Вы находитесь на странице: 1из 94

FUGRO – CITEPH GASSY SOILS – FINAL REPORT

RESEARCH PROJECT
CITEPH - GASSY SOILS

FINAL REPORT

Fugro Report No. CITEPH-REV04-23/06/2016


Digital Report Image The PDF document file held in Fugro’s archive represents Fugro’s formal
® ®
deliverable to the Client. It is designed for viewing with Acrobat Reader Version
®
9 operating under Windows XP
Confidentiality Report distribution restricted to project participants on CITEPH research project

Abstract

Gas-charged sediments are widely distributed throughout the world’s oceans and are considered a major
hazard by the oil and gas industry. Major oil companies have recently been faced with important engineering
problems related to the presence of free gas in marine sediments.
Research into gas-charged sediments has also gained renewed interest in close relation to environmental
issues and geological hazard assessment associated with gas release in marine sediments.

The gases found in shallow marine sediments (mainly methane) can originate from biogenic or thermogenic
processes. Gas can occur in the seabed in three ways: dissolved, undissolved in the form of gas-filled voids,
or as clathrates.

Free gas-filled voids, exsolution of dissolved gas and/or gas releases generated by gas hydrates dissociation
may all have significant implications for the behaviour and engineering properties of their host marine
sediments. While the behaviour of gassy sands has been characterised experimentally by various authors
data describing the behaviour of gassy clays are still limited, mainly because of the time demand of
experiments associated with the low permeability of such fine materials.

The document is the final report of CITEPH – GASSY SOIL research project.

Issue Report Status Prepared Checked Approved Date

01 Complete Report NDR SPO SPO 23/06/2016


FUGRO – CITEPH GASSY SOILS – FINAL REPORT

CONTENTS

1. INTRODUCTION 2

2. THE SIGNIFICANCE OF GAS FOR OFFSHORE OPERATIONS 3


2.1 Introduction 3
2.2 Structures of gassy soils 3
2.3 Physical characteristics of gassy soils 4
2.4 Fluid flow: gas and water 7
2.5 Undrained strength, compressibility and shear stiffness 7
2.6 Gas evidence 9
2.6.1 From seismic data 9
2.6.2 From drilling operations 9
2.6.3 From sample observation 11
2.6.4 From PCPT records 13
2.6.5 From laboratory test results 17

3. LABORATORY AND IN SITU TESTING 19


3.1 Experimental set-up 19
3.2 Index properties of the studied clays 20
3.3 Experimental test programme 21
3.4 Test results 26

4. MECHANICAL MODELS FOR GASSY SOILS 28


4.1 On the behaviour of gassy soils 28
4.1.1 Gas bubbles and structures of gassy soils 28
4.1.2 Consolidation behaviour of gassy soils 31
4.1.3 Elastic moduli of gassy soils 35
4.1.4 Undrained shear strength of gassy soils 38
4.1.5 3.5 Drained shear strength of gassy soils 46
4.1.6 Unloading behaviour of gassy soils 47
4.1.7 Models accounting for the behaviour of gassy soils 58

5. COUPLING SATURATED AND GAS EFFECTS MODELS AND PARAMETER IDENTIFICATION 67


5.1 Introduction 67
5.2 Modelling structure and its degradation: S3-SKH model 68
5.3 Modelling the effects of rate of strain: RASTRA model 73
5.4 Coupling RASTRA model and S3-SKH model 74
5.5 Effects of structure on the viscous response 75
5.6 Effects of strain rate on soil structure 76
5.7 Implementation of the models in a FEM code 78
5.8 Numerical results 80

6. CONCLUSIONS 83

7. REFERENCES 85

8. LISTE OF FIGURES 91

1 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

1. INTRODUCTION
The presence of free gas in marine sediments has for long been recognized in offshore oil and gas
exploration. “Shallow gas” can be detected by seismic reflexion surveys as causing the well known
phenomena called “acoustic masking” or “acoustic blanking”.

Until recently zones where presence of shallow gas was evidenced at an early stage were considered
at high risk by operators: drilling activities were as far as possible avoided and field developments
were disregarded. The move to deep water exploration and the rarefaction of oil and gas reserves
have progressively changed the attitude of the operators which consider more and more frequently the
development of fields in a gassy soil environment. Two main categories of risks are associated to oil
and gas developments in gassy soils:

- Operational risks (e.g. blow out during drilling): stringent HSE requirements must be put in place,

- Geotechnical risks: the effect of gas on the mechanical properties of the sediments should be
evaluated and their impact on foundation design taken into account.

Major oil companies recognize that i) the geological risk assessment associated to gas release in
marine sediments is still not reliable, and ii) the impact of free gas on the mechanical behavior of
marine sediments is poorly understood. Substantial R&D efforts are required.

Three major actions have been recently or are being conducted:

a Joint Industry Project sponsored by BP, EXXON-MOBIL and TOTAL dedicated to the acoustical and
mechanical characterization of gassy soil sediments;

a four-year multi-partner R&D project sponsored by CITEPH and entitled “Assessment of geotechnical
risks associated to the presence of gas in marine sediment”; the development of the GEOPS
equipment aimed at obtaining a continuous log of the compressive (V p) and shear (Vs) wave velocity
during penetration of a CPT like - tool into the sediment.

The present report is final report of CITEPH project.

The final objective of these successive investigation phases is that a complete geological/geotechnical
model of the site can be proposed so that i) operational risks can be thoroughly assessed and ii)
design parameters for the foundation of the seabed structures can be determined.

It is now clearly established that geophysical data are by themselves insufficient to assess the risk
associated to gassy sediments. Acoustic masking is a “proof” that free gas has accumulated in the soil
under the form of gas pockets but conversely experience shows that free gas may exist (at least in
small quantity) without being detected by seismic measurements. Many other signs can be revealed
by geophysical data which can demonstrate a geological context prone to gas rich sediments.

2 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

2. THE SIGNIFICANCE OF GAS FOR OFFSHORE OPERATIONS

2.1 Introduction
Gas can occur in the seabed in three ways: in solution in the pore water; undissolved in the form of
gas-filled voids: or as clathrates (gas hydrates). In the first case, the gas will have very little physical
effect unless the ambient pressures are reduced (for example, by drilling or sampling), when the gas
will then come out of solution. ln the second case, the presence of the gas will affect the engineering
properties of the seabed, since the gas is highly compressible. In addition, in some circumstances,
there is a serious blow-out risk associated with drilling. Gas hydrates are crystalline compounds
composed of natural gas and water at low temperatures. Again, they become potentially hazardous
only when they melt, releasing free gas.

The purpose of the present paper is to examine the effect that undissolved gas may have on the
engineering properties of the seabed, from laboratory tests, and to present, in addition, any insights
from these tests that might have relevance to applications other th an foundation design. It draws
together a number of conclusions, mainly although not entirely from previously published papers,
presenting them without the detailed back- ground that is presented in the original publications.

The biogenic and petrogenic origins of seabed gases (including methane and carbon dioxide) have
been considered in Floodgate and Judd (1992). This paper is concerned with the overall physical
consequences of the pressures of gas and not with specific features attributable to the precise nature
of the gas. However, different gases have different solubilities, and this could influence the time scale
of observed behaviour, so that, except where otherwise noted, the tests described in this paper use
methane, the most common seabed gas, as the gas in the soil samples.

2.2 Structures of gassy soils


Soils may be divided into the broad categories of fine and coarse grained-i.e. clays and silty clays on
the one hand and sands on the other. A similar categorization is possible in terms of the effects of
undissolved gas, reflected in the different structures that are observed.

Fine-grained

Figure 2-1 shows a muddy sand sample from the Witch Ground pockmark area in the North Sea
(Hovland and Judd, 1988), in which isolated voids exist that were gas-filled in situ. An example of a
fine-grained soil is presented in Yuan et al. (1992). It shows a scanning electron microscope
photograph of a void in silty clay in the western Irish Sea. It may be seen that the sail particles forma
bridging structure around the void, with the size of the void being considerably larger than that of the
individual particles.

Soil samples recovered from the seabed undergo decompression. In a gassy sail, gas comes out of
solution, causing overall volume expansion and the possibility of structural damage. An indication of
the original sail structure may be obtained from such samples, but their value is limited in the context
of understanding fundamental behaviour. Improvements in sampling techniques have reduced the
severity of sampling damage by keeping the sail sample pressurized during the coring operation, and
basic testing can then be carried out by transferring the sample from the corer to the testing apparatus
3 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

in a hyperbaric chamber. An alternative approach for a fundamental study is to produce reconstituted


gassy soli samples in the laboratory. This has the advantage that samples are homogeneous, with
uniform distributions of gas and of a known stress history. It is, of course, essential that the structure of
such artificially produced gassy soils is basically similar to the sail structure in situ.

Fine-grained soil samples containing undissolved gas can be produced in the laboratory by biogenic
methods, in which methane producing bacteria are used, by chemical methods, in which gas is
produced as a reaction between appropriate compounds, or using a technique based on zeolite, an
inert chemical with a strong affinity for water (Nageswaran, 1983; Sills et al., 1991). In the last method,
methane saturated zeolite is added to a soil slurry and the methane is released over a period of some
hours, producing soil samples with physical characteristics similar to those of samples recovered from
the seabed.A zeolite produced gassy soil using an estuarine silty clay from Combwich, Somerset,
where the uniform distribution of gas bubbles is clearly visible. Electron micrographs of this soil
showed a very similar structure to that of the seabed sample presented (Yuan et al., 1992). The
majority of the laboratory tests on fine-grained samples reported in this paper have been carried out
using this soil with varying gas contents.

Gassy sands

The gas in purely coarse grained samples (i.e. sands without mud, in contrast to the muddy sand of
Figure 2-1) does not produce changes in the structure in the way that it does for the fine-grained
samples. lnstead, it displaces water from between sand grains. Laboratory samples of this form have
been produced by flushing water that has been saturated with methane through sand samples,
maintaining a back pressure of up to 1000 kPa. One of two sequences was then followed: either the
back pressure was reduced, bringing methane out of solution, or free gas was introduced under
pressure at one of the sample boundaries, with the pressure maintained until gas was seen to be
escaping at the other boundary (Sills, 1989).

2.3 Physical characteristics of gassy soils


Relative gas and water pressures

The differences between saturated and gassy soils are related to the different compressibilities, which
are in turn a function of the gas volumes and pressures. The gas pressure will depend on external
factors such as the supply of gas available and the sail structure and permeability. Whatever the
differences between these factors, where water and gas share a common interface, a meniscus must
exist, through which surface tension acts. The pore water pressure uw and the pore gas pressure ug
are then necessarily different, and related through the radius of curvature R of the gas-water interface
and the surface tension T, so that

2T
ug  u w  (2-1)
R

4 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 2-1. Photography of gassy soil sample.

Figure 2-2 illustrates the extreme conditions that can occur for the curvature of the menisci between
gas and pore water. In both cases, the menisci are concave to the gas, since bound water existing
around the soil grain or floc provides a water/water contact at the end of the meniscus, so that the
contact angle is zero.

Figure 2-2a shows the condition where the menisci span between individual particles. Assuming that
the pore water pressure is uniform around the void, it follows that each meniscus bas the same radius
of curvature. The minimum possible value for this radius-associated with the largest pressure
difference-will be determined by the particle spacing, or the size of the throat, between the particles.
The critical value Re corresponds to a meniscus that is just able to bridge the largest gap that exists
between the soil particles.

The magnitude of Reis of the order of the particle size and is not related to the bubble cavity
dimensions. At the surface of a wetted soil particle, where the contact angle of the meniscus must be
zero, the smallest possible gas pressure, which occurs for the maximum value of R, is given by the
bubble cavity radius a, as shown in Figure 2-2b. Therefore the surface tension limits for the gas
pressure are given by

2T 2T
uw   ug  u w  (2-2)
a Rc

If the gas pressure drops to the lower surface tension limit (u w + 2T/a), a further reduction is prevented
by a decrease in the gas volume caused by pore water moving into the cavity, resulting in partial or
complete flooding. At the upper surface tension limit, (uw + 2T/Rc), the gas pressure is prevented from
increasing further by a volume expansion, as gas moves out of the bubble cavity into the sail pores.

5 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Where the gas exists in the form of comparatively large voids, a change in gas pressure may also
result from expansion or contraction of the bubble cavity itself. This possibility of cavity expansion
provides an alternative set of limits for the gas pressure, with the maximum possible difference
between gas pressure ug and confining stress being a function of the sail strength (Wheeler et al.,
1990).

Silts and clays

In clays and silts the critical meniscus radius R c is very small. This means that the upper surface
tension limit for ug is much higher than the pore water pressure uw. However, the lower surface tension
limit for ug is only slightly greater than uw because, in clays and silts, the bubble radius, a, is typically
several orders of magnitude larger than the particle size, and the term 2T/a, is therefore, very small.

The large range between upper and lower surface tension limits for ug means that cavity expansion or
contraction normally occurs before movement of gas into the soil matrix or flooding of water into a gas-
filled cavity. The range of allowable gas pressure is therefore generally a function of soil strength
rather than particle size and arrangement.

Sands

In sands, the minimum meniscus radius Reis of the same order as the radius a of the gas-filled voids,
and the upper and lower surface tension limits for ug are very close together and bath only slightly
greater than uw. This small range between upper and lower surface tension limits means that
movement of gas into an adjacent water-filled void, or flooding of water into a gas-filled void, normally
occurs before cavity expansion or contraction. As a consequence, the range of allowable gas pressure
is very limited and is controlled by the value of u w and the particle size and arrangement. It is,
however, likely that in many field situations, gas will be trapped in sand by an overlying layer of silt or
clay, so that the pressures at the interface may be dominated by the menisci that can develop in silts
and clay.

gas bubble gas bubble

a) R = Rc: minimum radius b) R = a: maximum radius

Figure 2-2. Idealization of gas-water meniscus.

6 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

2.4 Fluid flow: gas and water


Silts and clays

Thomas (1987) has shown experimentally that, in the consolidation of these soils, gas moves in
solution in the pore water but does not flow directly through the soil. The gas exists mainly as isolated
large bubbles and Wheeler et al. (1990) have shown theoretically that it is difficult for the gas to
displace the sail while maintaining the bubble structure. It might be possible for flow to occur out of the
bubble if local weaknesses exist that allow particles to be pushed apart, as might occur in highly
anisotropic soil, but, as has already been discussed, the more likely effect of an increase in the gas
pressure is to increase the overall cavity volume.

This will not lead to loss of gas from the bubble. Gas will therefore move through the soil either carried
in the pore water (where water flow is occurring) or diffusing through it, under the action of
concentration gradients caused by gas pressure differences between gas bubbles.

The flow of water is not rnuch affected by the presence of gas, and the water permeability will remain
close to that of a saturated soil under similar conditions. For his theoretical modelling of the
consolidation of gassy soil, Thomas (1987) calculated a revised water permeability based on the
reduction of available pore spaces for the water to flow through. However, he found that a better
prediction of the observed behaviour was obtained using the original water permeability rather than
the revised one.

Although the gas permeability in silts and clays is generally low through the intact material, gas will
flow at much higher rates through fissures, particularly in the stiffer clays where such features are not
uncommon.

Sands

In these soils, both gas and water are able to flow reasonably freely between sand grains. Water
permeabilities are lower than for a corresponding saturated soil, due mainly to the reduced volume of
water-filled voids.

2.5 Undrained strength, compressibility and shear stiffness


Silts and clays

The various effects of gas on the geotechnical behaviour of fine-grained soils have been reported in
detail in a number of publications, including Nageswaran (1983), Wheeler (1986), Thomas (1987),
Wheler (1988a, 1988b), Wheeler and Gardner (1989), and in a general overview, by Sills et al. (1991).
Some of the significant features are highlighted here.

Gas can have the effect either of increasing or of decreasing the undrained strength of a silt or clay,
depending on the consolidation history and the ambient water pressure (Wheeler, 1988b). The
weakening effect of gas on the undrained strength of soft soils becomes more pronounced as the
water pressure increases, i.e. with increasing water depth. Figure 2-3 shows results from triaxial tests

7 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

on kaolin samples conducted by Sham (1989) at an ambient water pressure of 500 kPa
(corresponding to a depth of 50 m below sea level), which indicate that a gas volume of only 1-2% of
the total sample volume can produce a 25% reduction in undrained strength. Related theoretical
studies, however, suggest that there is a limit to the weakening effect of increased water depth.

The compressibility of a sail is increased by the presence of gas. Thomas (1987) has shown that
changes in the void ratio of a gassy soil during one-dimensional consolidation are greater for the soils
with higher gas contents. He showed that the compressibility is predictable from the gas content if the
compressibility of the saturated sail is known. This is described in more detail in Sills et al. (1991).

Figure 2-3. Idealization of gas-water meniscus.

Sands

When a dense sand is loaded rapidly to failure in conditions simulating storm loading, the pore water
pressure drops, and this causes a high undrained strength. In a gassy sand, or, indeed, in one
containing pore water with dissolved gas, the gas comes out of solution and expands, limiting the drop
in pore water pressure that the sand can sustain before cavitation occurs, and hence limiting the
undrained strength.

Undrained triaxial tests have been carried out on dense sand of a particle size distribution typical of
North Sea sands (Sills, 1989), and show that this effect can be significant. For example, in a sand
tested at pressures equivalent to a water depth of 40 m, with a degree of saturation before shearing of
99% (i.e. 1% of the pore space was occupied by gas, 99% by water), the undrained strength was
reduced to about 60% that of a similar sand, containing no gas, either in the free state or in solution.

8 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

2.6 Gas evidence

2.6.1 From seismic data


Geophysical data have for long proved useful in identifying shallow gas areas in marine sediments.
Dramatic improvements have been made these last decades, namely with the implementation of AUV
mounted systems. This section is not intended to describe or compare seismic techniques, nor to
advise on data interpretation which should be made by geophysical experts. The aim of this section is
to:

- familiarize geotechnical engineers with terms or concepts used by geophysicists and often poorly
or even miss-understood,

- highlight the limitations of geophysical instruments with regard to gassy soils considerations for
geotechnical risk assessment and foundation design,

- emphasize the need to interpret geophysical data in the light of the general geological and
morphological context rather than just relying on first level reporting conclusions.

