marine sediments from three different view-
(e.g., porosity, pore pressure, and undrained shear strength) that points: geologists/sedimentologists; geotechnical engineers; and
are difficult to quantify. Where geophysical and geotechnical data geophysicists. It is worth noting that the different disciplines
have been integrated, the methods either are highly qualitative essentially rely on the same type of data; that is predominantly
(e.g., Vanneste et al. 2015) or suffer from large uncertainties due geophysical data (2D and sometimes 3D seismic reflection data,
to imprecisely calibrated relationships between the geophysical bathymetry data, and side-scan sonar or backscatter data) and 1D
and geotechnical parameters (e.g., Nauroy et al. 1998). in situ logs (e.g., CPTU data) or samples (or combined boreholes).
While geophysical data alone will never be able to geologi-
cally or geotechnically characterize shallow subsurface sedi- Geological/sedimentological viewpoint
ments in the same manner as ground truth data acquisition, these Geologists and sedimentologists analyse sediments on all scales
data are currently underused as a quantitative tool and can sig- to understand the processes (erosion, transport, deposition, and
nificantly contribute towards deriving more robust, high-fidelity post-depositional modification) that led to their formation. This
spatial maps of sediment properties. Here we seek to outline a information underpins a wide range of applications such as
procedure whereby geological, geotechnical, and geophysical reconstructing depositional environments to understand past cli-
data can be more effectively integrated to produce a coherent, mates (e.g., determination of ice grounding lines; Scourse et al.
detailed, and less subjective interpretation of the shallow subsur- 2009); determining palaeocurrent directions (e.g., Gardner,
face. We start by summarising the fundamental principles, meth- Richardson and Cacchione 1989); quantifying the quality, lateral
ods, and goals of each discipline, collating this information in an continuity, and connectivity of hydrocarbon reservoirs (Mayall,
attempt to reconcile differences in conventions and find common Jones and Casey 2006); assessing geohazard processes (e.g.,
ground. Based on this, a framework is derived that permits the landslides and debris flows; Solheim et al. 2005; L’Heureux et
key parameters that quantitatively describe the geological and al. 2012) based on quantitative interpretation of deposits, includ-
geotechnical properties of the subsurface to be identified. The ing frequency–magnitude estimations (Masson, Wynn and
sensitivity of the seismic wave field to these key geological/ Talling 2010; Clare et al. 2014); and determining organic carbon

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 388 12/07/17 09:56


Remote characterization of shallow marine sediments 389

transport between different physiographic settings (e.g., Hilton et In order to quantify these interpretations, a large range of spe-
al. 2011). cific measurements/primary characteristics are made on scales
To achieve this, geologists and sedimentologists aim to quan- ranging from micrometres to thousands of kilometres. The sedi-
tify the following: ment composition is expressed in terms of lithology, mineralogy,
• Facies: the fundamental building blocks of sedimentological and fossil content (Tucker 2003; Stow 2005). Sedimentary tex-
interpretation that provide the link to a physical mechanism for ture characterizes sediment in terms of sorting and grading in
transport, deposition, and/or post-depositional modification, grain size, grain morphology/surface texture, and the fabric.
such as understanding energy of environment and transport or Other key information is provided from the examination of the
erosion processes (e.g., bedload and suspended load; Tucker thickness, geometry, and nature of bed boundaries; inclination
2003; Stow 2005). and orientation of bedding; and the presence or absence of sedi-
• Facies sequences and cycles: provide information about tem- mentary structures, both on bedding surfaces and within beds.
poral variations in deposition, which may be related to cli- For instance, energetic processes can create erosional surfaces
matic, eustatic, tectonic, or other controls. that punctuate the depositional record, ranging from flute marks
• Lateral trends and geometry: inform our spatial and physi- at the base of turbidites to regional unconformities (e.g., related
ographic interpretation about the palaeogeography, such as to glacial processes), ploughmarks caused by iceberg movement,
what happened up-, down-, or along slope and address the incision of submarine canyons and channels, and landslide head-
sediment’s mobility (e.g., sand waves). scars and mass-transport deposits covering scales from tens to
• Facies associations and architectural elements: characterize hundreds of metres (e.g., 1996 slide near Finneidfjord; Vardy et
the relationships between individual facies both spatially and al. 2012) to hundreds of kilometres (e.g., Storegga slide com-
temporally. plex; Masson et al. 2010).
• Sequence stratigraphy and bounding surfaces: identify The nature of the deposits coupled with their resultant geo-
cycles, geometry, and unconformities to understand major morphology provides a qualitative idea on the depositional envi-
shifts and switches in deposition (e.g., sudden influx of sedi- ronment and of specific processes (e.g., en masse emplacement
ment or change in depositional regime) and erosional episodes of deposits or ripples that indicate flow direction, current veloci-
(e.g., sudden uplift and onset of channel incision). ties, and/or flow regime). Post- or syndepositional modification
• Controls, rates, and preservation: quantify the controls on is deciphered from the nature of sediment deformation and its
spatial and temporal variations in deposition and erosion, vari- interaction with the host sediment (e.g., mass-transport-related,
ation in sediment accumulation rates through time, and the glaciotectonism and halokinesis, water escape during consolida-
completeness of the depositional record (Sadler 1981). tion, or brecciation due to mud volcanism, biogenic, and/or
• Depositional environments: enable, with the integration of chemogenic alteration).
the aforementioned factors in this list, interpretation of the Further to the identification of environment and depositional
environment of deposition and erosion (e.g., alluvial, fluvial, processes, a temporal context is defined by dating undisturbed
Aeolian, and deep/shallow marine). sediments using biostratigraphy (microfossils and macrofossil

Table 1 Summary of geological, geotechnical, and geophysical characterization of the marine subsurface.
Geological Geotechnical Geophysical
Lithostratigraphic facies Soil units Seismostratigraphic facies
Units Thickness/Geometry Thickness/Geometry Thickness/Geometry
Lithology Lithology Impedance
Mineralogy Mineralogy Interval velocity
Fossil content Water content Bulk density
Biological alteration Index parameters Q-factor
Chemical alteration Undrained shear strength/Relative density
Boundaries Erosional features Soil type contrast (e.g., clay over sand) Impedance contrasts
Graded transitions Consistency contrast (e.g., high shear Q-contrasts
Bedforms strength over low shear strength soil) Phase reversals
Grain size contrasts
Internal Sorting direction Soil fabric Reflection character/attitude/sequence
architecture Bedding Permeability Deformation
Deformation Pore fluid Truncated reflections
Truncated beds On/off/downlap