The most commonly used sign to detect shallow gas is acoustic masking, also called acoustic
blanking. Acoustic masking appears in seismic reflection profiles as patches where reflections are faint
of absent. This results mainly from the dissipation of the signal energy in the gassy voids or also from
the disruption of sediment layering by the migration of gassy pore fluids (Judd, et al., 1992). Example
of seismic reflection profiles showing acoustic masking zones is given on Figure 2-4.

2.6.2 From drilling operations


Blowout is the most spectacular and dangerous risk related to shallow gas. Management and
mitigation of drilling risks associated to shallow gas are not within the scope of this report. This section
is intended to list the phenomena which can be observed during geotechnical drilling and highlight
how they can be interpreted with regard to the assessment of geotechnical risks.

Evidence of gas during geotechnical drilling operations can come from:

- gas or water discharge when drilling a pilot hole (predrilled hole of small diameter to investigate
possible presence of gas before starting with conventional drilling systems);

- gas or water discharge when drilling a borehole using conventional drill-strings;

- gas release when retrieving a wireline operated sampler or in situ tool.

Observations made on Site 1 are used for illustration. Gas eruptions lasting a few minutes to hours
were undergone during drilling. Figure 2-5 shows the most violent eruption where the gas gush
reached about 10m above the drill pipe. Discharges of water pulp containing methane were also
frequently observed following recovery of soil cores or retrieval of CPT tool. Lunne et al. (1996)
described on another site a gas kick with both gas and sand escaping from the drill string, whilst
attempting to sample a sandy layer.
9 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 2-4. Example of acoustic masking on sub-bottom profiles, AUV data (2012)

Figure 2-5. Gas eruption reaching 10m above drill pipe: mixed soils, seabed about 6m below MSL
(Caspian Sea).

Gas concentrations during discharges of gas were measured as percentage of “low concentration limit
for flame spreading” (LCLFS), 100% of LCLFS being considered as emergency concentration : high
LCLFS values were measured when penetrating shelly silty sand beds. In another close borehole, no

10 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

significant readings were observed when penetrating the same silty sand units. The first borehole
might have acted as “relief well” for the gas.

Sudden and violent eruptions of gas (blow outs) are generally interpreted as the result of perforation of
a pressurized gas pocket. This assumes that 1) there is an accumulation of free gas in the soil, 2) the
gas is under pressure i.e. in excess of hydrostatic and 3) the soil is sufficiently permeable to let the
gas migrate rapidly to the borehole and maintain the eruption active at least for a certain duration
(depending on whether the pocket is of limited extend or fed from other source(s).

However Kortekaas and Peuchen (2008) have shown that airlift pumping of gas at hydrostatic
pressure can be at the origin of gas discharge during drilling and other intrusive activities. Airlift
pumping is based on the pumping effect achieved when gas is injected (artificially or naturally) into a
liquid or a solid-liquid mixture. The source of airlift pumping is not the in situ overpressure of the free
gas, but the low unit weight of the free gas which generates an upward flow of the fluid/gas mixture in
the drill pipe by decreasing the pressure of the fluid column below hydrostatic.

Airlift pumping is frequently initiated by the retrieval of samples or in situ tools from the bottom of the
borehole during wireline operations. The suction caused by the extraction of the tool (also called
“swabbing” pressure) is responsible for the initiation of gas exsolution. In impermeable layers, the
phenomenon is likely to be limited to bubbling of low intensity but in permeable layers (sands) a
sustained airlift regime can establish and it may become difficult to differentiate from the case of a
pressurized gas pocket. Calculation of the minimum gas flow rate required to generate upward
circulation at the in-situ hydrostatic pressure can give a useful indication.

The above considerations are consistent with Sills et al. (1992) who concluded that blowout hazard
during drilling can be associated with high pressure gas pockets beneath impermeable beds whereas
gas distributed through fine-grained soils is unlikely to induce catastrophic events.

2.6.3 From sample observation


Visual observation of the core samples can give pertinent information on the presence of gas. When a
sample is recovered to the vessel deck, its pressure has fallen from the ambient in situ pressure to
zero (atmospheric pressure). In great water depths the drop in pressure may be expressed in MPa. In
presence of gas consequences are as follows:

- Volumes of free gas already contained in the sample will dramatically expand;

- Gas in dissolved form will come out of solution and exhibit in turn volume expansion;

In permeable materials (sands) gas exsolution and expansion may take place gently during core
retrieval so that no particular sign may be observed on the core once recovered on the deck and
examined. In semi-permeable materials (e.g. silty sands), bubbling may still be observed once core is
on deck but the sample may not have visibly suffered from the release of the gas.

11 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

In less permeable soils (silts and clays) gas release has necessarily damaged the material. To escape
from the soil mass, the gas must have created paths, fissures, interconnected cavities, etc. These
features are visible at the surface or inside the core. It should be emphasized that features observed
on cores are representative of the structure of the sample as recovered and not of the soil in situ.

Gas blisters are a common observation on gassy fine-grained samples. Blisters are cavities resulting
from gas exsolution and formation of a bubble bigger than intergranular spaces. The shape of blisters
can change in function of soil type: blisters observed by Boudreau, et al., 2005 presented clear shape
differences in function of the stiffness of the surrounding matrix (Figure 2-6).

Figure 2-6. Blisters shape differences in function of the surrounding matrix (X-ray tomography):
(A) spherical blisters in white loose carbonate silty sand; (B) oblate spheroid blisters in brown dense
silty sand Sub-tidal sediment of Southern UK (low water depth).

In deep waters sediments, gas bubbles are expected to be highly eccentric oblate spheroids
(disks) that grow either by fracturing the sediment or by reopening pre-existing fractures (Johnson
et al. (2002)). Examples of gas blisters and fissuring due to gas expansion observed during
geotechnical investigations in the Gulf of Guinea are presented on Figure 2-7.

The basic technique used to collect samples in deep water sediments is wireline push sampling: a 3”
diameter, 90cm long thin wall metallic tube is pushed into the sediment and retrieved on board the
vessel. A standard procedure on board geotechnical vessels consists of extruding samples from the
sample tube: the core is carefully pushed inside the tube by a well calibrated piston and progressively
extracted from the tube.

Bubbling is commonly observed in gassy soils at the surface of the core during stripping. This was in
particular the case for silty samples recovered during site 1 investigations. The reason is that a small
confining pressure in excess of the atmospheric pressure remains locked in the pore fluid/gas mixture
due to the friction between the sample and the tube. This pressure drops to zero during extrusion
causing residual exsolution. In some instances this excess pressure can be significant enough to
cause more or less “violent” expulsions of core sections. “Explosive extrusions” were observed during
stripping of site 2 investigations samples. They were accompanied by production of white smoke.

12 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

It is reminded that methane is odourless. Strong H2S odours are frequently noted by operators when
extracting cores. This is in particular the case for samples recovered from the Gulf of Guinea. This
observation has never been correlated to the presence of significant gas content in the sediment.

clean sample

gas blisters

Figure 2-7. Gas blisters and fissuring on core samples: Both core samples are from same borehole.
very dark greenish grey clay, with traces of shells and shell fragments, water depth 740m.

2.6.4 From PCPT records


Cone resistance

Based on triaxial testing, Sultan and Garziliglia. (2010) shows that gas exsolution in fine-grained
sediments is accompanied by damages to the soil structure. These damages can be quantified by:

- a decrease in the preconsolidation pressure compared to their fully saturated equivalents,

- a correlative decrease in the undrained shear strength.

Nageswaran (1983) reported similarly low shear strengths and underconsolidated sediments at zones
where gas content was high. As cone resistances are approximately proportional to undrained shear
strengths, weaker zones in a cone resistance profile may indicate local damages due to presence (or
past presence) of free gas in situ.

Pore pressures

Standard PCPT systems measure the pore pressure u 2 at the sensor level. In the absence of gas, u2
can be expressed as:

13 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

umeasured = u2 = u0 + ushearing (2-3)

where: u0 is the hydrostatic pressure at the sensor level, ushearing is the increase of pressure due to the
shearing of the soil by the CPT tool in undrained conditions, ushearing is a function of clay characteristics
(rigidity index, plasticity, permeability).

In presence of gas, the equation should be written as follows:

(2-4)
umeasured = u2 = u0 + ushearing + upressurized_gas

where u0 is the hydrostatic pressure (as bove), ushearing is the increase of pressure due to the
shearing of the soil by the CPT tool in undrained conditions. Compared to the equivalent
saturated sediment this value may change if the permeability is modified. upressurized_gas is the increase
of pressure resulting from the presence of gas.

Figure 2-8 illustrates the principle of pore pressure variations u2 in presence of gas. As both
ushearing and upressurized_gas depend independently of the presence of gas, the sole reading of u2 is
insufficient to conclude on the presence of pressurized gas in the sediment. Performing
dissipation tests (i.e. allowing ushearing to dissipate with time) is a requirement to identify the
presence of pressurized gas.

Figure 2-8. Schematic pore pressure record.

When interpreting pore pressure measurements in gassy sediments, extreme attention should be paid
on the reliability of the data as interferences in the pore pressure probe can alter the measurements
as outlined by Nageswaran (1983) or Peuchen et al. (2005).

Continuous PCPT records from site 2 are presented on Figure 2-9. The green profile shows an
anomalous low net cone resistance between 18 and 23m penetration which deviates from the linear
increase in cone resistance observed on nearby PCPTs and usually observed in normally
consolidated fine-grained soils of the Gulf of Guinea (as reported by Colliat et al., 2010). At the same
levels, pore pressure records disclose also an anomalous behavior, with values of Δu 2 = umeasured –
14 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

u0 lower than the expected linear trend. These observations may suggest that the sediment between
18 and 23m has been damaged by gas exsolution which has resulted in:

- a decrease in undrained shear strength (lower qnet values),

- an increase of the permeability (fracturing) causing a decrease in ushearing and in u2.

The low u2 value might indicate that no gas overpressure (no free gas?) is present at time of
measurement.

SB - seabed mode record, DH - downhole mode record, clayey soils.Site 2, deepwater Gulf
of Guinea Water depths about 800 m

Figure 2-9. Low net cone resistance and pore pressures on Seabed CPT records.

Deep-water CPTs can be carried out in:

- down-the-hole mode, or in

- seabed mode.

“Down-the-hole” mode, also called “Downhole” mode, is where a wireline operated cone penetrometer
is pushed into the ground from the bottom of a drill hole. Testing sequences are as follows, assuming
the drillstring is at level z below seabed and s is the PCPT stroke:

- a PCPT is performed between z and z+s;

- the PCPT rod is extracted from the seabed and the tool recover to the deck;
15 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

- drilling is performed from z to z+s;

- tool is reintroduced in the drillstring and clamped at the bottom of the drillstring;

- a new PCPT is performed between z+s and z+2s.

The drilling fluid is seawater so that the pressure at the bottom of the borehole during drilling is slightly
lower than the hydrostatic pressure. The stroke is normally 3m but 4.5m tools have been recently
made available.

“Seabed” mode means that the PCPT tool is located on a rig resting on the sea bottom and remotely
operated from the surface. The rod string is penetrated in a continuous mode from the sea bottom to
the end of the test. Up to 40m penetration can be achieved in a single stroke.

Downhole and Seabed mode CPT records usually disclose very similar cone resistance profiles (as
shown on Figure 27 between 0 and 13m penetration). On site 2, investigations were made at the same
locations using both CPT modes and surprisingly different records were obtained at certain levels
below seabed.

Instead of being gently aligned to provide a regular (i.e. globally linear) increase of qnet as a function
of penetration (as observed above respectively 35 or 40m but also for DH_3, between 0 and 13m),
records indicate sequences characterized by:

- higher than expected gradient in cone resistance over a particular stroke;

- losses in cone resistance between the end of a stroke and the beginning of the next stroke.

These sequences clearly visible between 35-40 and 60m below seabed have been called rebooting
sequences. Unfortunately no continuous CPT records are available for comparison below 40m (limit of
penetration for seabed mode CPTs).

A direct comparison between Downhole and Seabed mode records could be made at another location
between 18 m and down to 36 m:

- downhole records characterized by rebooting sequences;

- seabed records at the same location;

- an average of seabed records obtained in neighbouring sites;

- a 30m penetration seabed T bar recorded in the vicinity of the location and showing anomalous
low resistances.

It is observed that:

16 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

- the rebooting sequences tend to reach the continuous cone resistance values at the end of the
stroke;

- the cone resistance at the beginning of the rebooting sequences are close to the low profile
given by the anomalous Tbar.

An explanation of these apparent anomalies can be attempted in a gassy soil context where no free
gas is present in situ but fluid is saturated or close to saturation. It is anticipated that the loss in total
pressure at the bottom of the drillstring due to the only partially unbalanced soil column combined with
the swabbing pressures resulting from the extraction of the CPT tool can generate gas exsolution in
the soil located below the drillbit. Exsolution causes damage to the clay (fracturing). As time between
end of drilling and start of new CPT is reduced, exsolution remains confined in the upper few meters of
sediments. At the beginning of the new PCPT stroke, damaged soil properties are recorded, but
progressively the sensor encounters less damaged sediment and records quasi intact properties at
end of penetration. Quasi intact properties are recorded by seabed continuous PCPTs because cone
penetration generates increases in total pressure (i.e. no exsolution).

In contrast with the anomalous low cone resistances discussed previously, this explanation assumes
that no gas exsolution has taken place in situ before drilling operations. It is however interesting to
note that the degree of damage caused by exsolution due to drilling (cone resistance values at
beginning of rebooting sequences) might be similar to the degree of damage generated by “natural”
free gas generation in the soil (low T bar resistance values). Most likely both types of anomalies may
coexist at the scale of a site.

Sultan et al. (2007) present PCPT profiles recorded in deep waters offshore Nigeria on a zone prone
to presence of gassy soils (evidence of BSR, gas hydrates, anomalous Vp values). The black record is
considered as a reference PCPT profile corresponding to saturated sediment. The red record
corresponds to a PCPT performed in a zone of low Vp value (Vp < 1450m/s) suspected of presence of
free gas. Cone resistances are very similar for both records but pore pressures in excess of
hydrostatic Δu2 are higher for the red record.

The response of PCPT PM33-E may be explained assuming that small bubbles are in sufficient
quantity to lower the compressive velocity and generate overpressures but without destructuring the
material (cone resistance not affected). In the absence of dissipation test it is not possible to assess
whether the higher value of Δu2 is due to pressurized gas or to higher values of ushearing resulting
from a lower permeability.

2.6.5 From laboratory test results


High concentrations of dissolved gas in clays can be evidenced by comparing results of laboratory
tests performed on board or in the laboratory with results of in situ tests.

In situ tests are performed at ambient pressure in an intact soil whereas laboratory tests are performed
on samples already submitted to drastic drops in total pressure from the ambient pressure to the
atmospheric pressure. If high concentrations of dissolved gas are present in situ, exsolution will take

17 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

place during core recovery resulting in damages to the clay structure. These damages can be
considered as the most significant damages that the material can suffered from exsolution.

Example presented on Figure 2-10, undrained shear strengths measured from laboratory vane tests
are compared to undrained shear strength profiles derived from PCPT data using N k factors of 14 and
20 (qn = Nk.Su). In the upper 2 meters laboratory results match with in situ data using a N k factor of 14
which can be considered as a representative value for these fine-grained undisturbed saturated
sediments. Below 2.3m lab test results deviates from the N k = 14 trend to progressively pass below
the Nk = 20 trend around 3m penetration. Samples taken at the same location clearly show the
presence of gas bubbles around 3.15m. The drop of shear strengths below the N k = 20 profile can be
attributed to damages created by gas exsolution. Note that Nk = 20 may be considered as an upper
bound for correlating CPT and lab test results in this type of material when saturated.

Figure 2-10. Undrained shear strength of gassy sediment – CPT data versus lab test results.

18 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

3. LABORATORY AND IN SITU TESTING

3.1 Experimental set-up


A series of experimental tests was conducted using the axis-translation method to control
independently carbon dioxide and pore water pressures. Injecting carbon dioxide at several hundred
kPa was not sufficient to start sediment desaturation and water expulsion.

This suggests that in the natural environment, and for the studied sediments, the presence of free gas
in the upper porous layers seems to be possible only through

(a) local biogenic processes,

(b) gas flow through micro faults or

(c) upward circulation of deeper water saturated by gas.

Therefore the experimental set-up was modified in order to inject the gas in a dissolved phase and
generate free gas in the sediment by way of gas exsolution due to undrained isotropic unloading
following a decrease in the confining pressure (method already used by Lunne et al., 2001;
Amaratunga & Grozic, 2009). The new experimental triaxial set-up is presented in Fig. 3, showing the
carbon dioxide injected as a dissolved phase into the sediment.

Axial strain was measured by a local displacement gauge, and the volumetric strain was determined
from the volume change of the confining water (by taking out the volume change due to the cell
deformation) and also using the local displacement gauge on the sample surface. Carbonated water
was prepared separately in a cylinder several days before the triaxial test; this was done in order to
ensure proper dissolution of carbon dioxide gas at the desired applied pressure.

The circulation of carbonated water was carried out by applying a differential pressure (or by gravity
with differential hydraulic heads) at the top and bottom of the specimen, with these connected to
carbonated water cylinders. The bottom of the specimen was also connected to a pressure/volume
controller used with distilled or carbonated water during drained consolidation tests.

Compressional (P) and shear wave (S) velocities were measured using a bender element system
provided by GDS Ltd, UK. For these special triaxial tests, sediment samples were surrounded by two
latex membranes with a layer of grease in between in order to ensure minimum escape of dissolved
carbon dioxide and free gas during the exsolution phase.

The degree of water saturation, Sr, was calculated from the volume change of the confining water and
the local vertical displacement gauge. The water and sediment matrix compressibilities were
considered negligible.

During the special triaxial test, the change in the void ratio and the water content are measured
allowing the calculation of the degree of water saturation using the Fredlund and Rahardjo (1993)
equation.

19 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 3-1. Experimental set-up where carbon dioxide is injected as dissolved phase into sediment.
Decrease of confining pressure after dissolved gas circulation generates free gas into sediment pores
by exsolution. The triaxial cell is equipped with GDS bender elements.

3.2 Index properties of the studied clays


Experimental tests were carried out on natural intact sediments recovered from the GoG at a water
depth of about 2000 m and between 18 and 20 m below the seabed. Samples were taken from a
reference site without any indication of the presence of free or dissolved gas. In this work the fall-cone
method described by Feng (2001) was used to determine the Atterberg limits.

The natural water content and the Atterberg limits confirming the very high plasticity of the sediments
(PI ¼ 135).

Grain-size distribution curves (not presented here) were initially obtained by a laser diffraction grain-
size analyser. Surprisingly, the particle-size distributions showed a very low clay concentration (lower
than 10%) with respect to the measured plasticity index.

It was clear that for the studied site, and more generally for sediments from the GoG, clay contents
cannot be correctly detected by means of classical methods used for grain size distribution. Indeed, for
the GoG sediments, illite and smectite appear generally as aggregates (Thomas et al., 2005).

20 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

In order to overcome the inaccuracy of the clay content detection using the grain-size distribution, it
was decided to determine the clay composition of sediment samples by powder X-ray diffraction
(XRD) analysis. Results of XRD scans were performed on dried, crushed samples.

The XRD analyses obtained show that the clay content is much higher than that determined by the
laser diffraction grain-size analyser.

3.3 Experimental test programme


Four stress paths were defined to characterise the effect of the free gas on the mechanical properties
(compressibility, yielding and shearing) of the sediments. For the four tests (Tests A, B C and D), the
samples were saturated by carbonated water under low effective stress (usually less than 10 kPa),
and then consolidated to 100 kPa.

The pore carbonated water pressure (uc) was equal to 300 kPa for Test A, 400 kPa for test B and 600
kPa for tests C and D. The use of three different pore carbonated water pressures was necessary to
consider the dependence of the carbon dioxide solubility on the pore water pressure (see Figures 3-2,
3-3, 3-4 and 3-5)

Free gas generation within the tested samples occurred during the subsequent undrained unloading
phase. For the four stress paths, samples were then reloaded under drained conditions (isotropic
effective stress of around 200 kPa for Tests A, B and C, and around 100 kPa for Test D), and sheared
in undrained conditions and constant total confining pressure a3 in a displacement-controlled loading
by increasing axial deformation at an axial strain rate of 0.1%/h.

During the various phases of the tests, changes in pore pressure at the top (where carbonated water
was injected) and the bottom of samples, as well as P-wave and S-wave velocities, were measured.
In addition, and as reference tests, four oedometer tests and three triaxial CU tests were carried out
on water-saturated sediments.

21 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 3-2. Test A: (a) stress paths; (b) P-wave velocities; (c) S-wave velocities

22 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 3-3. Test B: (a) stress paths; (b) P-wave velocities; (c) S-wave velocities

23 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 3-4. Test C: (a) stress paths; (b) P-wave velocities; (c) S-wave velocities

24 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 3-5. Test D: (a) stress paths; (b) P-wave velocities; (c) S-wave velocities

25 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

3.4 Test results


Test A (sampled at 18.22 m below seabed) - the total duration of Test A was around 35 days. Sample
saturation, carbonated water circulation and consolidation took place during the first 10 days of the
test. During this first stage, the pore water pressure (uw) applied at the base of the sample was 280
kPa, the total confining pressure was 300 kPa, and the carbonated water pressure was 290 kPa. A
volume of about 1.5 times the initial pore water fluid content was passed through the specimen in the
phase of carbonated water circulation.

The sample was then isotropically consolidated up to 100 kPa by increasing the total confining
pressure to 400 kPa. Drainage of the sample during this phase of consolidation took place at the
bottom of the sample, while a continuous monitoring of the pore pressure at the top of the sample was
performed. The consolidation was considered achieved when the pore pressure at the top of the
sample had reached the imposed pore water pressure at the bottom. In order to generate free gas
within the sample, the confining pressure was then decreased by four subsequent steps down to 10
kPa under undrained conditions.

A signify cant decrease of the P- and S-wave velocities was observed immediately after the third
undrained unloading phase. The final unloading stage was maintained for 2.5 days. During the gas
exsolution phase, the degree of water saturation reached a minimum value of 0.95.

Next, the sample was isotropically consolidated in five steps to reach a mean effective stress of 150
kPa. Once again, the consolidation was considered achieved when the pore pressure at the top of the
sample had reached the imposed pore water pressure at the bottom of the sample. The increase of
the effective stress induced an important increase in the S-wave velocities (from 60 to 120 m/s). The
last phase of the test was the undrained shear test at a constant axial deformation rate of 0.1%/h.

Test B (sampled 18.35 m below seabed), Test C (sampled at 18.50 m below seabed) and Test D
(sampled at 19.56 m below seabed) The total durations of Tests B, C and D were 33 days, 35
days and 24 days respectively. The chronology and the applied pressures during the different tests
was : during the gas exsolution phase, the degree of water saturation reached a minimum value of
0.93 for Test B, 0.91 for Test C and 0.91 for Test D. For the three tests, a significant decrease
of the P and S wave velocities was observed during the undrained unloading phases. The last phase
of the three tests was again the undrained shearing at constant axial deformation rate of 0.1%/h.

The experimental results have shown the following.

(a) During gas exsolution generated by undrained unloading, the effective stress in the sediment
reduces dramatically, and the mean degree of saturation at the end of the gas exsolution is between
91% and 97%.

(b) Consolidation following gas exsolution shows a higher compressibility of the gassy sediments
and a lower preconsolidation pressure. The higher compressibility is the result of the presence of both
occluded gas in the sample, and destructuring of the clay caused by exsolution. The amount of
damage caused by exsolution seems to be a key parameter for the resulting compressibility (Figure 3-
6).
26 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

(c) The shear strength of sediments following exsolution and subsequent consolidation is affected
by the damage caused to the sediment structure during exsolution. Destructuring by gas exsolution
produces a characteristic change in the effective stress paths. This corresponds to a collapse of
fractures and pores, damaged by gas exsolution and reconsolidation. The extent of the discontinuity
increases with initial gas concentration, and therefore with the damage level.

The degradation of the undrained shear strength due to gas was evaluated in the framework of the
existing Wheeler (1988) theoretical model. The shear strength observed in the experimental work
following gas exsolution and in the absence of any additional mechanical consolidation was correctly
modelled by combining the lower and upper bounds of the Wheeler (1988) model.

For engineering and foundation design activities involving gassy sediments it is essential to keep in
mind that flooding of damaged areas and gas bubbles and dissolution of free gas do not imply an
immediate shear strength recovery of the sediments. Indeed, shrinkage and collapse of the dam- aged
zones generated by gas exsolution and expansion can be in part irreversible, and may lead to a
permanent decrease of the sediment strengths. At locations where the presence of free gas or
dissolved gas is suspected, the determination of Su from laboratory tests may lead to important
underestimations of the sediment shear strength.

Figure 3-6. Preconsolidation pressure against: (a) minimum observed degree of water saturation
during gas exsolution phase; (b) damage parameter d, which is directly linked to S r and the porosity

27 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

4. MECHANICAL MODELS FOR GASSY SOILS

4.1 On the behaviour of gassy soils

4.1.1 Gas bubbles and structures of gassy soils


Gas bubbles form in marine soils only if the water phase becomes oversaturated with gas (van
Kesteren and van Kessel, 2002). Gas bubbles formation thus relies on the solubility of gasses in water
given by the Henry’s law:

pi = ki  xi (4-1)

where pi is the partial vapour pressure of gas component i, k i is Henry’s constant for this component
and xi is the mol fraction of i in water.

Equation 4-1 shows that the vapour pressure is proportional to the concentration of dissolved gas. The
vapour pressure has to exceed the liquid pressure for gas bubbles to form (i.e. gas exsolution).
Another important implication of equation 4-1 is that the solubility of gas, captured by the Henry’s
constant, is proportional to the liquid pressure. For temperature in the range 0- 50°C (typical for the
upper 1000 m of sediments), the Henry’s constant, inversely related to the solubility of gas, typically
increases with temperature (van Kesteren and van Kessel, 2002; Smith and Arvey, 2007). Physically,
gas becomes less soluble as the temperature increases because the increasing kinetic energy allows
molecules to break intermolecular bonds in the liquid water and move into the gas phase (Waite et al.,
2009). Other factors including capillary pressure, and pore size influence gas solubility (Waite et al.,
2009).

As soon as pore water is oversaturated with gas, bubble nucleation may take place, generally
heterogeneously, as a significant energy barrier (activation energy) has to be overcome (van Kessel
and van Kesteren, 2002). This is caused by surface tension of water that limits the nucleation process.
The excess pore pressure Δp needed to overcome surface tension is the capillary pressure, given by
the Kelvin equation:

2T
p  u g  u w  (4-2)
R

where ug and uw are the pressures of the gas and water in the pores, R is the radius of curvature of the
gas-water interface through which surface tension, T, acts.

Hence, owing to the effect of surface tension, the pressure inside gas bubbles is higher than the
ambient pressure, by an amount equal to the capillary pressure Δp. From equation 4-2, it is clear that
capillary pressure (and gas oversaturation in water) increases with decreasing bubble radius. The
minimum possible value for this radius is determined by the size of the throat between particles (Sills
and Wheeler, 1992). The critical value, Rc, corresponds to a meniscus at the water-gas interface just
bridging the largest distance between two sediment particles (Wheeler, 1988a). The magnitude of R c
is of the order of the particle size and is not related to the bubble radius, a (Wheeler, 1988a). For the

28 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

maximum value of R the bubble cavity radius, a, gives the smallest possible gas pressure: from these
considerations, Wheeler et al. (1990) defined the surface tension limits for the gas pressure as:

In case of a drop in gas pressure to the lower surface tension limit (uw + 2T / a), a further reduction is
prevented by a decrease in the gas volume due to inflow of water into the bubble cavity and
consequent partial or complete flooding (Wheeler, 1988b). By contrast, at the upper surface tension
limit, (uw + 2T / Rc ), the gas pressure is prevented from increasing further by a volume expansion, as
gas moves out of the bubble cavity into the sediment pores (Wheeler, 1988b). The dependence of the
gas pressure on the value of R implies that the effects of free gas are different for clays or silts and for
sands (Sills and Wheeler, 1992).

4.1.1.1 Fine-grained soils


Within clays and silts, free gas normally occurs as relatively low concentrations of discrete bubbles
varying between 0.5 and 50 mm in diameter (Anderson and Hampton, 1980; Jones et al., 1986;
Anderson et al., 1998). Anderson and Hampton (1980) observed that bubbles of less than about 5 mm
in diameter are spherical or approximately spherical, whilst larger voids tend to be elongate.
Irrespective of their shape, bubbles in clays and silts are typically considerably larger than the soil
particles and the normal void spaces, thus forming significant cavities within the soil structure, as
illustrated in Figure 2-a (Wheeler et al., 1988b; Yuan et al., 1992; Abegg and Anderson, 1997;
Anderson et al., 1998; Gardner, 2000).

Figure 4-1. Schematic representation of the structures of soils containing a) gas bubbles much larger
than solid particles (in clays or silts) and b) gas bubbles much smaller than solid particles (in sands).
(Wheeler, 1988a).

According to equation 3, the small size of clay and silt particles and consequently of the small pore
throats between them implies that gas pressure may vary widely in fine-grained soils. This means that
bubble expansion or contraction normally occurs before movement of gas into the sediment matrix or
bubble flooding (i.e. if ug < uw) (Wheeler 1988a and 1988b). Bubble cavity expansion provides a limit
for gas pressure, with the maximum possible difference between gas pressure ug and confining stress
being a function of the sediment strength (Wheeler et al., 1990). It follows from the cavity expansion
theory (Vesic, 1972):

29 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

4  a  
3
 G  
p max  S u 1  ln   ln1    (4-3)
3 
  Su   a0   

in which G is the shear modulus, Su the undrained shear strength, a the bubble radius and a 0 the initial
bubble radius.

Since the overwhelming majority of pores in clays and silts is smaller than 10 μm and considering
undrained shear strength, Su, ranging between a few to a few tens of kPa in upper sedimentary layers,
equation 4-3 predicts that bubble growth may only occur by deforming the grain matrix (van Kessel
and van Kesteren, 2002). Experiments conducted with clay samples by van Kesteren and van Kessel
(2002) and van Kessel and van Kesteren (2002) suggested that bubbles grow in the direction of the
smallest principal stress until they reach a critical diameter and initiate cracks creating pathways for
gas transport in the sediment matrix. A number of studies based on laboratory and numerical
experiments have found extremely good agreement with the observations and theory presented by
van Kesteren and van Kessel (i.e. Sills and Gonzalez, 2001; Johnson et al., 2002; Boudreau et al.,
2005; de Vries et al., 2007; Jain and Juanes, 2009; Algar and Boudreau, 2009).

Van Kesteren and van Kessel (2002) defined three conditions for cracks formation according to
fracture mechanics theory (Broek, 1983):

- the ratio between the isotropic effective stress and the effective cohesion must be less than unity,

- the bubble size must be sufficient to allow for stress concentration,

- the plastic deformation zones around the bubbles must not overlaps, as crack will only form on the
boundary between the elastic and plastic zones.

According to the first point, cracks may easily be formed in overconsolidated sediments with large
effective cohesion, while in underconsolidated sediments, no crack can be formed. The second point
implies that no cracks form around small bubbles.

Experimental results reported by Sills and Gonzalez (2001) with clayey silts of estuarine origin
revealed that, by providing a quick route for pore water dissipation, the formation of cracks may
promote a new phase of sediment consolidation manifested by increased rate of settlement. The same
phenomena was observed by van Kessel and van Kesteren (2002) who noted as Sills and Gonzalez
(2001) that cracks formation and enhanced pore water drainage initiates at an average gas fraction of
about 0.9.

4.1.1.2 Coarse-grained soils


By contrast with fine-grained soils, free gas in purely coarse-grained soils (i.e. sands without mud)
most likely occurs initially as bubbles that are smaller than particle size and are entirely contained in
fluid within interparticle spaces (Sills and Wheeler, 1992). The resulting soil structure is thus likely to
be as illustrated in Figure 2-b. It is envisaged that the stable sand configuration is one in which gas
bubbles may move with the flow of pore water or under the influence of buoyancy but do not produce
30 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

changes in the structure in the way that it does for fine-grained soils (Sills and Wheeler, 1992;
Pietruszczak and Pande, 1996; Sills and Thomas, 2002; Grozic et al., 2005).

Owing to the larger size of the particles and pore throats, in purely coarse-grained sediments, gas
pressures vary less and remain lower than in silts and clays (Sills and Wheeler, 1992). This means
that bubble flooding normally occurs before bubble expansion or contraction and, in turn that the
sediment matrix is not affected by gas, displacing solely pore water (Wheeler, 1988a and 1988b).
Accordingly, Sobkowicz and Morgenstern (1984) and Fredlund and Rahardjo (1993) argued that for
water saturation higher than 0.85, the tension at the surface of occluded gas bubbles can be
neglected in analysing the sediment matrix parameters, such as effective stress and compressibility.

4.1.2 Consolidation behaviour of gassy soils


The presence of free gas may affect the consolidation behaviour of soils in different ways. First, it
clearly reduces the density and increases the volume a given mass of soil may occupy (Wichman et
al, 2000; Sills and Gonzalez, 2001). Hence, if self-weight consolidation of a gassy soil is compared
with that of a saturated soil containing the same mass of water and solids, the reduced bulk density
implies a smaller driving force for consolidation and a smaller initial excess pore-water pressure
(Wichman et al., 2000). Additionally, as shown experimentally by Orlob and Radhakrishna (1958),
Brooks and Corey (1964) and Elzeftawe and Cartwright (1983), any gas present in a porous material
may significantly increase the drainage path length as water must flow in the soil matrix around the
larger gas bubbles. As discussed in Wichman et al. (2000), these effects combine to retard the
consolidation process and reduce the final degree of consolidation.

On another hand, due to the high compressibility of gas bubbles, the application of a load to a gassy
soil causes volume change even when the soil is undrained (Thomas, 1987, Sills et al., 1991).
Accordingly, it is necessary to distinguish between the gas void ratio, e g, and the water void ratio, ew,
these being the ratio of gas volume to the volume of solids and the ratio of water volume to the volume
of solids, respectively. Both of these ratios are related to the degree of water saturation, S w, as:

eg = 1 – Sw) e (4-4)

ew = Swe (4-5)

eg + ew = e (4-6)

The volume fraction of bubbles f can be expressed in terms of total void ratio e and degree of water
saturation Sw, as:

f
1  S e   e
(4-7)
1 E

31 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Thomas (1987) performed a series of one dimensional tests on reconstituted samples of estuarine silty
clays (Combwich mud) made gassy by the addition of methane saturated zeolites. The results of these
tests, as presented in Figure 3, revealed that initial load applications under undrained conditions
resulted in an immediate change in total void ratio, e and gas void ratio, e g. Both of these changes
were ascribed solely to gas compression since no simultaneous change occurred in the water void
ratio, ew (Figure 3). As drainage was allowed during subsequent stages of the tests, the changes in e
and ew were almost equivalent, with very little change in eg during the consolidation stage (Figure 3).