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 389 12/07/17 09:56


390 M.E. Vardy et al.

analysis), radiometric and isotopic methods, magnetostratigra- plished using CPT data (with pore pressure measurements). The
phy, tephrochronology, and/or cyclostratigraphy (e.g., Bell and particle size distribution (clay, silt, sand, and gravel) largely gov-
Walker 2005). erns hydraulic conductivity and permeability and determines
In addition to the sedimentological (and geochemical) analy- whether the soil will behave in a drained or an undrained manner.
ses outlined above, the retrieved samples (in plastic liners) are For cohesive soils (e.g., clays), the clay mineralogy and activity
also typically scanned for physical properties using a multi-sen- are important (e.g., swelling) and the shear strength. For granular
sor core logging facility. This produces a continuous very high- soils (e.g., sands), the relative density is typically used. Routinely
resolution (sub-centimetre scale) record of P-wave velocity, measured index properties include water content, plasticity index,
P-wave attenuation, bulk density, and magnetic susceptibility, liquid index, wet and dry bulk density/unit weight, particle den-
providing further insights into the sedimentary history at the site. sity, void ratio, and porosity. The presence of biogenic material is
Together, these measurements permit a detailed interpretation also evaluated, as organic material can have a strong influence on
of the spatial and temporal evolutions in depositional environ- the particle–particle bonding and compressibility. A variety of
ment and processes on scales that span metres to whole basins strength–strain testing (e.g., triaxial testing or direct simple shear
and days to millennia (Rothwell and Rack 2006). testing, with various modes of application, like anisotropic con-
solidation under compression or extension, static, or cyclic)
Geotechnical engineering approach reveal the strength behaviour of the soil under specific conditions
Geotechnical engineers aim to quantify how Earth materials (soil (e.g., dilative or contractive). Such tests are typically combined
and rock) behave in response to both natural (e.g., slope failures with oedometer testing to evaluate the stress history of the given
and sediment loading/unloading) and/or anthropogenic (e.g., sample and provide an estimate on the soil’s permeability.
installation of foundations) processes. Example applications Geotechnical tests are often complemented with X-ray or CT
include 2D/3D slope stability analysis, dynamic modelling of imaging to reveal the sediment structure and fabric at the micro-
landslides and impact assessment, site response analysis during scale, e.g., grain packing and distribution, as well as morphology
earthquakes, foundation design, and engineering ground model- and connectivity of voids. Sometimes, specific geochemical tests
ling. However, quantifying this behaviour is difficult because are used to characterize the pore fluids, including salinity and
natural soils are complex, and therefore, their properties and presence (or absence) of gas(es), hydrates, and/or hydrocarbons.
behaviour are hard to predict (Terzaghi 1936). This means that, Advanced geotechnical tests are also used to calibrate the 1D in
in order to understand these processes, geotechnical engineers situ test measurements (CPTU) data and obtain a sample quality
look to quantify the following: value (e.g., Lunne 2012).
• Sample-specific nature: How the geotechnical properties vary Together, these measurements permit establishing a geotech-
from sample to sample, taking into account the quality of the nical soil profile and a detailed assessment of how the soil might
sample (which is affected by the sampling process). behave in response to future loading and/or unloading events,
• Soil units: Parameterise soil units based on a similar range in and they are the fundamental basis for foundation design.
geotechnical properties from the sample-by-sample testing or
in situ characterization (e.g., cone penetration testing (CPT) Geophysical methods
with pore pressure measurements; CPTU). The method by which the subsurface is quantitatively character-
• Soil profiles: Develop soil profiles that describe how a soil has ized using geophysical data is broadly termed seismic inversion.
been modified post- and syndeposition through processes such This is an immensely diverse subject but, at the most basic level,
as consolidation and determine representative trends. involves applying a physical understanding of how the seismic
• Soil provinces: Lateral demarcation of areas featuring similar wavefield propagates to predict/derive subsurface conditions
geotechnical conditions into soil provinces (e.g., Campbell based on the amplitude, phase, moveout, and frequency content
1984). of recorded reflections. While there is a well-known long track
• Location-specific nature: Predict the location-specific record of the inversion of marine geophysical data for hydrocar-
response to loading for a specific application, e.g., foundation bon reservoir characterization (e.g., Mallick 2001; Bosch,
installation. Mukerji and Gonzalez 2010; Wagner et al. 2012), there is also a
• Site-wide analysis: Predict the impact of external forces act- surprisingly long (if not so well-known) published record
ing upon the soils in the survey area, for example, liquefaction (Table 2) of the inversion of high- and/or ultra-high-resolution
due to earthquakes/vibrations, or slope stability assessment. marine geophysical data (e.g., Schock, LeBlanc and Mayer
• Soil-structure interaction: Assess how the soil influences the 1989; Panda, LeBlanc and Schock 1994; Nauroy et al. 1998;
infrastructure and vice versa. Pinson et al. 2008; Morgan et al. 2012, 2014; Holland and
Dettmer 2013; Ker, Gonidec and Gilbert 2013; Vardy 2015;
In order to interpret or determine these soil characteristics, a Vardy et al. 2015). Traditionally, however, geophysicists are
broad spectrum of measurements is made from laboratory tests on interested in quantitatively characterizing the shallow subsurface
soil samples or via in situ geotechnical tests typically accom- using parameters that control the propagation of the seismic

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 390 12/07/17 09:56


Remote characterization of shallow marine sediments 391

wavefield because at the heart of all seismic inversion approach- Table 2 Previously published work on quantitative geophysical
es lies a reliable forward model that predicts the observed seis- characterization of the marine near surface.
mic data. The approaches to the forward modelling of marine Reflectivity
seismic reflection data can be broadly divided into three princi-
Reflection coefficients: Chotiros (1994)
ple categories.
Bull et al. (1998)
• Reflectivity (or acoustic impedance): describes the interac-
tion between the seismic wavefield and impedance contrasts in Cevatoglu et al. (2015)
the subsurface. Most commonly, this is considered as a much Acoustic impedance: Panda et al. (1994)
simplified normal-incidence problem, limiting the seismic Sternlicht and Moustier (2003)
wavefield to P-waves alone and permitting the reflectivity to Guo et al. (2006)
be explicitly calculated as a reflection coefficient (e.g.,
Ker et al. (2013)
Chotiros 1994; Bull, Quinn and Dix 1998) or inverted as
reflectivity series using a computationally fast convolutional Zhang and Digby (2013)
approach (e.g., Panda et al. 1994; Ker et al. 2013; Vardy 2015). Vardy (2015)
For hydrocarbon applications, this has been expanded to mul- Vardy et al. (2015)
tiple incidence angles in the form of extended elastic imped- Wang and Stewart (2015)
ance or amplitude-versus-offset (AVO) inversion (e.g., Ma
Attenuation
2002), which derives S-wave and P-wave impedance. To date,
P-wave attenuation: Schock et al. (1989)
this has only been applied to shallow-section exploration data
(Sobreira et al. 2010) and is yet to be applied to high-resolu- Panda et al. (1994)
tion marine data. Stevenson et al. (2002)
• Intrinsic attenuation: model how the seismic wavefield loses Robb et al. (2006)
energy through viscous dissipation and heating at grain bounda- Pinson et al. (2008)
ries as it propagates through the subsurface. This is often calcu-
Morgan et al. (2012, 2014)
lated in terms of the seismic quality factor (Q), which is inversely
proportional to the attenuation (e.g., Stephenson et al. 2002; Vardy et al. (2012)
Pinson et al. 2008; Morgan et al. 2012). Traditionally, this is Cevatoglu et al. (2015)
treated as a purely P-wave problem; however, there has been some Lei and Morgan (2015)
limited work on S-wave attenuation (e.g., Lei and Morgan 2015). S-wave attenuation: Lei and Morgan (2015)
• Full-waveform (FW): full propagation modelling to describe
Full-waveform
how fast the seismic wavefield propagates, how energy is
P-wave velocity: Gerstoft (1994)
reflected and mode converted between P- and S-waves at inter-
faces (including surface waves and refractions), and how Ohta et al. (2005)
energy is lost through attenuation. FW derives multi-parameter Ballard, Becker and Goff (2010)
subsurface models in terms of P- and S-wave velocities, P- and Pinson et al. (2013)
S-wave Q, and bulk density but comes at a significant compu- Toth et al. (2014, 2015)
tational cost (e.g., Holland and Dettmer 2013; Provenzano,
Aleardi et al. (2016)
Vardy and Henstock 2016). Often, due to the computational
expense of FW, it is simplified to only consider velocity, either S-wave velocity: Ayres and Theilen (1999)
inverting for (e.g., Gerstoft 1994; Aleardi, Tognarelli and Huws et al. (2000)
Mazzotti 2016) or explicitly deriving (e.g., Pinson et al. 2013; Bohlen et al. (2004)
Toth et al. 2014), which provides improved imaging and lim- Winsborrow, Huws and Muyzert (2005)
ited quantitative information. Allouche et al. (2010, 2011)
Socco et al. (2011)
When combined with qualitative structural interpretation of
seismic reflection profiles and/or 3D volumes (and increasingly Vanneste et al. (2011)
with seismic attributes), these methods allow the shallow sub- Drijkoningen et al. (2012)
surface to be interpreted in terms of: large-scale seismic facies Full waveform: Seong and Park (2001)
distribution; heterogeneity and coherency of bedding within Holland and Dettmer (2013)
facies; dip, azimuth, and grading of coherent bedding; as well as
Dosso et al. (2014)
structure and material contrast at both conformable and uncon-
formable bounding surfaces, but also velocity and/or attenuation Provenzano et al. (2016)
values. Provenzano et al. (2017)