Based on this series of tests, Thomas (1987) confirmed Wheeler’s (1986) works showing that fine-
grained gassy soils behave as saturated soils containing discrete compressible inclusions. Most
importantly, Thomas (1987) pointed out that changes in gas volume are solely related to changes in
total stress and are not altered by water drainage during consolidation. He also emphasised that the
behaviour of the soil matrix is governed by the stress difference (σ-uw). Following Sills et al. (1991) this
stress difference is referred to as the ‘operative stress” rather than the ‘effective stress’ as the overall
gassy soil response to an applied load is not only dependent on the difference between total stress
and pore pressure. The independent deformation of a gassy soil due to changes of total stress and
operative stress led Thomas (1987) to develop a consolidation theory based on the concept of “two-
phase” or “double” compressibility, where the compressibilities of the gas and the surrounding
saturated soil matrix are considered to be independent.

Figure 4-2. Total, water and gas void ratios plotted against time for step-load consolidation tests.
(Thomas 1987).

32 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

As shown in Figure 4, another series of oedometer tests conducted by Thomas (1987) on Combwich
mud samples with initial degrees of saturation ranging from 0.87 to 1 revealed that the higher the initial
gas content, the greater the value of total void ratio for a given value of operative stress, and the more
compressible the soil.

Figure 4-3. Total void ratio plotted against operative stress for consolidation tests with
load increasing with time (Thomas 1987).

Thomas (1987) compared the results of this series of tests with two possible models. The first of them
assumed that gassy soils can be idealised as saturated soils but with a compressible fluid. It implies
that the strains are caused uniquely by changes in effective stress and that the compressibility of the
pore fluid influences the pore pressure response. The compressibility of the pore fluid βw’ is related to
the degree of saturation, Sw, following Verruijt (1969) as:

 'w   w 
1 S w 
pa (4-8)

where βw is the compressibility of gas free water, and pa the absolute pore water pressure. Based on
the ‘double compressibility’ theory proposed by Thomas (1987), the second model captured the
independent contributions of the saturated matrix and the gas to the overall gassy soil compressibility.
It implies that the strains are the sum of the strains in the saturated soil caused by change in operative
stress, and strains in the gas void caused by change in total stress.

Figure 4-4 shows comparisons between laboratory results and the predictions of the two models in
terms of pore pressure and pore pressure normalised by the applied vertical load.

It can be seen that the better match between predictions and measurements is obtained with the
double compressibility model, according to which the volumetric strain is totally absorbed in the gas
compression without any volume change of the soil matrix away from the voids.

33 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

As it requires there to be no change in operative stress, an increase in total stress is matched by an


increase in pore pressure and the predicted initial value of the ratio Δu w/Δσv is unity, as for a saturated
soil. It is clear, from Figure 4-4a, that incorporating a compressible pore fluid does not provide an
acceptable model for the fine-grained gassy soil tested.

This approach may, however, be considered as valid when gas exists as small bubbles within the soil
pores, as may be the case for coarse-grained soils (Pietruszczak and Pande, 1996; Sills and Thomas,
2002; de Groot et al., 2006). In this case, immediate volume changes occur in the pores of the soil and
are associated with corresponding strains in the soil matrix.

As these strains require an effective stress, the pore pressure increase in the compressible fluid has to
be lower than the total stress increase (Sobkowicz and Morgenstern, 1984; Sills and Thomas, 2002;
Amaratunga and Grozic, 2009).

Figure 4-4. Experimental results and model prediction using: a) compressible fluid model and
b) double compressibility model (Thomas, 1987).

Haththotuwa and Grozic (2008) reported results of triaxial tests aiming at highlighting the undrained
compressive behaviour of gassy Ottawa sand in response to isotropic step increases in total stress. As
shown in Figure 4-5, the behaviour of the tested gassy sample greatly differs from that of a theoretical,
ideal saturated soil by showing a significantly damped pore pressure response to increasing cell
pressure.

After immediate increases, the pore pressures equilibrated close to their initial value. Hence, at the
end of the test, the overall increase in cell pressure of 151 kPa resulted in a change in effective stress
of 134 kPa where the effective stress would have remained at 50 kPa for a saturated soil.

34 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-5. Pore pressure response due to isotropic step increases in cell pressure.
(Haththotuwa and Grozic, 2008)

4.1.3 Elastic moduli of gassy soils


As part of the extensive laboratory tests conducted at Oxford University on the behaviour of fine-
grained gassy soils, measurements of undrained elastic moduli have been reported by Wheeler and
Gardner (1989). Such as those presented by Thomas (1987), the experimental investigations
presented by Wheeler and Gardner (1989) were performed on reconstituted samples of Combwich
mud in which methane saturated zeolites were added for gas bubbles to form.

The laboratory programme included a series of undrained triaxial tests which did not allowed for the
accurate assessment of very small strain. Consequently, secant values of the undrained shear and
bulk moduli were determined at substantial strain magnitude, that is, at a deviator stress equal to half
the failure value (typically an axial strain of about 1%).

Wheeler and Gardner (1989) pointed out that the values of the elastic moduli from this series of triaxial
tests correspond to long-term undrained moduli since there was probably sufficient time for both
localised consolidation in the matrix around each bubble and the movement of gas into solution.

Figure 4-6 shows the secant values of the elastic moduli against the degree of saturation at the start of
shearing, for three distinct values of operative stress. Although there is some scatter of the results,
values of the undrained shear modulus, Gu, and of the undrained bulk modulus, Ku show clear
decreasing trends with decreasing degree of saturation.

Following Wheeler and Gardner (1989), Duffy (1990) reported a series of triaxial tests allowing
measuring accurate values of undrained Young’s modulus E u and Poisson’s ratio u for axial strains
from 0.02% to 1%. The tests were conducted of reconstituted kaolin samples containing methane
bubbles, as well as on reconstituted, fully saturated kaolin samples.

Figure 4-7a presents the experimental results for three different combinations of operative stress and
total stress. It reveals that the normalised Young’s modulus E u/Esat (where Esat is the undrained
Young’s modulus of a saturated sample consolidated to the same operative stress) measured at an

35 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

axial strain of 0.25% decreases with increasing volume fraction of gas f. As shown in Figure 4-7b, the
values of the undrained Poisson’s ratio also tend to decrease as the volume fraction of gas increases
and depend on the values of the operative and total stresses during consolidation prior to shearing.

Figure 4-6. Experimental results of undrained shear modulus (a) and undrained bulk modulus (b) from
triaxial tests with uw0 = 0 and values of σ3- uw0 varying from 100 to 400 kPa.
(Wheeler and Gardner, 1989).

Figure 4-7 also shows the curve corresponding to the theoretical expressions of E u/Esat and νu, based
on the equations of Gu and Ku derived by Wheeler and Gardner (1989). Assuming that, the saturated
soil matrix is a linear elastic material, and ignoring the undrained compressibility of the saturated
matrix and the stiffness of gas within bubble cavities, Wheeler and Gardner (1989) showed that the
theoretical expressions for Gu and Ku simplify to:

3G sat 1  2f 
Gu  (4-9)
3f

4G' 1  2f 1  f 
Ku 
f 3  f  (4-10)

Rearrangement of these equations gives the following expressions for E u and u:

E   E sat 1  1.917f  (4-11)

u 
1
1  0.75f  (4-12)
2

Analysis of Figure 4-7 reveals that the reductions in undrained Young’s moduli and Poisson’s ratios
with increasing gas content were generally greater than theoretical elastic predictions.

36 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Discrepancies between experimental measurements and theoretical elastic predictions, were also
emphasised by Wheeler and Gardner (1989) who attributed them to plastic straining concentrated in
the region of the matrix immediately surrounding each bubble cavity (Figure 4-8a). Duffy et al., (1994)
reported similar lack of agreement between theory and measurements conducted in a resonant
column apparatus at shear strains ranging from 0.0004% to 0.1%, suggesting that local yield zones
around the bubble cavities formed during consolidation prior to shearing (Figure 9-b).

This explanation is theoretically supported by the analyses presented by Wheeler et al. (1990) who
considered the formation of local yield zones around gas bubbles during isotropic compression as a
cavity contraction problem. In agreement with these analyses predicting that the yield zones forming
around each gas bubble during consolidation would tend to extend further into the surrounding matrix
with increased operative stress (Figure 4-9),

Figure 4-8 shows that the discrepancies between experimental results and theoretical elastic
predictions increase as the operative stress increases.

Figure 4-7. a. Variation of normalised undrained Young's modulus with gas content from triaxial
tests, at an axial strain ε1 of 0.25%. b. Variation of undrained Poisson's ratio with
gas content from triaxial tests, at an axial strain ε1 of 0.25%
(Modified from Wheeler et al., 1991).

37 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-8. a. Comparison of undrained shear moduli from triaxial tests with theoretical predictions
(Wheeler and Gardner, 1989). b. Comparison of undrained shear moduli from resonant column tests
with theoretical predictions (Duffy et al., 1994).

Figure 4-9. Schematic view of the development of plastic zones around oblate spheroidal gas bubbles
with increasing operative stress (σ-uw) during consolidation (After Thomas, 1987). a. Initial stage
where there is sufficient load to produce a small plastic zone around each bubble, without any
settlement of the soil matrix. b. Critical loading stage when the plastic zones are just connecting, still
without settlement. c. Connection of plastic zones producing a mechanism which then deform
plastically to allow load to be transferred to the gas resulting in an overall settlement of the soil.

4.1.4 Undrained shear strength of gassy soils

4.1.4.1 Fine-grained soils


Wheeler (1986) conducted a programme of undrained triaxial tests on reconstituted gassy samples of
Combwich mud prepared with the zeolite technique and consolidated under different values of
operative stresses (100, 200 and 400 kPa) and back pressure (0 and 100 kPa).

Figure 4-10 presents experimental results for samples whose degrees of saturation at the start of
shearing, Sw0, ranged from 0.90 to 1.00. As shown in figure 4-10a, values of undrained shear strength,
Su, vary in three different ways as a function of both the operative stress and the cell pressure. It

38 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

appears that the presence of gas bubbles increased the value of S u above the saturated value when
the cell pressure and the initial operative stress were both 400 kPa.

When the cell pressure and the operative stress were both 200 kPa, the gas bubbles appeared to
have little effect on the undrained strength. By contrast, the undrained shear strength was reduced
when the cell pressure was increased to 300 kPa with the operative stress held at 200 kPa and, when
the cell pressure and operative stress were 100 kPa.

Based on this series of tests Wheeler (1986, 1988a) concluded that the effect of gas bubbles on
undrained shear strength were most detrimental at low value of operative stress (corresponding to
shallow depth below seabed) and high values of total stress (corresponding to deep water location).

Figure 4-10b reveals that failure strain increases as the initial degree of saturation is reduced. This
trend is particularly marked for the series of tests with an initial operative stress of 400 kPa which
shows increases in failure strains reaching about 15 % for a decrease in Sw0 of 10%.

Wheeler (1986, 1988a) noted that the degree of saturation increased during undrained shearing so
that, as shown in Figure 4-10c, many gassy samples reached full saturation before failure occurred. All
the curves in Figure 4-10c indicate that full saturation was reached during the shear tests only if S w0
was above a certain critical value Swc.

It can be seen that the values of Swc decrease with increasing values of initial operative stress. By
comparing the degrees of saturation at the start of the tests and at failure with corresponding values of
undrained shear strength, Wheeler (1986, 1988a), observed that samples having reached a fully
saturated state before failure were weaker than the equivalent saturated samples.

Wheeler (1986, 1988a) took this to indicate that their behaviour was still affected by the fact that they
had previously contained gas bubbles, suggesting that the structure of a soil is permanently affected
by earlier presence of the bubbles.

Figure 4-11 shows the stress ratio at failure qf/(pf-uwf) - corresponding to the critical state parameter M
(Atkinson and Bransby, 1978) - plotted against the initial degree of saturation for the different tests
presented in Figure 4-10.

The results appear to be independent of the operative stress and the back pressure, but dependent of
the initial degree of saturation (Figure 4-11). The fact that the stress ratio at failure apparently
decreases as Sw0 is reduced was taken by Wheeler (1986) as evidence that there is not a unique
critical state line in q: (p-uw) space for soil samples containing gas bubbles.

Wheeler (1986, 1988a, 1988b) considered the effect of the gas bubbles on the strength of the gassy
soil in the light of a continuum model consisting of a matrix of saturated soil surrounding isolated gas-
filled cavities.

Based on this model assuming that the saturated matrix is a rigid perfectly plastic material, Wheeler
(1986, 1988a, 1988b) pointed out that the strength of fine grained gassy soils may be lower than that

39 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

of the saturated matrix because gas voids have a weakening effect. Alternatively, if the surrounding
soil consolidates by drainage into the voids, that is through bubble flooding, the matrix strength will
increase and as such could more than offset the weakening due to the presence of the voids.

In addition, reduction in the weakening effect of the bubble voids may arise from bubble shrinkage, as
the gas compresses due to the increasing total stress.

Theoretical analyses of yielding, bubble shrinkage and flooding led Wheeler (1986, 1988a, 1988b) to
show that the undrained shear strength is a function of six non-dimensional parameters:

Su   3  u g0 u g0Pa He w 0 2T 1  e w0 
 functionf 0 ; ; ; ; ;  (4-13)
S usat  S usat S usat 1  e w 0 R c S usat  

where Susat, is the undrained shear strength of a saturated sample consolidated to the same value of
operative stress as a gassy soil of shear strength S u, f0 is the volume fraction of bubbles at start of
shearing, ug0, is the gas pressure at start of shearing, Pa is the atmospheric pressure, H is the
solubility coefficient of gas, ew0, the water void ratio at start of shearing, T the surface tension acting
through the water-gas interface, Rc, the minimum radius of the menisci at water-gas interfaces and, λ
the slope of the critical state line for saturated soils in [e – log(ep’)] space.

The analysis of the yield stress indicated that the value of Su / Susat would be reduced by increasing
the value of (σ3 − ug0) / Susat. Conversely, the analysis of bubble flooding suggested that Su / Susat
would be increased by either reducing 2T / Rc, Susat (which encourages bubble flooding) or increasing
(1 + ew0 ) / λ (which maximises the effects of bubble flooding).

Wheeler (1986, 1988a, 1988b) recognised that the same type of prediction cannot be made for the
parameters (ug0 + Pa) / Susat and Hew0 / (1+ ew0) which have two counteracting effects.

Raising the first or lowering the second of these two parameters tends to promote the work hardening
effect produced by bubble shrinkage and thus to increase Su / Susat. At the same time, changes in
these two parameters tend, by contrast, to prevent widespread bubble flooding which implies a
decrease in Su / Susat. Wheeler (1986, 1988a, 1988b) thus noted that it is difficult to point which of
these two effects would dominate.

He also emphasised that, although the two processes of bubble shrinkage and flooding tend to
increase the value of Su, each process tends to impair the effect of the other.

As a result, the maximum value of Su / Susat would not occur when the two processes act
simultaneously. Wheeler (1986, 1988a, 1988b) points out that it may rather occur due to complete
bubble flooding, since the work- hardening effect induced by complete bubble shrinkage results in a
saturated soil that cannot explain experimentally observed values of Su / Susat rising above unity
(Figure 4-10a).

40 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-10. Undrained shear strength (a); failure strain (b); degree of saturation at failure (c).

41 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-11. Stress ratio at failure against degree of saturation at start of shearing for different
combination of operative stress and back pressure (Wheeler, 1986).

In practice, bubble shrinkage and bubble flooding may interact which, makes derivation of a rigorous
expression for the undrained shear strength impossible. The theoretical works conducted by (Wheeler
1986, 1988a, 1988b), suggest that the shear strength should reach a limiting lower bound value
as the parameter (ug0 + Pa) / Susat is increased. As shown in Figure 4-12, Wheeler’s experimental
results agree with theory and suggest that, for this particular test series, the effect of bubble flooding
dominate the effect of bubble shrinkage.

Figure 4-12. Experimental values of undrained shear strength (Wheeler 1986, 1988a).

42 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

4.1.4.2 Coarse-grained soils

4.1.4.2.1 Loose gassy sands


Pietruszczak et al. (2003) conducted an experimental programme consisting of a series of drained and
undrained triaxial compression tests on Ottawa sand at various confining pressure and initial degrees
of saturation. All samples were prepared in a very loose condition to approximately the same void ratio
of 0.76. Air inclusions entrapped in water allowed the initial degree of saturation to vary in the range
0.87-1. From Figure 4-13 presenting the results of this series of tests, it appears that, as the degree of
saturation is reduced, the development of excess pore pressure decreases rapidly and the response
tends towards that corresponding to drained conditions.

As noted by Pietruszczak et al. (2003), it can be seen that the response of the sample with an initial
degree of saturation of 0.92 is intermediate between the other tests showing responses tending
towards that of a saturated sample in undrained conditions, when Sw0 = 0.96, or in drained conditions,
when Sw0 =0.87 (Figure 4-13b).

Figure 4-13. Undrained triaxial tests results on Ottawa sand with initial void ratio of 0.76 and initial
degree of saturation varying between 0.87 and 1 (Pietruszczak et al., 2003). a. Effective stress paths.
b. stress- strain curves. c. Pore pressure-strain curves.

A similar dependence of the stress-strain response of loose gassy sand to the initial degree of
saturation was previously emphasised by Grozic et al. (1999) who conducted a series of triaxial
compression tests on Ottawa sand specimens charged with carbon dioxide. In view of the results
presented in Figure 4-14, one of the main conclusion of this experimental programme is that, if the
specimen is sufficiently loose (i.e. e0> 0.78) and if the initial degree of saturation is sufficiently high
(i.e. Sw0>0.9), then the gassy specimen can strain soften similarly to fully saturated specimens.

43 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

By contrast, gassy specimens with Sw0 less than 0.9 tend to respond in a strain-hardening manner as
fully saturated samples sheared in drained condition do (Figures 4-14a and 4-14b). As shown in
Figure 4-14-c, the fact that specimens with Sw0 greater than 0.9 nearly reached full saturation before
failure led Grozic et al. (1999) to suggest that a Sw0 of 0.9 was the minimum ( or “cut-off”) value that
samples must be at to strain soften.

Figure 4-14. Undrained triaxial tests results on loose samples of gassy Ottawa sand with initial void
ratio (e0) varying between 0.852 and 0.878 and initial degree of saturation (S w0) varying between 0.816
and 0.912 (Grozic et al., 1999). a. Stress-strain curves. b. Effective stress paths. c. Degree of
saturation-strain curves.