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 391 12/07/17 09:56


392 M.E. Vardy et al.

However, a major limitation with the quantitative methods to basic stratigraphic principles, specifically to simplify the sub-
applied to the marine shallow subsurface is that S-waves or sur- surface into a series of stratigraphic units described in terms of
face waves are only rarely considered. Notable exceptions are the their overall structure, the nature of the boundaries between
multi-model surface wave inversion work of Socco et al. (2011) them, and their detailed internal architecture. By determining
and Vanneste et al. (2011) and some limited mode-converted which parameters are consistently important in defining the
wave analysis (Bohlen et al. 2004; Allouche, Drijkoningen and nature of and boundaries between these stratigraphic units, we
van der Neut 2010; Allouche et al. 2011; Drijkoningen et al. can identify five fundamental parameterisations that could be
2012). As S-waves are transmitted through the grain skeleton, used to form the framework for an integrated interpretation:
they provide a more direct link between the geotechnical and 1. Structure: Understand the structural relationship between
geomechanical parameters than P-waves. In the marine environ- different stratigraphic units, including the nature and extent of
ment, however, they are both harder to generate and more com- any in situ deformation (faulting, folding, and remoulding).
plex to record (e.g., Huws, Davies and Pyrah 2000). This is key in determining how the stratigraphic units were
deposited and have been deformed/reworked subsequent to
INTEGRATED FRAMEWORK deposition.
A major obstacle in the successful integration of geological, 2. Lithology: Variations in lithology both between and within
geotechnical, and geophysical data is the sometimes-contradicto- the stratigraphic units, particularly the ratio of cohesive to
ry terminology employed. Geological and geotechnical engi- granular material, and both horizontal and vertical sorting. It
neering analysis and interpretation are fundamentally comple- provides critical information about the depositional environ-
mentary. They both make extensive use of isolated 1D direct ment and how a soil will behave under future loading.
sampling measurements, and while geological analysis focuses 3. Stress regime: Parameterise the relative volumes of pore
on understanding the depositional and environmental history of space and grain skeleton, identifying areas of overpressure
the sediments, the geotechnical analysis seeks to predict how the and degree of consolidation (relative to normal consolida-
sediments may respond to future changes in the environment. tion). This is a useful proxy for rates of deposition, erosion,
Geophysical data should be a highly useful bridge, enabling the and uplift and is critical in understanding how a soil will
detailed geological and geotechnical analysis at multiple isolated behave under future loading.
locations to be integrated into a robust, high-fidelity site-wide 4. Pore fluid: Quantify the nature of the fluid or fluids filling the
interpretation or ground model. However, in order to do this, pore space and how this spatially varies, which is critical in
from a geophysical perspective, it is necessary to derive param- predicting how a soil will behave under loading.
eters and/or an understanding that is of broader use than the 5. Strength/stiffness: Geotechnical characterization of the sub-
wavefield propagation according to acoustic impedance, P-wave surface critically requires the quantification of the strength
velocity, Q, etc. and stiffness of a soil unit, particularly under a shearing load.
Table 1 collates the geological, geotechnical, and geophysical
terminologies outlined in the previous section. Fundamentally, Over the following sections, each of these parameters is consid-
all three disciplines seek to characterize the subsurface according ered in turn using a range of different example datasets

Figure 1 Location map of data


examples used. Data, labelled in
order of appearance, includes (A)
Windermere, UK; (B) British
Columbia, Canada; (C) California,
USA; (D) Finneidfjord, northern
Norway; (E) Solent, UK; (F)
Oban, UK; and (G) eastern Irish
Sea, UK.

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 392 12/07/17 09:57


Remote characterization of shallow marine sediments 393

Figure 2 Example seismic reflec-


tion sections from: (a)
Windermere, UK; (b) British
Columbia, Canada; and (c)
California, USA. Data from
Windermere were acquired using
a Boomer source and a short mul-
tichannel streamer and have been
pre-stack depth migrated. Data
from British Columbia were also
acquired using a Boomer source
and a short multichannel streamer
but have been pre-stack time
migrated. Data from California
were acquired using the 3D Chirp
decimetre-resolution sub-bottom
profiler and have been normal
moveout corrected and stacked
onto a 25-cm binning grid.

(Figure 1). Where possible, the results from a range of different quency bandwidth of the waveform fundamentally controls the
seismic inversion methods are presented to outline the sensitivity detail of structure that can be quantified. Both the Windermere
of different components of the seismic wavefield (e.g., imped- and British Columbia datasets (Figure 2a and 2b) use a Boomer
ance, Q) to different sediment properties. This provides insights source (0.5–3.0 kHz) that can resolve structure on a 20- to 50-cm
into how the results from different seismic inversion approaches vertical scale in softer sediments, whereas the Californian dataset
can be combined to develop a more complete parameterisation of (Figure 2c) was acquired using a broadband Chirp source (1.5–
the subsurface. 13.0 kHz; Gutowski et al. 2002) that can resolve structure on the
5- to 15-cm scale. Additionally, the velocity of the subsurface
Structure also influences the resolution, with higher velocities stretching
The use of seismic reflection data for structural imaging is widely the wavelet and reducing the resolution. Notice the difference in
accepted. Figure 2 shows three examples of structural imaging at resolution between the glaciogenic landforms and the glaciola-
different scales. Figure 2a illustrates spatial variation and interac- custrine/lacustrine stratigraphic units in the depth-imaged seis-
tion of stratigraphic units on metre and decametre scales from a mic section from Windermere (Figure 2a). Furthermore, the
glacial lake in Windermere, UK, with a sequence of glaciogenic phase of the waveform is critically important. Airgun, Sparker,
landforms at depth, overlain by a thick sequence of glaciolacus- and Boomer sources are impulsive and, therefore, are approxi-
trine, in situ reworked (mass wasting), and modern lacustrine mately minimum phase, which means that the stratigraphic
sedimentation. Figure 2b shows a much simpler bedded stratigra- interface corresponds to the first zero-crossing, not an amplitude
phy from British Columbia, Canada, where pore fluids are peak or trough. In contrast, Chirp data are zero-phase, in which
trapped at several stratigraphic levels (amplitude anomalies) and case the stratigraphic interface does correspond to a peak or
the whole sequence from 380-ms two-way travel time (TWT) to trough in the seismic amplitude data.
the SF is bisected vertically by a focused fluid migration pathway. In the classical work on seismic reflection response to thin
Figure 2c shows bedding characterized by fine-scale grain size beds, Widess (1982) demonstrated the complex relationship
variations within a bedform sequence in California, USA. between two closely spaced reflective interfaces and the ampli-
Similar to illustrating different kinds of structure that can be tude/phase response recorded in the seismic reflection dataset. In
imaged using high- and ultra-high-resolution seismic reflection particular, it was demonstrated that, once the gap between the
data, the datasets presented in Figure 2 also demonstrate the reflective interfaces drops below the dominant wavelength (λ) of
importance of understanding the interaction between the source the seismic source waveform, tuning of reflected wavelets causes
waveform and the subsurface. At the most basic level, the fre- certain thin-bed thicknesses to have anomalously high and low