4.1.4.2.2 Dense gassy sands


Because of their dilative tendency, dense sands experience pore water pressure drops when sheared.
While this causes a high undrained strength in fully saturated conditions, in gassy sand it may implies
gas exsolution and expansion, thus limiting the drop in pore water pressure that the sand can sustain
before cavitation occurs (Sills and Wheeler, 1992).

Based on undrained triaxial tests carried out on dense sand, of a particle size distribution typical of
North Sea sands (i.e. d50 ~0.14mm), Sills (1989) showed that this effect can be significant. For
example, in a sand test at pressures equivalent to a water depth of 40 m, with a Sw0 as high as 0.99,
the undrained strength was reduced to about 60% to that of a similar but fully saturated dense sand
(Sills and Wheeler, 1992).

44 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Rad et al. (1994) complemented the study of Sills (1989) by performing triaxial tests on Baskarp sand
(d50 ~0.1 mm) specimens containing free gas and saturated specimens containing gas solely in
dissolved form in pore water.

Figure 4-15 synthesises results of this series of triaxial tests designed to study the effects of gas type
and amount, as well as the effect of pore pressure level on the behaviour of dense sand. Curves 6 and
13 in Figure 4-15 were respectively obtained from reference undrained and drained tests conducted
on specimens saturated with deaired water. Comparison of curves 6, 9, 12 and 13 allows to highlight
the effect of gas of different solubility (i.e. CH4 and CO2) on the behaviour of dense sand.

By contrast with tests 6 and 13, tests 9 and 12 conducted on specimens saturated with CH 4 or C02
saturated water were characterised by both volumetric and pore pressure variations. Rad et al. (1994)
pointed out that, as tests 9 and 12 were fully undrained, the volumetric strains experienced by the
specimens must have been related to gas exsolution and expansion.

According to these authors, the possibility for volumetric expansion, and thus, partial (internal)
drainage and less intense pore pressure reduction is the reason behind the observed reduction in
shear strength of gassy sand specimens compared with the reference undrained test (curve 6 in
Figure 4-15). When comparing tests 9 and 12, it is evident that, the more soluble the gas (i.e. CO2
relative to CH4), the stronger the gas exsolution and the subsequent volumetric expansion, the less
intense the pore pressure reduction, and thus the lower the undrained shear strength.

Comparison of curves 6, 13, 16 and 19 clearly indicates that the specimen containing free gas (curve
16) exhibits a lower peak strength than the one with gas solely in dissolved form (curve 19), though,
both specimens ultimately reach relatively similar values of residual strength and pore pressure.

When comparing the results of tests 4 and 9 conducted on identical specimens saturated with C02
saturated water, at different pore pressure level, it is clear that the pore pressure at which dilative
gassy sand is saturated has a strong effect on its strength under undrained loading.

In view of the volumetric strains and pore pressures recorded during these tests, Rad et al. (1994)
emphasised that the lower the initial pore pressure, the stronger the subsequent volumetric expansion
for a given pore pressure reduction and thus, the greater the potential for lower strength under
undrained shearing.

Reference curves 6 and 13 correspond, respectively, to undrained and drained tests on specimen
saturated with deaired water (η = degree of water-gas saturation = 0). Other tests were conducted on
specimens saturated with CO2 or CH4 saturated water (η =1, curves 4, 9, 12, 19) or on specimens
containing both dissolved and free gas (Sw0 < 1, η =1, curve 16).

45 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

4.1.4.3 Mixed soils

4.1.4.3.1 Loose gassy silt sands


Following Haththotuwa and Grozic (2009) who showed that the behaviour of fully saturated silt sands
depends on silt content, drainage conditions and loading path, Haththotuwa and Grozic (2010)
performed an experimental study on the influence of the presence of plastic fines on the behaviour of
gassy Ottawa sand.

The experimental programme implied triaxial testing of CO2 charged mixtures of Ottawa sand and
Penticton silt (10 and 20% of silt by weight). After isotropic consolidation, all the specimens had low
gas content (i.e. 0.842 < Sw0 < 0.989) at the start of shearing in undrained conditions. Analysis of the
results of triaxial tests presented in Figure 4-16 shows a significant drop of peak strength of the gassy
silt sands compared to the gassy sand.

The strength reduction is the greatest at the lowest initial confining pressure of ~100 kPa, with similar
percentage strength reductions for silt contents of 10% and 20% compared with pure sand (Figure 4-
16). Haththotuwa and Grozic (2010) ascribed the drop in shear strength to the higher compressibility
of gassy silt sands compared to gassy Ottawa sand. As already noticed by Grozic et al. (1999) after
testing pure Ottawa sand specimens, the test C23 shows that for a S w0 below 0.9, the presence of gas
has the effect of changing the shear behaviour from strain softening to strain hardening.

In order to capture the critical state behaviour of gassy silt sands, Haththotuwa and Grozic (2010) paid
attention to the void ratio values as a function of the mean effective stress for the end of test
conditions.

According to Figure 4-17 presenting these results, the trend line obtained for gassy sand possesses a
shape similar to that of gassy sand with fines. It is clear from Figure 4-17 that increasing the fines
content has the effect of shifting downwards the critical state line without changing its slope.

However, when compared with the results of tests conducted under similar conditions on fully
saturated silt sands (Figure 4-17; Haththotuwa and Grozic, 2009), it appears that the presence of gas
bubbles has an influence on both the position and slope of the critical state line.

4.1.5 3.5 Drained shear strength of gassy soils


Wheeler et al. (1991) reported results of a drained triaxial testing programme carried out by Boden on
Combwich mud containing undissolved methane produced by the zeolite technique.

As shown in Figure 4-18, there appears to be no significant effect of gas on the stress-strain curves or
on the peak value of stress obtained on three samples with degree of saturation varying between
0.999 and 0.923. Besides, the results of two tests conducted by Vanhoudeusden et al. (2004) on fully
saturated and gassy silty clay samples suggest that under drained conditions, a fine-grained gassy
soil is more resistant than a saturated soil (Figure 19-b).

46 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

The same tendency is observed when comparing the stress-strain results for the virtually saturated
sample (i.e. Sw0 = 0.999) and the more gassy samples (i.e. Sw0 = 0.923) presented in Figure 4-12a.
However, the number of tests performed on fine-grained gassy soils appears too limited to draw firm
conclusion as to the effect of gas on their drained strength.

4.1.6 Unloading behaviour of gassy soils


From the early research works of Hardy and Hemstock (1963) to the most recent works of
Amaratunga and Grozic (2009) it has been widely recognised that the most detrimental effects of gas
on marine soils are encountered during unloading as it implies gas exsolution and expansion.
Because of these two processes, under undrained unloading conditions, changes in volume of gassy
soil may occur and pore pressure responses are dramatically different from that of saturated soils
(Sobkowicz and Morgenstern, 1984; Amaratunga and Grozic, 2009).

Figure 4-15. Results of triaxial tests on dense Baskarp sand specimens (Rad et al., 1994). a. Stress-
strain curves. b. pore pressure-strain curves. c. volumetric against axial strains curves.

47 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-16. Results of undrained triaxial compression test on gassy Ottawa sand and gassy
mixtures of Ottawa sand and Penticton silt Haththotuwa and Grozic (2010).
a. Effective stress paths. b. Stress-strain curves.

Figure 4-17. Void ratio versus mean effective stress plot for gassy sand, gassy silt sand and fully
saturated silt sand Haththotuwa and Grozic (2010).

4.1.6.1 Effect of unloading on gassy soil structure


Samples expanding out of core liners while brought from the seabed to the surface have provided the
most striking evidence for the effect of unloading on the structure of gassy soils (Esrig and Kirby,
1977). Efforts conducted in an attempt to avoid such sample disturbances have resulted in the
development of in situ pressure samplers (Denk et al., 1981; Johns et al., 1985 among others).

By using silty clay samples collected with such coring devices in the Mississippi Delta, Chiou et al.
(1991) performed microfabric studies on gassy soils under in situ pressures as well as under ambient
pressure, after controlled degassing.

The results of these studies revealed that samples collected and maintained under in situ pressures
were characterised by relatively well oriented clay particles even from the walls of large gas voids as
shown in Figure 20. By contrast, depressurised gassy samples were typified by random arrangements

48 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

of clay particles with a degree of randomness that appeared to decrease with overburden pressure
(Figure 21; Chiou et al., 1991). By comparing clay fabric microfeatures in different stages of sediment
degassing Chiou et al. (1991) noted that samples degassed for less than one hour still showed fairly
well preserved clay fabric compared to samples that were left degassing for more than 2 hours.

Accordingly, they concluded that the depressurising effect on the clay fabric texture in a sediment
matrix is time dependent.

Figure 4-18. Drained triaxial tests on a. fully saturated silty clay and silty clay containing methane
(Wheeler et al., 1991); and b. fully saturated silty clay and silty clay containing C0 2
(Vanhoudeusden et al., 2004).

Figure 4-19. a. Scanning electron microscope (SEM) micrograph showing a clay fabric around a oval-
shape depression created by a large gas bubble in a sampled maintained under in situ pressures.
b. SEM micrograph showing well-oriented clay fabric from the wall of the gas bubble depression in a.
(Chiou et al., 1991).

49 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-20. SEM micrograph showing the clay fabric of a depressurised sample (Chiou et al., 1991).
Somewhat oriented clay arrangement is shown in the upper portion of the micrograph whereas
randomly arranged clay flakes associated with channel-like features meandering through the central
portion of the micrograph.

4.1.6.2 Effect of unloading on gassy soil compression behaviour


Hight et al. (2002) reported the results of a series of conventional oedometer tests to highlight the
effects of gas exsolution on the compression behaviour of Nile Delta clays.

As shown in Figure 4-21, the higher the gas content, the higher the compressibility until the
compression curves become nearly parallel once the gas bubbles have collapsed. Hight et al. (2002)
ascribed the higher initial compressibility to the compressibility of gas bubbles as well as destructuring
implied by gas expanding the soil skeleton.

Johns et al. (1985) drew similar conclusions while working on Mississippi Delta sediments collected
and tested at in situ pressures as well as at ambient pressures, after controlled degassing. Figure 4-
22a presents the results of a consolidation test on a clayey sample with a in situ degree of saturation
of 0.97.

It can be seen that, after complete degassing, such a clayey sample exhibits a [e-log(σ-uw)] curve
clearly indicating greater amounts of compression for a given load than a sample maintained under in
situ pressure or a remoulded sample.

The results of this series of oedometer tests revealed large initial differences for the coefficient of
consolidation, cv, of gassy and degassed samples (Figure 4-22b).

Initially, the degassed values of cv calculated by the square root of time method may be up to 4 times
greater than the values for samples maintained under in situ pressure. However, as shown in Figure
4-22b, both in situ and degassed log and square root of time of cv tend to converge at a vertical
operative stress in excess of the effective preconsolidation pressure p’c.

50 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-22b and Figure 24-a reveal that, with some exceptions, the values of effective
preconsolidation pressure are larger for the pressurised tests than those for the degassed tests. At the
same time, Figure 4-22b shows that compressions indices, cc, are generally higher for the degassed
tests than for the pressurised ones.

Figure 4-21. Effects of gas exsolution on oedometer tests on samples of Nile Delta Clay.
(Hight et al., 2002).

Lunne et al. (2001) presented a series of constant rate of strain (CRS) oedometer tests on samples of
Lierstranda and Bothkennar clays prepared with solely dissolved gas allowing to reach degrees of gas
saturation ranging from 0 to 1. Prior to consolidation testing, these samples were unloaded to simulate
stress relief and gas exsolution due to reduced effective stresses from tube sampling and water depth
change.

As illustrated in Figure 4-24, one of the main results of the series of consolidation tests is that the
preconsolidation pressure shows some tendency for reduction with increasing degree of gas
saturation.

51 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-22. Results of oedometer tests conducted on gassy (pressurised), degassed (depressurised)
and remoulded clayey sediments of the Mississippi Delta (Johns et al., 1985). a. Comparative plots of
oedometer results in a e-log(σ-uw) graph. b. Values of the coefficient of consolidation, c v, calculated
with the Log time and Square root time methods versus the vertical operative stress.

4.1.6.3 Effect of unloading on the undrained shear strength of gassy soil

4.1.6.3.1 Fine-grained soils


In addition to oedometer tests, Lunne et al. (2001) performed a series of anisotropically undrained
triaxial tests sheared in compression (CAUC) on samples of Lierstranda and Bothkennar clays. Such
as for oedometer tests, the samples were prepared with degrees of gas saturation ranging from 0 to 1
(i.e. initially dissolved gas) and then subjected to unloading to simulate gas exsolution due to core
sampling.

The effect of gas exsolution on the results of triaxial tests on Liestranda clay is illustrated in Figure 4-
25a as causing increasing decreases in shear strength with increasing degree of gas saturation. The
Bothkennar data are less conclusive, with the sample with a degree of gas saturation of 20% seeming
to be less affected than the specimen free of any gas (Figure 4-25a).

Figure 4-25b clearly illustrates that destructuring by gas exsolution also produces a characteristic
change to the effective stress path. This manifests by a discontinuity in the effective stress path when
excess pore pressures rapidly increase with no increase in shear stress.

52 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-23. Effective preconsolidation pressure (a.) and compression indices (b.) versus depth for
samples collected and tested at in situ pressure (pressurised) and for degassed samples
(depressurised). (Johns et al., 1985).

Figure 4-24. Preconsolidation pressure versus degree of gas saturation for Lierstranda and
Bothkennar CRS tests (Lunne et al., 2001).

According to Figure 4-25b, the extent of this discontinuity increases with initial degree of gas
saturation and, therefore, with the amount of damage. Comparison of Figure 4-25b and Figure 4-26b
reveals that for similar degree of gas saturation, the discontinuity is less pronounced for the
Bothkennar clay than for the Lierstranda clay.

53 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

As suggested by Hight (1999), this difference in response to gas exsolution for the two clays could be
related to differences in structure with the pore sizes of the Lierstranda clay being larger than those of
the Bothkennar clay.

Hight et al. (2002) noted similar discontinuity in the effective stress path during undrained shearing in
triaxial compression of Nile Delta clay samples and presumed that such a characteristic change
corresponds to the collapse of the pores enlarged by gas exsolution (Figure 4-27).

Figure 4-25. Results of CAUC triaxial tests on Lierstranda clay with different degrees of gas saturation
(Lunne et al., 2001). a. Stress-strain curves. b. Effective stress paths.

Figure 4-26. Results of CAUC triaxial tests on Bothkennar clay with different degrees of gas
saturation (Lunne et al., 2001). a. Stress-strain curves. b. Effective stress paths.

54 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-27. Results of CIU triaxial tests on Nile Delta clay (Hight et al., 2002). The red segment
highlights the discontinuity in the effective stress path of samples damaged by gas exsolution.

4.1.6.3.2 Coarse-grained soils


Sobkowicz and Morgenstern (1984) presented results of laboratory tests conducted on dense triaxial
samples of Ottawa sand that did not initially contained free gas but were imbued with a gas-rich (CO2)
pore liquid.

Figure 4-28 presents the pore pressure response for a typical test which was subjected to a set of
isotropic total stress decrements. The results of this test allow to highlight the effect of the transient
process of gas exsolution initiating once the pore pressure drops below the liquid-gas saturation
pressure.

As can be seen in Figure 4-28, upon unloading the immediate pore pressure response is less than the
change in total stress (Skempton’s B parameter <1). Sobkowicz and Morgenstern (1984) ascribed this
particular response to increasing pore fluid compressibility owing to the presence of free gas but not to
any gas exsolution and noted that it is not an equilibrium response for the soil unless the gas is totally
insoluble in the pore fluid.

They emphasised that the true equilibrium response of the soil is that behaviour observed after all gas
exsolution has been completed, for a given decrease in total stress. However, both immediate and
long term pore pressure responses show a similar trend. For decreasing total stress, the short as well
as the long term values of the Skempton’s B parameter decrease, level off, then increase back to a
value of 1. According to Sobkowicz and Morgenstern (1984), the level off stage occurs as the pore
pressure reaches the liquid-gas saturation pressure.

Subsequently, the pore pressure begins to decrease when the effective stress approach 0, that is,
when σ>0 and u=σ. With a Skempton’s B parameter back to a value of 1, any further unloading

55 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

causes the generation of large volumes of gas and a disruption of the soil structure Sobkowicz and
Morgenstern (1984).

Figure 4-28. Observed undrained pore pressure response (Sobkowicz and Morgenstern, 1984).

Figure 4-29. Effective stress path for q-ramping and q-constant unloading of Ottawa sand specimens
saturated with CO2 saturated water (Amaratunga and Grozic, 2009).

56 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-30. Comparison of collapse lines derived from unloading and shearing of Ottawa sand
specimens. (Amaratunga and Grozic, 2009)

Amaratunga and Grozic (2009) furthered the works of Sobkowicz and Morgenstern (1984) by
experimentally studying the effect of q-constant unloading of Ottawa sand specimens saturated with
CO2 saturated water. Figure 4-29 shows that the response of such specimens was completely
different from that of saturated soils without any gas saturated water which, when subjected to a q-
constant unloading stress path, should theoretically not exhibit change in mean effective stress.

Amaratunga and Grozic (2009) pointed out that even though pore pressure initially dropped with cell
pressure, the ensuing gas exsolution process resulted in increases in pore pressure close to the
liquid-gas saturation pressure. Hence, the mean effective stresses of the gassy specimens reduced
with unloading so that the stress paths moved towards the left along a q-constant line. It can be seen
that all the gassy specimens collapsed during unloading, at a stress state specific to their respective
initial condition.

By joining the collapse points in a q:p’ space, Amaratunga and Grozic (2009) defined the collapse
line for gassy specimens of void ratios 0.75 to 0.79 (Figure 4-30). When comparing this collapse line
with that obtained by undrained shearing of saturated specimens of Ottawa sand, they found that the
two collapse lines coincided (Figure 4-30).