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 393 12/07/17 09:57


394 M.E. Vardy et al.

amplitudes, as well as wavelet peaks, troughs, and zero-crossings data (S/N ~ 10), a reflection coefficient of 0.05 would be
to become mislocated relative to the true reflectivity interface. approaching the minimum that can be reliably inverted. In real-
This complex interaction between the seismic waveform and the ity, therefore, only larger contrasts in grain size associated with
subsurface means that, when compared against cores, boreholes, sudden changes in deposition style and/or non-conformable
and CPT data that sample at centimetre resolution, misinterpreta- boundaries could be quantified with any level of confidence. The
tion of the geophysical data at the wavelet scale can cause the velocity values in Figure 3b similarly suggest that it is hard to
location of subsurface interfaces to mismatch by tens of centime- differentiate lithologies based on velocity alone, unless large
tres, whereas very fine-scale structures may not be imaged at all. contrasts in grain size are considered. This is best illustrated in
Figure 3c, where the interval velocities are used to calculate the
Lithology magnitude of the difference in travel time moveout for a reflec-
Lithology is a critical parameter for both geological and geotech- tion from the base of a 10-m-thick sediment layer in 40 m of
nical applications. The grain size distribution and sorting (in both water recorded with a source–receiver offset of 50 m (i.e., offset
vertical and horizontal directions) provides information about equal to the imaging depth). These data indicate that the differ-
the depositional environment, whereas differentiating between ence in recorded moveout, were this layer to comprise very
cohesive and granular materials has major implications for the coarse sand rather than clay, would be c. 1.0-ms TWT and, there-
behaviour of a soil under loading. Lithology, however, is a fore, resolvable by most high- and ultra-high-resolution sources.
parameter that traditional marine seismic reflection techniques However, subtler differences in the grain size of this layer are
(i.e., using P-wave) are poorly sensitive to, as harder to resolve, with the l/2 for a typical Boomer/Sparker
waveform (c. 0.5 ms) equating to c. 4 φ.
When considering real sediments, the ability to discern subtle
(1) changes in lithology using traditional seismic reflection tech-
niques is further complicated by changes in porosity. The major-
where Vp is the P-wave velocity, K is the bulk modulus (inverse- ity of seismic inversion techniques applied to marine near-sur-
ly proportional to compressibility), m is the shear modulus (i.e., face data consider only P-waves (see Table 2), whose propaga-
rigidity), and ρ is the bulk density. tion is controlled by bulk density and velocity, which combine
Figure 3a and 3b plot the magnitudes of the normal incident influences from the grain skeleton and the infilling pore fluid.
reflection coefficient and contrast in interval P-wave velocity for These bulk properties are therefore strongly influenced by small
mean grain sizes ranging from φ = 0 (c. 1.0 mm; very coarse changes in porosity, with just a 5% change in porosity, equating
sand) to φ = 10 (c. 0.004 mm; clay) assuming a consistent poros- to a reflection coefficient of c. 0.04 assuming no change in grain
ity. This illustrates that, while we might expect to see reflections size. Figure 4 shows example data from Finneidfjord, northern
from subtler contrasts in grain size (a change in grain size of 1 φ Norway, where the derived impedance structure is the opposite
roughly equates to a reflection coefficient of 0.05), the robust expected from the lithology, demonstrating a discrete decrease in
quantitative detectability of these subtler contrasts is questiona- acoustic impedance associated with a thin sand seam. This can
ble. Vardy et al. (2015) demonstrate using acoustic impedance be shown to correlate with excess pore pressure generated fluids
inversion of synthetic data that, while under excellent signal-to- accumulating in and immediately above the sand seam, rather
noise (S/N) conditions, it is possible to successfully invert for than the pure lithological response of the complex interbedded
reflection coefficients down to 0.025, under more realistic noise clay/sand unit (L’Heureux et al. 2012; Vardy et al. 2012; Vardy
conditions for high- and ultra-high-resolution seismic reflection 2015).

Figure 3 Multi-panel figure illus-


trating (a) the magnitude of nor-
mal incidence reflection coeffi-
cients at an interface with a range
of difference grain size contrasts;
(b) the magnitude of the change
in interval velocity for the same
range of grain size contrasts; and
(c) the magnitude of the change
in reflector moveout at 50-m
source–receiver offset from the
base of a 10-m-thick layer of
sediment under 40 m of water.

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 394 12/07/17 09:57


Remote characterization of shallow marine sediments 395

Figure 4 Panel (a) shows an arbitrary line through a 3D Chirp seismic volume crossing two ground truth sites where long piston cores and CPT data
have been acquired. Panels (b) and (c) show the acoustic impedance and porosity measured from the Calypso 2 core (red) and predicted using the
acoustic impedance inversion (black) of Vardy (2015). Panel (d) compares these physical properties against the core lithostratigraphy, whereas panel
(e) shows trace-by-trace variation in Q calculated using the match-filtering technique of Cheng and Margrave (2012).

Lithological characterization of the subsurface is most robust- tering method of Cheng and Margrave (2012). Broadly, these
ly achieved using either S-waves and/or Q to complement the results agree with the lithology, identifying the thin bed at c.
common P-wave measurements. Both S-wave and Q properties 69-ms TWT as having a significantly higher Q (>300) and
are primarily sensitive to the grain skeleton rather than the skel- therefore implying it is clay rich (Pinson 2010).
eton/pore space ratio:
Stress regime
(2) The stress regime is geologically useful in understanding how
rapidly a stratigraphic unit was buried and if it has been subse-
where Vs is the S-wave velocity and m is the shear modulus quently uplifted or exposed (e.g., by scouring or faulting),
(rigidity). whereas, geotechnically, it provides insights into the stability and
S-waves have long been known to have an increased sensi- in situ strength of the soils. Traditionally, the stress regime
tivity to lithology (as well as pore pressure conditions) over within a fully saturated soil is considered as a balance between
P-waves (e.g., Ayres and Theilen 1999), but quantifying S-wave the compressive loading force of the soil skeleton and the expan-
structure in the marine setting is difficult, relying on either sive force of the incompressible pore water (Terzaghi 1925):
using complex and often unwieldy source/receiver geometries
(e.g., Vanneste et al. 2011) or estimating mode-conversion (3)
through AVO effects as part of an FW inversion (e.g.,
Provenzano, Vardy and Henstock 2017). Q-factor, in contrast, where s is the total stress (lithological loading), u is the pore
can be estimated using several different techniques (e.g., pressure, and s’ is the effective stress.
Schock et al. 1989; Stevenson, McCann and Runciman 2002; A by-product of the integrated relationship between P-wave
Pinson et al. 2008; Morgan et al. 2012) and has been shown velocities and the pore space is a sensitivity to the modern in situ
empirically to have a strongly bi-modal relationship with grain stress regime. Figure 5a shows a section of multichannel Boomer
size, in essence, differentiating between cohesive and granular seismic reflection data from Solent, UK. Velocity analysis of
materials (Pinson et al. 2008; Vardy et al. 2012). However, these data identifies a very high-velocity (c. 1900 m/s) surface
accurately quantifying Q is difficult. It can only be reliably layer consistent with a laterally extensive stratigraphic unit with
calculated as a property between discrete reflections (with a a chaotic internal architecture, underlain by a sequence of dip-
minimum travel-time difference to make the attenuation resolv- ping beds with medium-to-high velocities (1700–1750 m/s) for
able) and often demonstrating significant trace-to-trace varia- such shallowly buried sediments. Empirical relationships
tion due to non-white contamination of the spectrum due to between P-wave velocity and lithology (e.g., Hamilton and
noise and/or thin-bed tuning. Figure 4e shows Q-estimates Bachman 1982; Richardson and Briggs 1993) would suggest the
made for the seismic section in Figure 4a using the match-fil- surface layer consists of gravels, whereas the dipping beds con-

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 395 12/07/17 09:57


396 M.E. Vardy et al.

Figure 5 Panel (a) shows a sec-


tion of Boomer multichannel data
from Solent, UK, overlain with
the interval velocity model used
to perform pre-stack depth imag-
ing and the location of a rotary
borehole. Panel (b) shows the
lithological log for the top 15 m
of the borehole, identifying a
gravel layer at the seabed, overly-
ing a sequence of interbedded
siltstones and mudstones, with
occasional shelly layers. Panel (c)
shows the vertical component of
the pore pressure estimated using
the same methodology as Vardy
et al. (2015).

sist of coarse sands. While a coincident borehole confirms the Pore fluids
chaotic surface layer as being modern alluvial gravels, the dip- Quantitatively differentiating between different pore fluids (and
ping beds are over-consolidated Tertiary siltstones and mud- their in situ pressure) is critical to a wide range of geological and
stones (Figure 5b). Estimation of Q from the seismic reflection geotechnical applications, such as mapping methane hydrates,
data derives a bulk value of 147 for the dipping beds, which is foundation design, drilling hazard assessment (shallow water
consistent with the silt and clay lithology (i.e., cohesive; Pinson flows or shallow gas accumulations), organic carbon cycling/
et al. 2008) and therefore demonstrates the usefulness of com- burial, and anthropogenic carbon sequestration. Shallow gas, in
bining velocity and attenuation analysis for identifying over- particular, is hugely important, and therefore quantifying the in
consolidated stratigraphic units. situ gas saturation has received a significant amount of research
With the application of more advanced inversion strategies interest using a variety of different techniques, including vertical
that derive porosity and density information, the in situ stress seismic profiling (e.g., Holbrook et al. 1996), P-wave seismic
regime can be quantitatively estimated. If it is assumed that the reflection coefficients (e.g., Cevatoglu et al. 2015) and interval
stress regime is dominated by vertical loading, then the vertical velocities (e.g., Leighton and Robb 2008; Toth et al. 2014, 2015;
component of the total (sv) and effective (s’v) stresses can be Vardy et al. 2015), P-wave attenuation measurements (e.g.,
calculated by integrating the densities: Morgan et al. 2012, 2014; Cevatoglu et al. 2015), P- and S-wave
attenuations (e.g., Lei and Morgan 2015), changes in Poisson’s
(4) ratio from FW inversion (Provenzano et al. 2016), and a variety
of downhole logging tools (e.g., Lu and McMechan 2002; Lee
(5) 2004).
Figure 6 shows two examples of these techniques. Figure 6a
where g is the gravitational loading and ρg is the grain density. and 6b shows a Chirp seismic reflection section from the latter
Vardy et al. (2015) demonstrated the application of this tech- stages of a time-lapse controlled CO2 release experiment
nique to the Finneidfjord dataset shown in Figure 4. The top (Blackford et al. 2014; Cevatoglu et al. 2015; Lichtschlag et al.
10 m of subsurface was characterized as being under-consolidat- 2015) where the vertical migration of CO2 injected into the bot-
ed (probably due to a combination of rapid deposition and high tom of the soft sediments can be traced. Polarity reversals, high-
clay content; therefore, low permeability), with a concentrated amplitude reflections, and attenuation anomalies (amplitude
thin bed of excess pore pressure associated with the complex loss, blanking, and frequency distortion) can be observed at a
clay/sand bed at c. 3.5-m depth (Fig 4). This demonstrated good variety of stratigraphic levels, including at the SF. These can be
agreement with local CPT data. In Figure 5c, we apply the same quantified using reflection coefficients calculated for the SF and
technique to the Solent data, identifying the over-consolidated reflector H2, as well as Q between the SF and H2 (Figure 5b;
nature (i.e., pore pressure less than hydrostatic) of the dipping Cevatoglu et al. 2015). Figure 6c shows a section of pre-stack
siltstone and mudstone layers. Alternatively, similar results can time-migrated Sparker multichannel data from Finneidfjord,
be obtained using burial models constrained by the seismic data Norway, overlain by gas saturation estimates derived from
(e.g., Marin-Moreno, Minshull and Edwards 2013). inverted interval velocity changes using the Wood’s equation

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 396 12/07/17 09:57


Remote characterization of shallow marine sediments 397

(Leighton and Robb 2008; Vardy et al. 2015), whereas Figure 6d gassy sediments (e.g., Toth et al. 2015). However, if geophysical
shows Poisson’s ratio curves for inside and outside the gas front measurements can be made across a broad-enough bandwidth,
derived using an FW inversion (Provenzano et al. 2016). The this frequency-dependent response can be highly diagnostic
Poisson’s ratio change is a particularly intuitive gas indicator as (Marin-Moreno, Sahoo and Best 2017).
the decrease from 0.485 to 0.470 (which is significant) reflects
the increased sensitivity of P-waves to small gas saturations Strength and stiffness
through increased compressibility. While the strength and/or stiffness of a soil can be geologically
The limitation of all of these techniques is that they initially interesting in understanding the depositional history of a strati-
quantify the gas in terms of the acoustic and/or elastic signature, graphic unit (such as levels of ice loading), for geotechnical
which requires relating to a gas saturation using a rock physical applications, it is a fundamental component to a huge range of
model. There are a large number of published rock physical applications, from foundation design to predicting trenching
models for gassy sediments (e.g., Hamilton 1972; White 1975; resistance during cable/pipeline laying. When more advanced
Mavko and Nur 1979; Winkler and Nur 1979; Carcione and inversion techniques are used and S-wave velocity is derived,
Picotti 2006; Rossi et al. 2007), many of which require a large this offers a direct dependence on various basic geotechnical
number of parameters (e.g., porosity, permeability, grain bulk strength/stiffness parameters (e.g., undrained shear strength and
and shear moduli, grain density, temperature, and pressure) that plasticity index) through the shear modulus. Although there are
cannot be easily constrained. This is particularly true at the fre- no published soil mechanical relationships, there are numerous
quencies used for high- and ultra-high-resolution seismic reflec- empirically derived relationships based on terrestrial studies
tion profiling (broadly, 0.4–20.0 kHz), where the seismic wave- (e.g., L’Heureux and Long 2016) and some with marine sedi-
lengths (3.750–0.075 m at 1500 m/s) are approaching the same ments (Ayres and Theilen 1999).
scale as the gas inclusions and therefore the topology and distri- However, as previously discussed, S-wave measurements of
bution of the inclusions become important (e.g., Hart and the marine near-surface data are rare (see Table 2). Even when
Hamilton 1993; Mathys et al. 2005). Bubble resonance, for an acquisition geometry/method is used whereby S-wave veloc-
example, can lead to a significant underestimate of the gas satu- ities can be derived, the uncertainties compared to P-wave
ration due to a reduction in the apparent velocity change of the velocities are often very high (Provenzano et al. 2017).

Figure 6 Panel (a) shows a single-channel Chirp profile from a time-lapse-monitored and controlled CO2 release experiment study in Oban, UK, where
a range of gas accumulation and migration features can be observed. Red circle indicates the location of gas injection. Panel (b) shows reflection coef-
ficients calculated for the SF and reflection H2. Panel (c) shows a section of pre-stack time-migrated Boomer multichannel data from Finneidfjord,
Norway, overlain by gas saturation estimates made using inverted interval velocities and Wood’s equation. Panel (d) compares Poisson’s ratio estimates
made inside (red) and outside (black) the gas front using FW inversion (after, Provenzano et al. 2016).

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 397 12/07/17 09:57


398 M.E. Vardy et al.

Figure 7 Panel (a) shows a short


section of single-channel Boomer
seismic reflection data from
Walney, UK, that intersects a CPT
location. Panel (b) cross-plots the
CPT measured tip resistance (qt)
against the acoustic impedance
inverted using the Vardy (2015)
algorithm. Panels (c) and (d) com-
pare the CPT measured qt against
qt values predicted using an ANN.

While P-wave properties such as P-wave velocity and imped- Figures 7c and 7d present the results from using an artificial
ance can be expected to demonstrate some relationship to the neural network (ANN) to perform multi-attribute regression and
soil strength/stiffness, it will be a much more complex one that predict CPT measurements from a range of geophysical inputs.
is both site-specific and, in all likelihood, facies-specific. Figure 7c cross-plots the measured qt versus the qt predicted by
Nauroy et al. (1998) attempted to derive such a relationship by the ANN, showing strong clustering of the data around the 1:1
cross-plotting P-wave velocity and impedance with CPT tip line of perfect fit. Figure 7d compares the measured qt (red line)
resistance (qt). These results show a lot of scatter (probably a and predicted qt for the CPT profile location indicated on Figure
reflection of facies dependence) but do demonstrate enough of 7a. Broadly, these data demonstrate an excellent agreement in
a relationship between the geophysical and geotechnical param- both metre- and sub-metre-scale structures and therefore present
eters to warrant continued interest. one possible solution for deriving soil strength/stiffness informa-
Figure 7 cross-plots a range of geophysical parameters tion from P-wave seismic reflection data.
against qt using single-channel Boomer data from Walney, UK.
To produce the cross-plots, seismic trace data coincident with a DISCUSSION AND CONCLUSIONS
CPT test location (Figure 7a) have been inverted for acoustic In reviewing the published literature regarding quantitative char-
impedance using the genetic algorithm inversion of Vardy (2015) acterization of the marine shallow subsurface, we have demon-
and the resulting impedance model used to derive a high-resolu- strated that a large range of different techniques have success-
tion interval velocity model to convert all the geophysical data fully been applied to derive various parameters, including imped-
(impedance and trace amplitudes) from travel time to depth ance, velocity, Q, bulk density, and Poisson’s ratio. By collating
below SF. Figure 7b plots the acoustic impedance (colour coded specific examples from a range of different sites in different
by seismic facies) against qt. These results demonstrate enough environments and acquired using different sources and source/
coherency to support a relationship between soil strength/stiff- receiver arrays, we have explored the practical limitations of
ness and acoustic impedance, but that this can be different for what can and cannot be derived from high- and ultra-high-reso-
different stratigraphic units (notably, facies 5 is markedly dis- lution geophysical data in terms of five key subsurface parame-
tinct): and the change in the acoustic properties of the soils terisations.
between difference facies is much less pronounced than the • Lithological characterization of the shallow subsurface is dif-
changes in qt, with significantly less facies segregation on the ficult using seismic reflection data, which commonly only
Y-axis of Figure 7b compared to the X-axis (notably, facies 3). records the P-wave response. Where Q-estimation can be per-
Any mechanism to derive a relationship between the P-wave formed, it is possible to broadly distinguish between granular
response and geotechnical measurements (such as qt) therefore and cohesive materials, but more detailed lithological informa-
needs to be capable of handling a complex multi-parameter tion can only reliably be derived using the S-wave response of
dependence on acoustic property, facies, and geolocation. the sediments.