Accordingly, Amaratunga and Grozic (2009) concluded that such coincidence indicates that it may be
possible to predict the collapse of gassy-soils with a collapse surface derived from saturated-soil
testing.

57 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

4.1.7 Models accounting for the behaviour of gassy soils

4.1.7.1 Excess pore pressure and free gas generation during gas exsolution: Sobkowicz and Morgenstern
(1984) model
Sobkowicz and Morgenstern (1984) presented a theoretical model based on a closed-form solution for
undrained pore pressure responses to changes in total stress, by applying volume compatibility and
considering gas expansion and exsolution. Gas exsolution is initiated when pore pressure drops below
liquid-gas saturation pressure – the pressure at which the liquid is fully saturated with dissolved gas.

The quadratic equation describing change in pore pressure equates change in total specimen volume
to change in fluid volume and assumes negligible change in soil particle volume:

Au 2  Bu  C  0 (4-14)

where

A   T  n0 S w0  w (4-15)

(where βT is the total soil compressibility, βw, the compressibility of the water, n0 is the initial porosity
and Sw0 the initial degree of saturation)

B   T u0  Patm    n0  w S w0Patm0  1  S w0  S w0H (4-16)

(where u0 is initial gauge pore gas pressure, total stress change, Patm is the atmospheric pressure, and
Δσ is the total stress change)

C   T u0  Patm  (4-17)

This solution assumes that the exsolution process is at equilibrium, which remains exact as long as βT
is constant. In practical term, this would require that sufficient time be given during one phase of total
stress change so that at the beginning of the ensuing phase, the exsolution process would have been
completed.

However, providing that the liquid-gas saturation pressure is known for the pore fluid or, that
measurements of pressure and saturation are made at some point of the equilibrium, Sobkowicz and
Morgenstern (1984) suggested that a solution to the expected equilibrium response can be found by
adopting the following form for C:

C  u 0  Patm   T   n1  S w  S w H 


K'
VT (4-18)

where VT is the total volume of soil element and n and Sw are measured at (u0 + Patm). K’ is a constant
for the system and is equal to the total equilibrium volume of dissolved and free gas in the pore space,
measured at a given pressure P = (u + Patm).

58 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Sobkowicz and Morgenstern (1984) used equation 14 to predict the behaviour of a gassy sand sample
subjected to a set of isotropic total stress decrements as reported in Figure 4-28. Two theoretical
analyses were conducted based on equation 14. Analysis number 1 considered the effects of both air
and CO2 exsolution due to step decreases in total stress matching those imposed in the laboratory.

Analysis number 2 attempted to account for both non-equilibrium at the end of each unloading phase
and for loss of gas by diffusion through the sample membrane. Figure 4-31, which only shows the
predicted points differing the most significantly from the observed behaviour, reveals the extremely
good predictive capability of the theory for the equilibrium (or long term) pore pressure response.

4.1.7.2 Undrained shear strength of gassy soil: existing models


Several theoretical models have been developed during the last four decades to predict the behaviour
of fine grain soils partially saturated by gas. Those theoretical models can be divided into two main
groups.

The first model group was developed to predict the behaviour of sediments with low degree of water
saturation and where the gas and water phases are continuous. In this case, the effective stress
theory is no longer valid (Coleman, 1962; Matyas and Rahakrishna, 1968; Fredlund and Morgenstern,
1978) and it is essential to consider the suction (difference between pore gas pressure and pore water
pressure) as an additional independent stress state variable.

Figure 4-32 illustrates schematically the invalidity of the effective stress theory for this first model
group and where, under a constant effective stress, a change of the capillary pressure induces a
decrease of the void ratio. Therefore, there is no unique relationship between void ratio and effective
stress as it was defined by Terzaghi, in 1927.

The second model group was developed to predict the behaviour of sediments with high degree of
water saturation (i.e. Sw > 0.8): a saturated matrix contains occluded gas bubbles and large
macropores are filled by gas (see for instance Wheeler, 1988b and Grozic et al., 2005).

In this case the water phase is continuous and the principle of effective stress remains valid. This
second group of models has often used the effective stress theory by including the gas compressibility
and the gas solubility in the theoretical developments.

59 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-31. Comparison of measured and predicted pore pressures for an undrained unloading test
on dense triaxial samples of Ottawa sand (Sobkowicz and Morgenstern, 1984).

Figure 4-32. a) Theoretical stress path in q-p’-pc space and b) theoretical void ratio changes
versus effective stress.

60 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

4.1.7.2.1 Barcelona model (Alonso et al., 1990; Gens et al., 1992)


For sediments with continuous gas phase (low degree of water saturation, Sw <0.8) models proposed
by UPC in Barcelona (Alonso et al., 1990; Gens et al., 1992) have been used in the past to accurately
predict the hydro-mechanical behaviour. The Barcelona Basic Model (BBM, Alonso et al., 1990) is
able to predict volumetric strains and shear strains for any stress paths under drained conditions.
However, it cannot be used to predict the undrained behaviour because it does not provide any
information on the variation of the degree of saturation (Wheeler, 1996; Wheeler et al., 2003). In order
to take into account the behaviour of a gassy sediment during sudden unloading (undrained
conditions), additional equations considering the change of the degree of saturation during different
stress paths have to be included in the BBM (see for instance: Gallipoli et al., 2003; Wheeler et al.,
2003; Sheng et al., 2004; Vanoudheusden et al., 2004).

The use of the Barcelona model modified for undrained tests to simulate the observed behaviour
obtained within the present project seems inappropriate because:

- the main Barcelona model parameters are determined at constant capillary (or suction) pressures
whereas gas exsolution process (in natural environment and laboratory) occurs at constant water
content.

- for gassy soils, where the free gas is expected to be in the form of occluded bubbles, the lowest
degree of water saturation is expected to be greater than 0.8 and therefore the use of a new state
variable (water capillary pressure or suction), where the free gas phase is supposed to be
continuous, seems inappropriate to describe correctly the behaviour of the gassy sediments.

4.1.7.2.2 Grozic, Nadim and Kvalstad model


Grozic et al. (2005) presented a constitutive model for determining the undrained shear strength of
fine-grained gassy soils by accounting for the peculiar pore pressure response due to the presence of
free gas. Following Fredlund and Rahardjo (1993), in the model proposed by Grozic et al. (2005), the
theoretical pore pressure response due to change in total stress is calculated by considering the initial
and final conditions of an applied increment of total stress in undrained loading.

To this aim, the compressibility of the pore fluids as described in equation 21 (Fredlund and Rahardjo,
1993) is used to determine the volume change. In order to account for the effect of gas on the
undrained shear strength of fine-grained soils, Grozic et al. (2005) incorporated, into a Cam Clay
model, the equations 21 and 22 which describe the volume change due to the compression and
solution of gas in term of change in void ratio, Δe :

 du   dug 
 gw  S w   w   w   1  S w  hS w    g  
 d   d  (4-19)
 

where βgw is the compressibility of the water/gas mixture, βg is the compressibility of gas, h is the
volumetric coefficient of solubility (i.e. the ratio between the volume of dissolved gas and the volume of

61 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

water), duw/dσ is the water pressure change with respect to a total stress change, dug/dσ is the gas
pressure change with respect to a total stress change, and σ is the total stress.

  ug 
e     1  S w 0  k i S w 0 e 0
 ug0   ug  (4-20)
 

where  ug , is the average change in pore gas pressure, ug0 is the average initial pore gas pressure,

ki is the Henry’s constant. Equation 19 is valid for dissolved gas and free gas in the form of occluded
bubbles, that is for Sw greater than 0.8.

In order to verify the reliability of their fine-grained gassy soil model, Grozic et al. (2005) compared
model predictions with triaxial tests results on Combwich mud presented by Wheeler et al. (1998).
Figure 4-33-a, -b & -c show that the model predictions reasonably fit with Wheeler’s results.

Grozic et al. (2005) ascribed the difference between model and experimental results to differences in
the Henry’s coefficient of solubility. It is however observed in Figure 4-33-d, -e and -f that the results of
the model predictions in term of undrained shear strength plot higher than the experimental results.
The best match was obtained for the higher operative stress results showing increases in strength due
to compression and solution of the gas (Figure 35-d).

Besides, the model predictions greatly differ from experimental results obtained for operative stresses
of 200 and 100 kPa. The worth prediction is observed in Figure 4-33-f revealing that the model
predicts increase in shear strength with decreasing degree of saturation while the exact opposite trend
was experimentally obtained by Wheeler (1988a). As pointed out by Grozic et al. (2005) such a
discrepancy may be ascribed to the model simplifications assuming that the change in pore gas
pressure is equal to the change in pore water pressure which disregards Wheeler’s key findings.

4.1.7.2.3 Wheeler’s model


Wheeler (1986, 1988a, 1988b) proposed a theoretical model attempting to deal with the presence of
discrete gas bubbles in fine-grained soils. The basis of this model is that, since in fine- grained soils,
the individual soil particles are generally much smaller than the bubble cavities (see Figure 4-34-a),
the water and solid phases may be combined to form a single phase of saturated soil which can be
treated as a continuum surrounding large isolated gas bubbles (Figure 4-34-b).

In this view, the bubble cavities form internal boundaries to the saturated soil matrix, with the boundary
stress on these internal surfaces given by the gas pressure u g. By analysing this model, Wheeler
(1986, 1988a) derived theoretical expressions for the normalised undrained shear strength S u/Susat
(where Susat is the undrained shear strength of a saturated soil consolidated to the same value of
operative stress σ3-uw0).

62 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-33. Comparison of predictions of a Cam Clay model modified by Grozic et al. (2005) with
experimental tests results of Wheeler (1988a).

Figure 4-34. a. Schematic representation of the structure of a soil containing large bubbles. b.
Idealised model of a soil containing large bubbles (Wheeler 1988b).

For the analysis of undrained shear strength, the saturated matrix was assumed to be a rigid, perfectly
plastic von Mises-type material following the work of Green (1972) who proposed an elliptical yield

63 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

curve for material containing empty spherical cavities. According to this analysis, Wheeler (1986,
1988a) acknowledged that an exact relationship for S u/Susat cannot be obtained, but that useful upper
and lower bound limits can be approximated by assuming a specific type of failure mechanism. Hence,
by ignoring the increase in strength caused by bubble flooding and shrinkage, so that Su is simply half
the value of the yield stress qy, Wheeler proposed the following theoretical lower bound expression,
consisting of a quadratic equation for Su/Susat:

 Su 
2
 3  3  u g0  Su   2

a  b   (4-21)


 Su sat   2Su sat Su sat 

where both a and b are functions of the volume fraction of gas bubble f:

2
 3  2f 01/ 4 
a

 3 1  f 0
1/ 3



(4-22)

2
 1 
b  (4-23)
 2 loge f 0 

By arguing that the contact angle of the menisci between the wetted surface of a clay particle must be
zero, Wheeler et al. (1990), showed that the menisci between the gas and water must be concave
when viewed from inside the bubble and therefore the gas pressure ug is always greater than the pore
water pressure uw. Thus, in the lower bound expression, the value of the initial gas pressure u g0 can,
conservatively, be replaced by that of the initial water pressure u w0:

 3 3  u w 0 
2 2
 Su  Su 
a  b   (4-24)
 Su sat   2Su sat Su sat 

Accordingly, the lower bound values of normalised strength predicted by equation 4-21 depend not
only on the value of f controlling the factors a and b, but also on the soil type and overconsolidation
ratio which control the value of Susat/(σ3-uw0). Figure 4-35 shows how the theoretical lower bound
varies with the soil type (high or low plasticity index) and stress history (OCR value).

Wheeler (1986, 1988a) argued that the two processes of bubble shrinkage and bubble flooding tend to
increase the value of Su above the lower bound in equation 4-21. However, the maximum possible
value of Su does not occur when the two processes act simultaneously but rather when one of the
processes occur without interference from the other.

Wheeler showed that the simplest upper bound expression for S u/Susat may correspond to complete
closure of bubbles cavities resulting in a saturated soil such that:

Su
1 (4-25)
Su sat

64 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-35. Theoretical lower bound curves for undrained strength for different soil types and
overconsolidation ratios (Wheeler et al., 1991).

Alternatively, by considering complete flooding of all the bubbles cavities without any shrinkage
Wheeler derived a theoretical upper bound expression for S u/Susat in terms of gas volume fraction and
the parameter (1+em0)/λ, where em0 is the void ratio of the saturated matrix at the start of shearing,
and λ, the slope of the critical state line in e:log e(p-uw) space for a saturated soil:

  f 
1/ 3 
3 1   0 


  1  f0    1  e m0 f 0 
Su   (4-26)
 exp 
3  21  f 0   1  f 0  
Su sat 1/ 4

The actual upper bound value of Su/Susat is given by either equation (25) or (26) whichever is the
greater.

Figure 4-36-a shows comparison between experimental values of undrained shear strength of
normally consolidated Combwich mud samples and theoretical lower and upper bound curves
calculated for these samples. It can be seen that experimental results fit neatly within the theoretical
upper and lower bounds, providing evidence for the validity of the conceptual model.

A similar good agreement is obtained between the theoretical bounds and experimental results of
Sham (1989) who extended the measurements of Su to higher values of total stress, by conducting a
series of triaxial compression tests on isotropically consolidated samples of kaolin containing helium
bubbles (Figure 4-36-b). It is however noteworthy that the lower bound theoretical expression is
conservative in its description of the undrained strength loss.

The use of the Wheeler (1988b) model to simulate the experimental data obtained from the JIP
“Gassy Soil” (Sultan et al., 2009) show that the Wheeler model can predict correctly the undrained

65 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

shear strength of gassy sediments if the sediment is directly sheared after the gas exsolution process.
However, for more complex stress paths it was shown that:

- it is essential to combine flooding process to the damage;

- if consolidation occurs after gas exsolution, the Wheeler (1988b) model needs to be modified in
order to include the shrinkage and collapse of the gas bubbles and damaged zones during the
subsequent loading.

Figure 4-36. Experimental and theoretical values of undrained shear strength. a. CH 4 charged
Combwich mud (Wheeler 1988a). b. He charged Kaolin (Sham 1989).

66 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

5. COUPLING SATURATED AND GAS EFFECTS MODELS AND PARAMETER IDENTIFICATION

5.1 Introduction
Quantitative studies of the dynamics of gas hydrate in marine sediments can be grouped into two
categories. The first of these models take into account methane conservation, methane advection-
diffusion coupled with heat transfer (Rempel and Buffett (1998) and Xu and Ruppel (1999)). An
alternative simpler category, deals with models accounting for thermal non steady-state regime
(Chaouch and Briaud (1997) Delisle et al. (1998)). By neglecting the effect of gas components and
concentration on the gas hydrate stability law, the second category of models are inadequate by
regarding mainly the following three points:

- the low concentration of gas in the upper meters of sediment, which can be related to the methane
exchange between bulk water and seawater column, can prevent the formation of gas-hydrates.
Thus, the gas-hydrate stability zone from p-T conditions does not coincide with the hydrate
occurrence zone.

- the hydrate fraction, which depends on the gas concentration within the sediment column, is
often improperly considered as constant.

- the excess pore pressure generated by the melting of the gas hydrate depends on i) the hydrate
fraction and ii) on the gas solubility. By considering only the energy conservation equation, it is
impossible to evaluate this excess pore pressure.

Therefore, a numerical model of the formation or dissociation of gas hydrate, which takes into account
the influence of temperature, pressure, pore water chemistry, and the pore size distribution of the
sediment is developed. This model fully accounts for the latent heat effects. The model allows for the
evaluation of the excess pore pressure generated during hydrate melting using the Soave’s equation
of state.

The mechanical behaviour of marine sediments from the Gulf of Guinea has been characterised at
CERMES during a previous project, with a par- ticular focus on time effects and structure degradation.
The experimental campaign included isotropic and triaxial tests on intact and reconstituted marine
sediments. Except for the isotropic compression tests performed at controlled rate of stress, the tests
were performed at controlled and varied rates of strain. Experimental results have been published in
Géotechnique and interested readers are referred to Sendir-Torisu et al. (2012) for more details.

During this project, a constitutive model accounting for these phenomena (that is effects of strain rate
and structure on the mechanical behaviour of marine sediments) has been developed. The model is
presented in detail in the final report of the project, see Pereira et al. (2010). In parallel, an other
constitutive model has been developed by IFREMER to account for the presence of gas and its effects
from the mechanical point of view, when gas exsolves from the interstitial water phase.

As a sequel of the first project described above, the present project aims at merging both modelling
frameworks and at simulating realistic cases encountered in practice by offshore geotechnical
engineers. This report present the theoretical and numerical developments carried out to date. Its aim

67 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

is to present the constitutive framework that has been developed to merge the two models developed
previously and to discuss the imple- mentation of the complete model into a Finite Element Code
developed at Laboratoire Navier, namely the code BIL.

5.2 Modelling structure and its degradation: S3-SKH model


The saturated model accounts for the effects of strain rate (based on RAS- TRA model: (De Gennaro
et al., 2009; Pereira and De Gennaro, 2009; De Gennaro and Pereira, 2013)) and of soil structure and
its degradation (based on S3-SKH: (Baudet and Stallebrass, 2004)) on the mechanical be- haviour of
the sediments. RASTRA and S3-SKH models have been merged in a previous project and will be
briefly presented hereafter (see (Pereira et al., 2010) for a more detailed presentation).

The 3-SKH model was developed by Stallebrass and Taylor (1997) to simulate the non-linearity of the
stiffness of clays. The additional surface to the original “bubble” kinematic hardening model (Al
Tabbaa and Muir Wood, 1989) allows effects of recent stress history on the stiffness to be modelled.
For example, a soil element subjected to a stress reversal will have a much higher stiffness than a soil
element that is simply sheared in a single direction (Figures 6-1 and 6-2).