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 398 12/07/17 09:57


Remote characterization of shallow marine sediments 399

• P-wave seismic reflection data, however, demonstrate high specific, “true” value of a property. Realistically, due to the
sensitivity to variations in porosity, making both qualitative complexity of the problems being solved (often being severe-
(using velocity and Q) and/or quantitative (using bulk density ly under-parameterised) and the significant noise contamina-
and porosity) analysis of the in situ stress regime possible. tion of high- and ultra-high-resolution geophysical data, it
• Numerous examples have shown that seismic reflection data will perhaps never be possible to derive the “true” in situ
are highly sensitive to the nature of the pore fluids and will value for the desired property. It is therefore more pragmatic
demonstrate noticeably anomalous amplitude responses even to develop and apply a more thorough parameter sensitivity
in the presence of small quantities of free gas. A number of analysis to quantify realistic errors/confidence limits on
different inversion schemes (velocity, Q, and FW) have been derived parameters, which is also precisely the kind of sub-
used to derive estimates for the in situ gas saturation, produc- surface parameterisation of most use to geologists and geo-
ing sensible numbers. However, due to a lack of reliable technical engineers.
ground truth, it is difficult to comment on the accuracy of these iv. S-wave characterization: S-waves offer a more direct route
methods beyond their self-consistency. into quantifying both the lithology and also the fundamental
• Several methods have been shown that offer the potential for geotechnical properties such as undrained shear strength.
deriving geotechnical strength/stiffness parameters such as However, the challenges in reliably generating and recording
plasticity and undrained shear strength. Where S-wave data S-waves in the shallow marine environment mean they are
can be derived, there are existing empirical relationships, currently underused and therefore offer significant potential
whereas it has been shown that computationally advanced for improvement.
multi-attribute regression techniques can be used to predict
CPT measurements based on P-wave velocity information on ACKNOWLEDGEMENTS
a site- and facies-specific basis. The authors wish to thank Kerry Campbell, Ranajit Ghose and an
anonymous reviewer for their useful comments, which greatly
Based on these results, we identify four research questions/ave- improved the quality of the manuscript. The field data presented
nues critical to developing a better integration between geologi- was processed using a combination of Landmark’s ProMAX
cal, geotechnical, and geophysical site characterizations. software and Seismic Unix, while the inversion results used
i. Soil mechanical models: Currently, the derivation of subsur- custom-written algorithms. Clare was supported by the NERC
face parameters from marine geophysical data involves a Environmental Risks to Infrastructure Innovation Programme
two-stage process inversion of the geophysical data in terms (NE/N012798/1).
of seismic wavefield properties (e.g., impedance, velocity,
density, and Q), followed by the calculation of more useful REFERENCES
sediment properties (e.g., lithology, effective stress, and und- Aleardi M., Tognarelli A. and Mazzotti A. 2016. Characterisation of shal-
rained shear strength), often using empirically derived rela- low marine sediments using high-resolution velocity analysis and
genetic algorithm-driven 1D elastic full-waveform inversion. Near
tionships. Fundamentally, there is a limit as to how far Surface Geophysics 14(5), 449–460.
empirical relationships, be they globally or site specifically Allouche N., Drijkoningen G.G. and van der Neut J. 2010. Methodology
derived, can be trusted to be reliable, and there is therefore a for dense spatial sampling of multicomponent recording of converted
need to develop a better physical understanding of the rela- waves in shallow marine environments. Geophysics 75(6), WB29–
tionships between a range of seismic wavefield and sediment WB37.
Allouche N., Drijkoningen G.G., Versteeg R. and Ghose R. 2011.
properties. This is particularly true for quantifying gas satura- Converted waves in a shallow marine environment: experimental and
tion, where better calibrated soil mechanical models would modeling studies, Geophysics, 76(1), P.T1-T11.
significantly aid application. Ayres A. and Theilen F. 1999. Relationship between P- and S-wave
ii. Single-stage inversion: The development of improved soil velocities and geological properties of near-surface sediments of the
mechanical models could also offer the possibility of reduc- continental slope of the Barents Sea. Geophysical Prospecting 47,
431–441.
ing the current two-stage inversion process to a single inver- Ballard M., Becker K.M. and Goff J.A. 2010. Geoacoustic inversion for
sion explicitly cast in terms of the most useful sediment the New Jersey shelf: 3D sediment model. IEEE Journal of Oceanic
properties (e.g., lithology, effective stress, and undrained Engineering 35(1), 28–42.
shear strength). Holland and Dettmer (2013) demonstrate the Bell M. and Walker M.J. 2005. Late Quaternary Environmental Change:
potential of this approach using the current Biot and viscous Physical and Human Perspectives. Pearson Education.
Blackford J., Stahl H., Bull J.M., Berges B.J.P., Cevatoglu M., Lichtschlag
grain shearing models, but there is still significant scope for A. et al. 2014. Detection and impacts of leakage from sub-seafloor
improvement. deep geological carbon dioxide storage. Nature Climate Change 4,
iii. Property envelopes: Historically, the majority of quantitative 1011–1016.
geophysical characterization of the marine shallow subsur- Bohlen T., Kugler S., Klein G. and Theilen F. 2004. 1.5D inversion of
face has been on either explicitly calculating (e.g., Q and/or lateral variation of Scholte-wave dispersion. Geophysics 69(2), 330–
344.
reflection coefficients) or inverting for (e.g., impedance) the

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 399 12/07/17 09:57


400 M.E. Vardy et al.

Bosch M., Mukerji T. and Gonzalez E. 2010. Seismic inversion for res- Holbrook W.S., Hoskins H., Wood W.T., Stephen R.A. and Lizarralde D.
ervoir properties combining statistical rock physics and geostatistics: 1996. Methane hydrate and free gas on the Blake Ridge from vertical
a review. Geophysics 75, A165–A176. seismic profiling. Science 273, 1840–1843.
Bull J.M., Quinn R. and Dix J.K. 1998. Reflection coefficient calculation Holland C. and Dettmer J. 2013. In situ sediment dispersion estimates in
from marine high resolution seismic reflection (Chirp) data and appli- the presence of discrete layers and gradients. The Journal of the
cation to an archaeological case study. Marine Geophysical Researches Acoustical Society of America 133, 50–63.
20, 1–11. Huws D.G., Davies A.M. and Pyrah J.R. 2000. A nondestructive tech-
Campbell K.J. 1984. Predicting offshore soil conditions. Proceedings - nique for predicting the in situ void ratio for marine sediments. Marine
Offshore Technology Conference, Houston, TX, OTC Paper 4692, pp. Georesources and Geotechnology 18(4), 333–346.
391–398. Ker S., Gonidec Y.L. and Gilbert D. 2013. Multiresolution seismic
Carcione J.M. and Picotti S. 2006. P-wave seismic attenuation by slow- data fusion with a generalized wavelet-based method to derive sub-
wave diffusion: effects of inhomogeneous rock properties. Geophysics seabed acoustic properties. Geophysical Journal International 195,
71, O1–O8. 1370.
Cevatoglu M., Bull J.M., Vardy M.E., Gernon T.M., Wright I.C. and L’Heureux J.-S. and Long M. 2016. Correlations between shear wave
Long D. 2015. Gas migration pathways, controlling mechanisms and velocity and geotechnical parameters in Norwegian clays. In:
changes in acoustic properties observed in a controlled sub-seabed Proceedings of the 17th Nordic Geotechnical Meeting, Reykjavik,
CO2 release experiment. International Journal of Greenhouse Gas Iceland.
Control 38, 26–43. L’Heureux J.-S., Longva O., Steiner A., Hansen L., Vardy M., Vanneste
Cheng P. and Margrave G.F. 2012. A match-filter method for Q-estimation. M. et al. 2012. Identification of weak layers and their role for the
2012 SEG annual meeting, Expanded Abstracts. stability of slopes at Finneidfjord, northern Norway. In: Submarine
Chotiros N.P. 1994. Reflection and reverberation in normal incidence Mass Movements and Their Consequences, Advances in Natural and
echo-sounding. The Journal of the Acoustical Society of America Technological Hazards Research, Vol. 31, pp. 321–330. Heidelberg,
96(5), 2921–2929. Germany: Springer.
Clare M.A., Talling P.J., Challenor P., Malgesini G. and Hunt J. 2014. Lee M.W. 2004. Elastic velocities of partially gas-saturated unconsoli-
Distal turbidites reveal a common distribution for large (>0.1 km3) dated sediments. Marine and Petroleum Geology 21, 641–650.
submarine landslide recurrence. Geology 42(3), 263. Lei X. and Morgan E.C. 2015. Characterization of gas-charged sedi-
Clayton C. 2011. Stiffness at small strain: research and practice. ments from joint inversion of Qp and Qs. Proceedings of the SEG
Géotechnique 61(1), 5–37. Annual Meeting, 2765–2770.
Clayton C.R.I., Hight D.W. and Hopper R.J. 1992. Progressive destruc- Leighton T. and Robb G. 2008. Preliminary mapping of void fractions
turing of Bothkennar clay: implications for sampling and reconsolida- and sound speeds in gassy marine sediments from subbottom profiles.
tion procedures. Géotechnique 42(2), 219–239. JASA Express Letters 124(5), EL313–EL320.
Dosso S.E., Dettmer J., Steininger G. and Holland C.W. 2014. Efficient Lichtschlag A., James R.H., Stahl H. and Connelly D. 2015. Effect of a
trans-dimensional Bayesian inversion for geoacoustic profile estima- controlled sub-seabed release of CO2 on the biogeochemistry of shal-
tion. Inverse Problems 30. low marine sediments, their pore waters, and the overlying water col-
Drijkoningen G.G., el Allouche N., Thorbecke J. and Bada G. 2012. umn. International Journal of Greenhouse Gas Control 38, 80–92.
Nongeometrically converted shear waves in marine streamer data. Lunne T. 2012. The Fourth James K. Mitchell Lecture: The CPT in off-
Geophysics 77(6), P45–P56. shore soil investigations - a historic perspective. Geomechanics and
Evans T. 2011. A systematic approach to offshore engineering for multi- Geoengineering 7(2), 75–101.
ple-project developments in geohazardous areas. In: Frontiers in Lu S.M. and McMechan G.A. 2002. Estimation of gas hydrate and free
Offshore Geotechnics II (ed D. White), pp. 3–32. CRC Press. gas saturation, concentration, and distribution from seismic data.
Gardner W.D., Richardson M.J. and Cacchione D.A. 1989. Geophysics 67, 582–593.
Sedimentological effects of strong southward flow in the Straits of Ma X.-Q. 2002. Simultaneous inversion of prestack seismic data for rock
Florida. Marine Geology 86, 155–180. properties using simulated annealing. Geophysics 67(6), 1877–1885.
Gerstoft P. 1994. Inversion of seismoacoustic data using genetic algo- Mallick S. 2001. Prestack waveform inversion using a genetic algorithm
rithms and a posteriori probability distributions. The Journal of the - The present and the future. CSEG Recorder 26(6), 78–84.
Acoustical Society of America 95(2), 770–782. Marin-Moreno H., Minshull T.A. and Edwards R.A. 2013. Inverse mod-
Guo Y.-G., Li F.-H., Liu J.-J. and Li Z.-L. 2006. Time-domain geoacous- elling and seismic data constraints on overpressure generation by
tic inversion based on normal incidence reflection from layered sedi- disequilibrium compaction and aquathermal pressuring: application to
ment. Chinese Physics Letters 23(9), 2483–2486. the Eastern Black Sea Basin. Geophysical Journal International 194,
Gutowski M., Bull J., Henstock T., Dix J., Hogarth P., Leighton T. et al. 814–833.
2002. Chirp sub-bottom profiler source signature design and field test- Marin-Moreno H., Sahoo S.K. and Best A.J. 2017. Theoretical modelling
ing. Marine Geophysical Researches 23, 481–492. insights into elastic wave attenuation mechanisms in marine sediments
Hamilton E.L. 1972. Compressional-wave attenuation in marine sedi- with pore-filling methane hydrate. Journal Geophysical Research,
ments. Geophysics 37, 620–646. Solid Earth, 122, 1835–1847, doi:10.1002/2016JB013577.
Hamilton E.L. and Bachman R.T. 1982. Sound velocity and related prop- Masson D.G., Wynn R.B. and Talling P.J. 2010. Large landslides on pas-
erties of marine sediments. The Journal of the Acoustical Society of sive continental margins: processes, hypotheses and outstanding ques-
America 72, 1891–1904. tions. In: Submarine Mass Movements and Their Consequences, pp.
Hart B.S. and Hamilton T.S. 1993. High resolution acoustic mapping of 153–165. Springer Netherlands.
shallow gas in unconsolidated sediments beneath the Strait of Georgia, Mathys M., Thiessen O., Theilen F. and Schmidt M. 2005. Seismic char-
British Columbia. Geo-Marine Letters 13, 49–55. acterisation of gas-rich near surface sediments in the Arkona Basin,
Hilton R.G., Galy A., Hovius N., Horng M.J. and Chen H. 2011. Efficient Baltic Sea. Marine Geophysical Researches 26, 207–224.
transport of fossil organic carbon to the ocean by steep mountain rivers: Mavko G.M. and Nur A. 1979. Wave attenuation in partially saturated
an orogenic carbon sequestration mechanism. Geology 39(1), 71–74. rocks. Geophysics 44, 161–178.