This feature is well captured by the 3-SKH model but not in simpler single surface models. Principal
stress directions can change significantly during the loading history of a site, and such changes in
stiffness can occur. It is common in practice to reset the stiffness to high values when the stress
direction changes, but if a model like the 3-SKH model is used the change in stiffness is included in
the formulation of the model, which simplifies matters and saves on computing time.

Figure 5-1. Effect of recent stress history on stiffness (Stallebrass and Taylor, 1997)

68 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 5-2. Schematic diagrams of MCC and 3-SKH yield surfaces.

The S3-SKH model was developed by Baudet and Stallebrass (2004) to simulate soft structured clays.
It is based on the 3-SKH model, but in- cludes structure and degradation of structure in its formulation.
The main parameters of the S3-SKH model are the same as in the 3-SKH model, and they are derived
from test data on the reconstituted clay. This is convenient as tests on reconstituted soil are more
easily prepared and they are repeatable.

There are three additional parameters needed to describe structure and its degradation that are
derived from test data on intact soil. In the aim to include parameters that have a physical meaning,
the Sensitivity framework proposed by Cotecchia and Chandler (2000) was used as a theoretical base
to include structure in the model (Figure 6-3).

In the Sensitivity framework, a correspondence between the effects of structure in the stress and
volumetric planes is suggested, so that structured clay has a state boundary surface that has the
same shape as that of its reconstituted counterpart, but is larger in size by a factor equivalent to
sensitivity. The S3-SKH model follows this up by using sensitivity (initial sensitivity, s0) as one of the
ad- ditional parameters to simulate structure (see Figures 6-4 and 6-5).

Sensitivity can be derived from laboratory tests (isotropic compression, undrained compression) or
field tests. An ultimate sensitivity (sf) is also included, to account for effects of fabric that are not
removed with reconstitution. The value of sf is however equal to unity in most clays, i.e. the intact clay
converges to reconstituted states at large strains.

The third parameter, k, is used to describe the degradation of structure, or sensitivity, with strain level.
This is less straightforward to derive, and a typical method would be to use trial and error on isotropic
compression data.

69 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 5-3. Sensitivity framework (Cotecchia and Chandler, 2000).

Figure 5-4. Schematic diagram of S3-SKH model (Baudet and Stallebrass, 2004).

70 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 5-5. Definition of S3-SKH model parameters describing structure: initial and ultimate
sensitivities, rate of degradation.

Figure 5-6. Ratio of influence of plastic shear to volumetric strain in damage strains used in different
models for structured soils.

71 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Table 1. Parameters of the Modified Cam Clay model

Parameter Description
N intercept (at pf =1kPa) of the NCL in v − ln pf plane
λ gradient of the NCL in v − ln pf plane
κ gradient of swelling lines in v − ln p f plane
M gradient of Critical State Line (CSL) in q − p f plane
f
p0 size of the bounding "yield" surface (isotropic pre-consolidation
pressure)

Table 2. Parameters of the Sensitivity 3-Surface Kinematic Hardening


model: S3-SKH

Parameter Description
N∗ intercept (at pf =1kPa) of the NCL in ln v − ln pf plane
λ∗ gradient of the NCL in ln v − ln pf plane
κ∗ gradient of swelling lines in ln v − ln pf plane
M gradient of Critical State Line (CSL) in q − p f plane
f
p0 centre of the bounding “yield” surface; corresponds to half the
isotropic pre-consolidation pressure
T ratio of size of history surface to size of bounding surface
S ratio of size of yield surface to size of history surface
ψ stiffness degradation coefficient
s0 initial sensitivity (degree of structure)
sf ultimate sensitivity (generally equal to 1)
krate of structure degradation with strain level

72 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

5.3 Modelling the effects of rate of strain: RASTRA model


The “isotach approach” proposed by Šuklje (1957) is schematically sum- marised in Figure 6-7. The
state of the soil specimen is described by the relationship (p f: ev : e˙v) among stress-strain-strain rate
(see also Leroueil, 2006). Following this approach soils may experience both compression un- der
constant strain rate (i.e. following a rate-line) or delayed compression under sustained constant load
(i.e. decreasing the strain rate and crossing multiple rate-lines).

The amount of delayed strain accumulated during this latter phase will depend on the elapsed time of
sustained loading imposed to the soil. At the end an equilibrium void ratio is obtained. It is worth noting
that each isotach (rate-line) has an equivalent rate- dependent yield stress (pc, pc1, pc2 . . . ) which is
larger for the higher applied strain rate.

Leroueil (2006) observed that an average increase in the yield stress from 7% to 12% per log cycle of
strain rate is generally obtained for inorganic clays. In the usual range of interest of the applied strain
rates during one-dimensional compression tests at Constant Rate of Strain (CRS tests) (i.e. e˙ varying
−8 −1 −4 −1
between 10 s and 10 s ) the following relationship between the yield stress p0 and the strain rate
e˙ holds true (Ladd, 1971):

log(p 0 )  A   log  (5-1)

where A and α are materials constants.

Mesri and Godlewski (1977) showed that each class of soil is characterised by typical α values (Table
3). The validity of Equation (3) and the relationship α = C αe/Cc has been verified for a rather wide
range of geomaterials and strain rates (Leroueil et al., 1985). Equation (3) and relationship α = C αe/Cc
will be the tenet of the theoretical framework proposed in the next section for the constitutive modelling
of time-dependent behaviour of geomaterials.

Figure 5-7. Ratio of influence of plastic shear to volumetric strain in damage strains used in different

73 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Table 3. Parameter α for various geomaterials (data from Mesri and Godlewski, 1977;
Leroueil, 2006).
Material α = C α e /C c

Granular soils, rockfill 0.02 ± 0.01


Shale and mudstone 0.03 ± 0.01
Inorganic clays and silts 0.04 ± 0.01
Organic clays and silts 0.05 ± 0.01
Peat and muskeg 0.06 ± 0.01

Time effects include various phenomena (strain rate effects, creep, ageing, thixotropy, etc.). The
RASTRA model presented here only accounts for strain rate effects and may simulate creep stages
for normally consoli- dated soils.

The rate dependent model RASTRA (RAte of STRAin) has been developed adopting the “isotach” or
“reference time” models proposed by Šuklje (1957) and Bjerrum (1967) within the general theoretical
frame- work based on the theory of strain hardening plasticity. For the sake of simplicity, the
formulation of the RASTRA model will be presented in the triaxial stress space.

The Modified Cam-Clay model (MCC) proposed by Roscoe and co-workers (Roscoe et al., 1958,
1963) has been chosen as the reference constitutive framework enabling to encompass in a relatively
simple and straightfor- ward formulation the effects of time on the mechanical behaviour of ge-
omaterials. In triaxial stress space, the stresses are described by means of the effective stress pf and
the deviatoric stress q.

The strain tensor is decomposed into volumetric and deviatoric components, ev and eq respectively.
Thus, the elastic behaviour is expressed as follows:

dp f 1
deev  f
, deeq  (5-2)
vp 3G

where κ is the elastic stiffness parameter for changes in mean effective stress; G is the shear
modulus; v = 1 + e is the specific volume (e being the void ratio). The yield surface is defined by the
following expression:

   
F  f , p 0  q2  M2p f p 0  p f  0 (5-3)

where p0 is the isotropic yield stress.

5.4 Coupling RASTRA model and S3-SKH model


Constitutive models have been developed for structured clays (e.g. Baudet and Stallebrass, 2004)
independently of time.

74 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

In parallel, constitutive models for strain rate and time effects have typically been derived for that
purpose only i.e. applicable to reconstituted clays with no consideration of structure (e.g. Yin and
Graham, 1999; Bodas Freitas et al., 2011; Clarke and Hird, 2012).

Even if there is a wealth of data showing the effects of strain rate and time on natural clay behaviour
(e.g. Graham et al., 1983; Vaid and Campanella, 1977; Leroueil et al., 1996; Sorensen et al., 2007),
very few modelling work accounting for both features has been done to date. Oka and Kimoto (2005);
Karstunen and Yin (2010) assume that the visco-plastic strain contributes to damage strain that
causes degradation of structure.

Hinchberger et al. (2010) model the degree of structure as the ratio of viscosity of natural soil to
reconstituted soil. Tatsuoka et al. (2008) include ageing in their model’s formulation, mostly applicable
to artificially cemented soil. Apart from Tatsuoka et al’s unique formulation that includes permanent
and non-permanent effects of strain rate, models generally use Perzyna’s overstress theory (i.e.
isotach type of behaviour).

When developing a model accounting for both structure and time effects, the most important question
concerns the interplay between the degree of structure and the time sensitivity of the material’s
response to mechani- cal loadings.

However this has to be validated according to experimental data. Data available in the literature and
tests performed at CERMES (see Pereira et al., 2010; Sendir-Torisu et al., 2012) on marine sediments
from the Gulf of Guinea have been analysed for that purpose.

The main conclusion are presented hereafter, starting by the role played by soil’s structure on its
viscous response and continuing by the effects of strain rate on soil’s structure and its degradation.

5.5 Effects of structure on the viscous response


Concerning the effects of structure on the viscous response of soils, the fol- lowing comments can be
made. Sorensen et al. (2007) tested intact and reconstituted London clay.

Triaxial compression tests showed an isotach (permanent effects of strain rate) behaviour in natural
state but a gen- eral TESRA (increasingly non-permanent effects of strain rate) response when
reconstituted clay is tested. Interestingly, this latter observation is not obtained during isotropic
compression tests on reconstituted London clay for which an isotach behaviour is observed.

For Kitan clay (Japan), Li et al. (2003) observed different amounts of creep axial strain in natural and
reconstituted clay under same stress (higher creep strain in reconstituted clay).

For Kitan clay and Oimachi clay (Japan), Komoto et al. (2003) observed different gains in strength with
increase in strain rate in natural and reconstituted clay.

Less clear evidences are observed for GoG sedi- ments. However, a slight tendency to switch from
isotach to TESRA type of viscous behaviour has been observed.

75 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

At least, one can conclude from the available experiments on GoG clays that rate effects (or soil
sensitivity to changes of rate of strain) seem to decrease with accumulated strain.

As a consequence, the rate sensitivity parameter α could be assumed to depend on the damage strain
according to consider the following relationship:

e d    0 exp bed  (5-4)

where ed is the accumulated damage strain.

5.6 Effects of strain rate on soil structure


Concerning the effects of strain rate on the degradation of the degree of structure, the following
comments can be made.

A quantitative estimation of the combined effects of structure and strain rate can be obtained using the
sensitivity as a proxy for the degree of structure of the soil. In this re- port, it is assumed that the
sensitivity can be obtained using the following equations for either triaxial tests (for which peak
strength is determined) or for isotropic tests (for which yield stress is determined):

qpeak p if,y
s t,q  *
, s t,  (5-5)
qpeak p if,y *

Computed values of st,q or st,σ as a function of total (accumulated) strain for different soils and strain
rates are reported in Figures 5-8, 5-9 and 5-10. From these analyses, it can be concluded that the
most important factor affecting the degradation of structure is the accumulated strain and not the strain
rate at which the specimen is loaded.

In a first step or when not enough experimental data is available, parameter α describing the
sensitivity of the sediment to time effects1 is supposed to be independent of the degree of structure.

Figure 5-11 shows the coupled effects of structure and strain rate changes on isotropic compressions.
A soil initially structured (initial degree of structure s 0 = 3) is compared to a soil with no initial structure.
Dotted lines represent compression tests at constant rate of strain at three different rates.

Continuous lines represent isotropic compression tests performed at varying strain rates. The coupled
model is thus able to capture both effects of destructuration of the initial soil structure and the
sensibility of the soil to strain rate changes.

Since these simulations were performed using a constant α parameter (no effect of structure on the
viscous response), the three trend curves plotted in dot- ted lines keep a constant “distance” between
each other.

76 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 5-8. Evolution of strength sensitivity as a function of total strain and effects of strain rate
changes on London clay (experimental data from Sorensen et al. (2010)).

Figure 5-9. Evolution of strength sensitivity as a function of total strain and effects of strain rate
changes on Grand Baleine clay (Québec) (experimental data from Lefebvre and LeBoeuf (1987)).

77 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 5-10. Evolution of strength sensitivity as a function of total strain and effects of strain rate
changes on GoG sediments (experimental data from Sendir-Torisu

et al. (2012)).

5.7 Implementation of the models in a FEM code


This chapter corresponds to the task 3 of the project. The aim is to present the implementation of the
coupled model into a Finite Element code. BIL code, developed at Laboratoire Navier by Patrick
Dangla has been chosen for its modular nature which permits accessing to the solver to add new
balance equations. This was necessary since including the gaseous phase requires solving the
corresponding mass balance equation.

The most important stage of the implementation consists in extending the code to solve the three
balance equations corresponding to the mechanical problem (balance of momentum) and the fluids
phases, namely balance of water mass and balance of gas mass. The corresponding unknowns are

78 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

the displacement vector (3 components in 3D problems), the pressure of the liquid phase and the
pressure of the gaseous phase, respectively.

This equation, whose primary variable is the displacement vector, is coupled with the water and gas
pressures through the definition of the effective stress and the effects of the presence of gas that
appear in the definition of the yield stress and the apparent compressibility of the fluid mixture. The
elastoplastic models included in the FEM code include rate and time effects on one hand and gas
effects on the other hand. Merging all these components is ongoing work.

Figure 5-11. Coupled effects of structure and strain rate effects.

Two equations are implemented, one for each phase. The phenomena in- cluded in the solved
equation include:

- state equations of the fluids;

- hydraulic constitutive equation (water retention curve);

- Henry’s law, expressing the dependency of gas solubility as a function of gas pressure;

- mechanical coupling (through porosity changes);

- convective and advective fluxes of water or gas through generalised Darcy equations;

- Fick’s diffusion of dissolved gas in the water phase;

It is worth noting that the definition of the water retention curve, that expresses the evolution of the
degree of saturation of water as a function of the capillary pressure3 is a requirement to close the
problem when both water and gas pressures are unknowns. Some correlations between the water
retention curve and the pore size distribution or basic geotechnical properties of soils have been
proposed in the literature.
79 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

5.8 Numerical results


Figures 5-12 and 5-13 shows results using a model accounting for both strain rate and structure
effects. It can be observed that the trends obtained from the simulations are conform to the theoretical
frameworks included in the models RASTRA and S3-SKH and qualitatively correspond to
experimental observations on soils.

Figure 5-12. RASTRA formulation: influence of strain rate on void ratio–effective stress response.

Figure 5-13. Illustration of structure effects and its degradation: influence of initial degree of structure
on void ratio–effective stress response.

Simulations using Modified Cam Clay model and the effects of the presence of gas are now
presented. The water retention curve used in the cal- culations is presented in Figure 15. It is worth
noting that no hysteresis effects are accounted for.

80 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 16 shows the evolution of the degree of saturation during undrained unloading of a clay with
dissolved gas at initial state. Three tests corre- sponding to three different unloading increments are
presented. It can be observed that all the responses follow the same curve, with the higher un- loading
giving lower final degrees of saturation in water. These unloading stages were followed by isotropic
compressions. Results are shown in Fig- ure 17. It can be seen that the yield stress decreases for the
specimens that were previously subjected to higher unloading steps, which corresponds to the model
assumptions and the experimental observations.

Figure 5-14. Water retention curve used to simulate gas effects on the mechanical behaviour of a soil.

Figure 5-15. Undrained unloading of a gassy clay: evolution of the degree of saturation.

81 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 5-16. Mechanical isotropic compression after undrained unloading of a gassy soil.

82 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

6. CONCLUSIONS
After a review of previous studies from the related literature, it was clear that existing theoretical
models cover only limited parts of the main aspects of the behaviour of fine-grained gassy soils. Thus,
a new simple constitutive model was proposed to capture the effect of gas exsolution and expansion
on the mechanical behaviour of fine-grained gassy soils.

An analytical expression was derived to link the preconsolidation pressure to a damage parameter
which depends on the degree of water saturation. In the deviatoric plane, the analytical expression of
the preconsolidation pressure was coupled to the Banerjee et al. (1985) model which was developed
in order to take into account the initial anisotropy of natural clay and the enhanced anisotropy
generated by different mechanical stress paths.

The proposed model involved only one additional parameter with respect to classical constitutive
models for saturated soils. The three main features reproduced by the proposed model are
summarized below:

- Undrained unloading causes a drastic decrease in effective stress. Parameters governing such
behaviour include the in situ state of stress, the liquid/gas saturation pressure, the gas solubility,
the degree of saturation and the sediment and liquid compressibility.

- Consolidation processes following gas exsolution imply a low preconsolidation pressure for the
fine-grained gassy sediments with respect to the saturated one. The amount of damage caused by
exsolution is a key parameter in the proposed model.

- The shear strength of fine-grained gassy sediments following exsolution and subsequent
consolidation is affected by the damage caused to the soil structure during exsolution. These
sudden jumps in pore pressure are related to the collapse of the pores damaged by gas
exsolution.

The developed model was used to simulate results of tests on fine-grained gassy soils reported in
literature. Two set of data were considered and the proposed model was able to reproduce the
following four main phases characterizing the stress paths of gassy soils in undrained shear tests:

- The collapse during shearing of damaged zones generated by gas expansion and exsolution
implies an important increase in pore pressure.

- The presence of free gas bubbles during elasto-plastic undrained shearing increases the
compressibility of the bulk fluid and induces an only slight increase in pore pressure.

- After the complete collapse and disappearance of gas bubbles, the gassy sediment behaves as a
saturated one, with an important increase in pore pressure during shearing.

- The drained and undrained peak shear strengths increases with the initial degree of water
saturation.

83 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

The proposed model can be improved by considering more accurately the compressibility and
consolidation of the gassy soils as a function of the damage generated during gas exsolution and
expansion. In the present work two model classes were considered (with and without collapse) while
those two classes must be considered as limiting conditions to the gassy soil behaviour, an
intermediate behaviour depending on the state of the sediment structure (Lunne et al. 2001) should be
introduced.