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 400 12/07/17 09:57


Remote characterization of shallow marine sediments 401

Mayall M., Jones E. and Casey M. 2006. Turbidite channel reservoirs— Schock S., LeBlanc L. and Mayer L. 1989. Chirp subbottom profiler for
Key elements in facies prediction and effective development. Marine quantitative sediment analysis. Geophysics 54(4), 445–450.
and Petroleum Geology 23(8), 821–841. Scourse J.D., Haapaniemi A.I., Colmenero-Hidalgo E., Peck V.L., Hall
Morgan E., Vanneste M., Lecomte I., Baise L., Longva O. and McAdoo I.R., Austin W.E. et al. 2009. Growth, dynamics and deglaciation of
B. 2012. Estimation of free gas saturation from seismic reflection the last British-Irish ice sheet: the deep-sea ice-rafted detritus record.
surveys by the genetic algorithm inversion of a P-wave attenuation Quaternary Science Reviews 28, 3066–3084.
model. Geophysics 77(4), R175–R187. Seong W. and Park C. 2001. Geoacoustic inversion via genetic algorithm
Morgan E., Vanneste M. and Vardy M.E. 2014. Characterisation of the and its application to manganese sediment identification. Marine
slope destabilizing effects of gas charged sediments via seismic Georesources and Geotechnology 19(1), 37–51.
surveys. Offshore Technology Conference, 14OTC-P-963-OTC. Sobreira J.F.F., Lipski M., Carvalho L.A. and Marquez E. 2010.
Nauroy J.-F., Dubois J.-C., Colliat J.-L., Kervadec J.-P. and Meunier J. Geotechnical characterization based on seismic data approaches
1998. The GEOSIS Method for Integrating VHR Seismic and applied in Campos Basin, southeastern Brazilian Margin. The Leading
Geotechnical Data in Offshore Site Investigations. Offshore Site Edge 29(7), 842–846.
Investigation and Foundation Behaviour - New Frontiers: Socco V.L., Boiero D., Maraschini M., Vanneste M., Madshus C.,
Proceedings of an International Conference, pp. 175–198. Westerdahl H. et al. 2011. On the use of the Norwegian Geotechnical
Ohta K., Okabe K., Morishita I., Ozaki S. and Frisk G.V. 2005. Institute’s prototype seabed-coupled shear wave vibrator for shallow
Inversion for seabed geoacoustic properties in shallow water experi- soil characterization –II. Joint inversion of multimodal Love and
ments. Acoustical Science and Technology 26(4), 326–337. Scholte waves. Geophysical Journal International 185,
Panda S., LeBlanc L. and Schock S. 1994. Sediment classification 237–252.
based on impedance and attenuation estimation. The Journal of the Solheim A., Bryn P., Sejrup H.P. and Berg K. 2005. Ormen Lange—an
Acoustical Society of America 96(5), 3022–3035. integrated study for the safe development of a deep-water gas field
Pinson L. 2010. Derivation of acoustic and physical properties from within the Storegga Slide Complex, NE Atlantic continental margin;
high-resolution seismic reflection data. Unpublished PhD thesis, executive summary. Marine and Petroleum Geology 22, 1–9.
University of Southampton, UK. Sternlicht D.D. and de Moustier C.P. 2003. Remote sensing of sediment
Pinson L., Henstock T., Dix J. and Bull J. 2008. Estimating quality fac- characteristics by optimized echo-envelope matching. The Journal of
tor and mean grain size of sediments from high-resolution marine the Acoustical Society of America 114(5), 2727–2743.
seismic data. Geophysics 73(4), G19–G28. Stevenson I., McCann C. and Runciman P. 2002. An attenuation-based
Pinson L., Vardy M., Dix J., Henstock T., Bull J. and Maclachlan S. sediment classification technique using Chirp sub-bottom profiler data
2013. Deglacial history of glacial lake Windermere, UK: implica- and laboratory acoustic analysis. Marine Geophysical Researches 23,
tions for the central British and Irish Ice Sheet. Journal of 277–298.
Quaternary Science 28(1), 83–94. Stoker M., Bradwell T., Howe J., Wilkinson I. and McIntyre K. 2009.
Power P., Clare M., Rushton D. and Rattley M. 2011. Reducing geo- Lateglacial ice-cap dynamics in NW Scotland: evidence from the
risks for offshore developments. In: Geotechnical Safety and Risk fjords of the Summer Isles region. Quaternary Science Reviews 28,
(eds N. Vogt, B. Schuppener, D. Straub and G. Bräu), pp. 217–224: 3161–3184.
3rd International Symposium on Geotechnical Risk and Safety. Stoker M. and Bradwell T. 2009. Neotectonic deformation in a Scottish
Bundesanstalt fur Wasserbau. fjord, Loch Broom, NW Scotland. Scottish Journal of Geology 45(2),
Priest J., Druce M., Roberts J., Schultheiss P., Nakatsuka Y. and Suzuki 107–116.
K. 2015. PCATS Triaxial: a new geotechnical apparatus for charac- Stow D.A. 2005. Sedimentary Rocks in the Field: A Color Guide. Gulf
terizing pressure cores from the Nankai Trough, Japan. Marine and Professional Publishing.
Petroleum Geology 66(2), 460– 470. Terzaghi K. 1925. Principles of soil mechanics. Engineering News-
Provenzano G., Vardy M.E. and Henstock T.J. 2016. Pre-stack wave- Record 95, 19–27.
form inversion of VHF marine seismic reflection data – A case study Terzaghi K. 1936. Simple tests determine hydrostatic uplift. Engineering
in Norway. Near Surface Geoscience 2016 – Second Applied News-Record 116(25), 872–875.
Shallow Marine Geophysics Conference. Toth Z., Spiess V., Mogollon J.M. and Jensen J.B. 2014. Estimating the
Provenzano G., Vardy M.E. and Henstock T.J. 2017. Pre-stack full free gas content in Baltic Sea sediments using compressional wave
waveform inversion of ultra-high-frequency marine seismic reflec- velocity from marine seismic data. Journal of Geophysics: Solid
tion data. Geophysical Journal International 209, 1593–1611.. Earth.
Richardson M.D. and Briggs K.B. 1993. On the use of acoustic imped- Toth Z., Spiess V. and Kiell H. 2015. Frequency dependence in seismoa-
ance values to determine sediment properties. Proceedings of the coustic imaging of shallow free gas due to gas bubble resonance.
Institute of Acoustics 15(2), 15–24. Journal of Geophysics: Solid Earth.
Robb G.B.N., Best A.I., Dix J.K., Bull J.M., Leighton T.G. and White Tucker M.E. 2003. Sedimentary Rocks in the Field. John Wiley & Sons.
P.R. 2006. The frequency dependence of compressional wave veloc- Vanneste M., Madshus C., Socco V.L., Maraschini M., Sparrevik H.,
ity and attenuation coefficient of intertidal marine sediments. The Westerdahl K. et al. 2011. On the use of the Norwegian Geotechnical
Journal of the Acoustical Society of America 120(5), 2526– Institute’s prototype seabed-coupled shear wave vibrator for shal-
2537. low soil characterization –I. Acquisition and processing of multi-
Rossi G., Gei D., Böhm G., Madrussani G. and Carcione J.M. 2007. modal surface waves. Geophysical Journal International 185,
Attenuation tomography: an application to gas-hydrate and free-gas 221–236.
detection. Geophysical Prospecting 55, 655–669. Vanneste M., L’Heureux J.-S., Brendryen J., Baeten N., Larberg J., Vardy
Rothwell R.G. and Rack F.R. 2006. New techniques in sediment core M. et al. 2012. Assessing offshore geohazards: a multi-disciplinary
analysis: an introduction. Geological Society of London Special research initiative to understand shallow landslides and their dynamics
Publications 267(1), 1–29. in coastal and deepwater environments, Norway. In: Submarine Mass
Sadler P.M. 1981. Sediment accumulation rates and the completeness Movements and Their Consequences, Advances in Natural and
of stratigraphic sections. The Journal of Geology 569–584. Technological Hazards Research, Vol. 31 (eds Y. Yamada, K.

© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 401 12/07/17 09:57


402 M.E. Vardy et al.

Kawamura, K. Ikehara, Y. Ogawa, R. Urgeles, D. Mosher et al.), pp. Wagner C., Gonzalez A., Agarwal V., Koesoemadinata A., Ng D., Trares
29–41. Heidelberg, Germany: Springer. S. et al. 2012. Quantitative application of poststack acoustic imped-
Vanneste M., Sultan N., Garziglia S., Forsberg C.F. and L’Heureux J.-S. ance inversion to subsalt reservoir development. The Leading Edge
2014. Seafloor instabilities and sediment deformation processes: The 31(5), 528– 537.
need for integrated, multi-disciplinary investigations, Marine Geology Wang J. and Stewart R. 2015. Inferring marine sediment type using Chirp
352, p. 183–214. sonar data: Atlantis field, Gulf of Mexico. Proceedings of the SEG
Vanneste M., Forsberg C., Knudsen S., Kvalstad T., L’Heureux J.-S., Annual Meeting, 2385–2390.
Lunne T. et al. 2015. Integration of very high-resolution seismic and White J. 1975. Computed seismic speeds and attenuation in rocks with
CPTU data from a coastal area affected by shallow landsliding–the partial gas saturation. Geophysics 40, 224–232.
Finneidfjord natural laboratory. Proceedings of the Annual Offshore Widess M.B. 1982. Quantifying the resolving power of seismic systems.
Technology Conference, OTC-TC-P-686. Geophysics 47(8), 1160–1173.
Vardy M. 2015. Deriving shallow-water sediment properties using post- Winkler K. and Nur A. 1979. Pore fluids and seismic attenuation in
stack acoustic impedance inversion. Near Surface Geophysics 13(2), rocks. Geophysical Research Letters 6, 1–4.
143–154. Winsborrow G., Huws D.G. and Muyzert E. 2005. The estimation of
Vardy M., L’Heureux J.-S., Vanneste M., Longva O., Steiner A., Forsberg shear-wave statics using in situ measurements in marine near-surface
C. et al. 2012. Multidisciplinary investigation of a shallow near-shore sediments. Geophysical Prospecting 53, 557–577.
landslide, Finneidfjord, Norway. Near Surface Geophysics 10, 267–277. Zhang A. and Digby A. 2013. Analysis of amplitude, reflection strength,
Vardy M., Vanneste M., Henstock T., Morgan E. and Pinson L. 2015. Can and acoustic impedance of AUV sub-bottom profiles with application
high-resolution marine geophysical data be inverted for soil proper- to deepwater near-surface sediments. Proceedings of the Annual
ties? Proceedings of the Institute of Acoustics 37(1), 149–156. Offshore Technology Conference OTC-23978.

Knowledge
empowers
you
www.earthdoc.org Over 63,000 papers are now available online!

EDC17 V2H.indd 3 12/07/17 09:54


© 2017 European Association of Geoscientists & Engineers, Near Surface Geophysics, 2017, 15, 387-402

19693-NSG17 August BOOK.indb 402 12/07/17 09:57

View publication stats

You might also like