The modelling framework presented here is based on experimental data on marine sediments. It
accounts for the effects of strain rate, of soil structure and its degradation, and also of the presence of
gas which degrades the solid matrix. We presented some improvements to the saturated model that
add a better coupling between strain rate and structure effects. We finally described the coupling
strategy that enables us to derive the com- plete model for gassy soils.

The implementation of the model in an in-house FEM code has also been presented. Some illustrative
results have been presented. They show reasonable trends. The next step is to merge the saturated
elastoplastic model (strain rate and structure effects) and the elastoplastic model for gassy soils.

The numerical tool developed in this project is quite modular and could be used or easily extended to
study several advanced geotechnical applications, such as unsaturated soil mechanics or
geomechanical applications, such as reservoir geomechanics.

For engineering and foundation design activities involving gassy sediments it is essential to keep in
mind that flooding of damaged areas and gas bubbles and dissolution of free gas do not imply an
immediate shear strength recovery of the sediments. Indeed, shrinkage and collapse of the dam- aged
zones generated by gas exsolution and expansion can be in part irreversible, and may lead to a
permanent decrease of the sediment strengths.

84 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

7. REFERENCES

Abegg F., Anderson A.L. (1997). The acoustic turbid layer in muddy sediments of Eckernförde Bay,
Western Baltic: methane concentration, saturation and bubble characteristics. Marine Geology no.
137, pp. 137-147.

Alonso, E. E., A. Gens, A. Josa, (1990). A constitutive model for partially saturated soils..
Geotechnique, 40, 405- 430.

Amaratuunga A., Grozic J.L.H. (2009). On the undrained unloading behaviour of gassy sands.
Canadian Geotechnical Journal 46 (10), 1267-1276.

Anderson A.L., Hampton L.D. (1980). Acoustics of gas-bearing sediments I. Background and II.
Measurements and models. Journal of the Acoustical Society of America. vol. 67, pp. 1865-1903.

Anderson A.L., Abegg F., Hawkins J.A., Duncan M.E., Lyons A.P. (1998). Bubble populations and
acoustic interaction with the gassy sefloor of Eckernförde Bay. Continental Shelf Research, vol. 18,
pp. 1807-1838.

Banerjee, P.K., Stipho, A.S. et Yousif, N.B. (1985). A Theoretical and Experimental Investigation of the
Behaviour of Anisotropically Consolidated Clay. Developments in Soil Mechanics and Foundation
Engineering- 2 (Elsevier Applied Science Publishers), 1-41.

Baudet, B. A. and Ho, E. W. L. (2004). On the behaviour of deep-ocean sediments. Géotech- nique,
54(9):571–580.

Baudet, B. A. and Stallebrass, S. (2004). A constitutive model for structured clays. Géotechnique,
54(4):269–278.

Bjerrum, L. (1967). Engineering geology of Norwegian normally-consolidated marine clays as related


to settlement of buildings. Géotechnique, 17:81–118.

Boudreau, B.P., Algar, C., Johnson, B.D., Croudace, I., Reed, A., Furukawa, Y., Dorgan, K.M.,
Jumars, P.A., Grader, A.S. 2005. Bubble growth and rise in soft sediments. Geology 33 (6).

Broek, D. 1983. Elementary engineering fracture mechanics. 3rd ed. Martinus Nijhoff Publisher. The
Hague, Netherlands.

Chiou, W.A., Bryant, W.R., Bennett, R.H., (1991). Clay fabric of gassy submarine sediments. In:
Bennett, R.H., Bryant, W.R., Hulbert, M.H. (Eds.), Microstructures of Fine-Grained Sediments: From
Mud to Shale. Springer, New York, pp. 333-352.

Coleman J.D., (1962). Stress strain relations for partly saturated soil. Géotechnique 12(4), 348-350.

85 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Colliat J.L., et al. (2010). Gulf of Guinea deppwater sediments: geotechnical properties, design issues
and installation experiences. Frontiers in Offshore Geotechnics II. CRC Press.

De Gennaro, V. and Pereira, J. M. (2013). A viscoplastic constitutive model for unsatu- rated
geomaterials. Computers and Geotechnics, accepted.

De Gennaro, V., Pereira, J. M., Gutierrez, M. S., and Hickman, R. J. (2009). Viscoplastic modelling of
fluids filled porous chalks. Italian Geotechnical Journal, 1/2009:44– 64.

De Groot M. B., Bolton M.D., Foray P., Meijers P.A., Palmer C., Sandven R., Sawicki A., Teh T.C.
(2006). Physics of liquefaction phenomena around marine structures, Journal of Waterway, Port
Coastal Ocean Engineering, 132(4), 227- 243

De Vries, M.H., Svanø G., Tjelta, T. I. 2007. Small scale model testing of gas migration in soft seabed
as a basis for developing a mechanical model for gas migration. In Proecedings of the 6th
International conference on Osshore Site Investigation and Geotechnics, London, 11-13 Sepctember
2007. Society for Underwater Technology, London, 231-236.

Denk, E., Dunlap, W.A., Bryant, W.R., Milberger, L.J , Whelan, T.J. (1981). A pressurized core barrel
for sampling gas-charged marine sediments. In Proceedings of the 13th Annual Offshore Technology
Conference, p. 43-52.

Duffy, S.M., Wheeler, S.J., Bennel, J.D. (1994). Shear modulus of kaolin containing methane bubbles.
Journal of geotechnical engineering 120, 781-796.

Duffy, S.M. (1990). Small strain behaviour of soil containing gas bubbles. PhD thesis, University of
Sheffield.

Esrig, M. I., Kirby, R. C. (1977). Implications of gas content for predicting the stability of submarine
slopes. Marine Geotechnology. 2, 81-100.

Floodgate G., Judd A. G. (1992). The origins of shallow gas. Continental Shelf Research, no. 12, pp.
1145-1156.

Fredlund D.G., Morgenstern N.R., (1978). Stress state variables for unsaturated soils. ASCE
103(GT5), 447-466

Fredlund, D.G., Rahardjo, H. (1993). An overview of unsaturated soil behavior. Edited by S.L. Houston
and W.K. Wray. Unsaturated soils. Geotechnical Special Publication (39), 1-31.

Gens, A., E. E. Alonso, (1992). A Framework for the behaviour of unsaturated expansive clays.
Canadian Geotechnical Journal, 29, 1013-1032.

Grozic, J.L.H., Nadim, F., Kvalstad, T.J. 2005. On the undrained shear strength of gassy clays.
Computers and Geotechnics. 32, 483-490.

86 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Grozic, J.L.H, Robertson, P.K., Morgenstern, N.R. 1999. The behaviour of loose gassy sand.
Canadian Geotechnical Journal. 36 (3), 482-492.

Hardy, R.M., and Hemstock, R.A. (1963). Shearing strength characteristics of Athabasca oil sands.
K.A. Clark Volume, Research Council of Alberta, Information Series No. 45, p. 109.

Haththotuwa, C.K., Grozic, J.L.H. (2008). On the undrained compressive behaviour of gassy sand. In:
Proceedings of the 61st Canadian Geotechnical Conference, Edmonton, Alberta, pp. 370-374.

Haththotuwa, C.K., Grozic, J.L.H. (2009). Behaviour of loose silty sands. In: Proceedings of the 62nd
Canadian Geotechnical Conference, Halifax, Nova Scotia. Paper ID. 175.

Haththotuwa, C.K., Grozic, J.L.H. (2010). The undrained response of silt sands containing gas
bubbles. In: Proceedings of the 63rd Canadian Geotechnical Conference, Calgary, Alberta.

Hight,D.W. (1999). Effect of sample disturbance on measured soil parameters. A review report to NGI
on their Report No. 521676-7.

Hight, D.W. Hamza, M.M., El Sayed, A.S. (2002). Engineering characterization of the Nile Delta clays.
In Coastal geotechnical engineering in Practice, proceedings of the International Symposium
Yokohama 2000. Nakase, A., Tsuchida, T., eds. Swets & Zeitlinger, Lisse, pp 149-162.

Hight, D.W., Leroueil, S., (2003), Characterisation of soils for engineering purposes. Characterisation
and Engineering Properties of Natural Soils, Tan et al. (eds), Vol. 1, 255 – 362, Balkema.

Hovland M., Judd A. G. (1988). Seabed pockmarks and seepages: impact on geology, biology and the
marine environment, Graham and Trotman ed., 293 pp.

Jain, A. K., Juanes, R. 2009. Preferential Mode of gas invasion in sediments: Grain-scale mechanistic
model of coupled multiphase fluid flow and sediment mechanics. Journal of Geophysical Research
(114), B08101.

Johnson, B.D., Boudreau, B.P., Gardiner, B.S., Maass, R. 2002, Mechanical response of sediments to
bubble growth: Marine Geology (187), 347-363.

Jones G.B., Floodgate G.D., Bennell J.D. (1986). Chemical and microbial aspects of acoustically
turbid sediments: preliminary investigations. Marine Geotechnology, vol. 6, no. 3, pp. 315-332.

Judd A.G., Hovland, M. (1992). The evidence of shallow gas in marine sediments. Continental Shelf
Research. Pergamon Press, 1992, Vol. 12, no. 10, pp. 1081-1095.

Kortekaas S., Peuchen J. (2008). Measured Swabbing Pressures and Implications for Shallow Gas
Blow-out. Offshore Technology Conference.

87 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Leroueil, S. (2006). The isotache approach. Where are we 50 years after its development by Prof.
Šuklje? In Proc. 13th Danube-European Conf. on Geotechnical Engineer- ing, volume 2, pages 55–88,
Ljubljana, Slovenia.

Leroueil, S., Kabbaj, M., Tavenas, F., and Bouchard, R. (1985). Stress-strain-strain rate relation for
compressibility of sensitive natural clays. Géotechnique, 35(2):159– 180.

Leroueil, S., Perret, D., and Locat, J. (1996). Strain rate and structuring effects on the compressibility
of a young clay. ASCE Geotechnical Special Publication, 61:137– 150.

Lunne T., Isa O.M., Tan M. (1996). Shallow gas problem at Duyong B offshore Malaysia. OSEA
96065.

Matyas, E.L., Radhakrishna, H.S. (1968). Volume change characteristics of partially saturated soils,
Géotechnique, Vol.18, pp 432-448

Nageswaran S. (1983). Effects of gas bubbles on the sea-bed behaviour, D.Phil, thesis, Oxford
University, 250 pp.

Peuchen J., Jaeck C., Sills G. (2005). Measurement of pore water pressure in offshore gassy soils.
Leidschendam. Report, Fugro, 2005.

Pietruszczak S, Pande GN. (1996). Constitutive relations for partially saturated soils containing gas
inclusions. Journal of Geotechnical Engineering, ASCE. 122, 50–59.

Pietruszczak; S., Pande, G.N., Oulapour, M. (2003). A hypothesis for mitigation of risk of liquefaction.
Géotechnique 53, 833–8.

Rad, N.S., Lunne, T. (1994). Gas in soil. I: detection and n-profiling. Journal of Geotechnical
Engineering, ASCE 120(4), 697-715.

Rad, N.S., Vianna, A.J.D., Berre, T. (1994). Gas in soils. II: effect of gas on undrained static and cyclic
strength of sand. Journal of Geotechnical Engineering, ASCE, 120(4): 716-737.

Sham W. K. (1989). The undrained shear strength of soils containing large gas bubbles. Ph.D. thesis,
Queen's University of Belfast.

Sills G. C. (1989). Triaxial testing of gassy sand. Oxford University Soil Mechanics Contract Report.

Sills G.C., Gonzalez R. (2001). Consolidation of naturally gassy soft soil. Géotechnique, vol. 51, no. 7,
pp. 629 - 639.

Sills, G.C., Thomas, S.D. (2002). Pore pressure in soil containing gas. Edited by Di Maio C., Hueckel
T., Loret B. Chemo-Mechanical Coupling in Clays; From nano-scale to engineering applications, pp.
211-222.

88 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Sills G. C., Wheeler S. J., Thomas S. D., Gardner T. N. (1991). The behaviour of offshore soils
containing gas bubbles. Geotechnique, no. 41, pp. 227-242.

Sills G.C., Wheeler S.J. (1992). The significance of gas for offshore operations. Continental Shelf
Research. Pergamon Press Ltd, 1992, Vol. 12, no. 10.

Smith F.L., Harvey A.H. (2007). Avoid Common Pitfalls When Using Henry’s law. Chemical
Engineering Progress 103, no. 9, pp. 33-40.

Sobkovicz J.C., Morgenstern N.R. 1984. The undrained equilibrium behaviour of gassy sediment.
Canadian Geotechnical Journal 21(3), 439-448.

Šuklje, L. (1957). The analysis of the consolidation process by the isotache method. In Proc. 4th Int.
Conf. on Soil Mech. and Found. Engng., volume 1, pages 200–206, London.

Sultan N., Garziliglia G. (2010). Elaboration d'une loi de comportement pour les sédiments gazeux -
(Projet CITEPH, Rapport Tâches 4 et 5). Ifremer, 2010. Fiche CITEPH N° 23-24/2008.

Thomas S. D. (1987). The consolidation behaviour of gassy soil. D. Phil. thesis, Oxford University,
264 pp.

Van Kessel T., Van Kesteren W.G.M. (2002). Gas production and transport in artificial sludge
deposits. Waste Management (22), pp. 19-28.

Vanoudheusden, E., Sultan, N., Cochonat, P. (2004). Mechanical behaviour of unsaturated marine
sediments: experimental and theoretical approaches. Marine Geology 213, 323-342.

Verruijt A. (1969). Elastic storage of aquifers. In: DeWiest RJM (ed) Flow through porous media.
Academic Press, New York, pp 331–376.

Vesic A.S. (1972). Expansion of cavities in infinite solid mass. ASCE Journal of the Soil Mechanics
and Foundations Division no. 98, pp. 265-90.

Waite W.F., et al. (2009). Physical properties of hydrate-bearing sediments, Reviews of Geophysics
47 (RG4003), pp. 1-38.

Wheeler S. J. (1986). Stress-strain behaviour of soils containing gas bubbles. D. Phil. thesis, Oxford
University, 257 pp.

Wheeler S. J. (1988a). A conceptual model for soils containing large gas bubbles. Geotechnique, no.
38, pp. 389-397.

Wheeler S. J. (1988b). The undrained shear strength of soils containing large gas bubbles.
Geotechnique, no. 38, pp. 399-413.

89 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Wheeler S. J., Gardner T. (1989). The elastic moduli of soils containing large gas bubbles.
Geotechnique, no. 39, pp. 333-342.

Wheeler S. J., Sham W. K., Thomas S. D. (1990). Gas pressure in unsaturated offshore soils.
Canadian Geotechnical Journal, no. 27, pp. 79-89.

Yuan F., Ben Nell J. D., Davis A. M. (1992). Acoustic and physical characteristics of gassy sediments
in the western Irish Sea. Continental Shelf Research, no. 12, pp. 1121-1134.

90 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

8. LISTE OF FIGURES

Figure 2-1. Photography of gassy soil sample. 5


Figure 2-2. Idealization of gas-water meniscus. 6
Figure 2-3. Idealization of gas-water meniscus. 8
Figure 2-4. Example of acoustic masking on sub-bottom profiles, AUV data (2012) 10
Figure 2-5. Gas eruption reaching 10m above drill pipe: mixed soils, seabed about 6m below MSL (Caspian
Sea). 10
Figure 2-6. Blisters shape differences in function of the surrounding matrix (X-ray tomography): (A) spherical
blisters in white loose carbonate silty sand; (B) oblate spheroid blisters in brown dense silty sand Sub-tidal
sediment of Southern UK (low water depth). 12
Figure 2-7. Gas blisters and fissuring on core samples: Both core samples are from same borehole. very dark
greenish grey clay, with traces of shells and shell fragments, water depth 740m. 13
Figure 2-8. Schematic pore pressure record. 14
Figure 2-9. Low net cone resistance and pore pressures on Seabed CPT records. 15
Figure 2-10. Undrained shear strength of gassy sediment – CPT data versus lab test results. 18
Figure 3-1. Experimental set-up where carbon dioxide is injected as dissolved phase into sediment. Decrease of
confining pressure after dissolved gas circulation generates free gas into sediment pores by exsolution. The
triaxial cell is equipped with GDS bender elements. 20
Figure 3-2. Test A: (a) stress paths; (b) P-wave velocities; (c) S-wave velocities 22
Figure 3-3. Test B: (a) stress paths; (b) P-wave velocities; (c) S-wave velocities 23
Figure 3-4. Test C: (a) stress paths; (b) P-wave velocities; (c) S-wave velocities 24
Figure 3-5. Test D: (a) stress paths; (b) P-wave velocities; (c) S-wave velocities 25
Figure 3-5. Preconsolidation pressure against: (a) minimum observed degree of water saturation during gas
exsolution phase; (b) damage parameter d, which is directly linked to Sr and the porosity 27
Figure 4-1. Schematic representation of the structures of soils containing a) gas bubbles much larger than solid
particles (in clays or silts) and b) gas bubbles much smaller than solid particles (in sands). (Wheeler, 1988a). 29
Figure 4-2. Total, water and gas void ratios plotted against time for step-load consolidation tests. (Thomas
1987). 32
Figure 4-3. Total void ratio plotted against operative stress for consolidation tests with load increasing with time
(Thomas 1987). 33
Figure 4-4. Experimental results and model prediction using: a) compressible fluid model and b) double
compressibility model (Thomas, 1987). 34
Figure 4-5. Pore pressure response due to isotropic step increases in cell pressure. (Haththotuwa and Grozic,
2008) 35
Figure 4-6. Experimental results of undrained shear modulus (a) and undrained bulk modulus (b) from triaxial
tests with uw0 = 0 and values of σ3- uw0 varying from 100 to 400 kPa. (Wheeler and Gardner, 1989). 36
Figure 4-7. a. Variation of normalised undrained Young's modulus with gas content from triaxial tests, at an
axial strain ε1 of 0.25%. b. Variation of undrained Poisson's ratio with gas content from triaxial tests, at an axial
strain ε1 of 0.25% (Modified from Wheeler et al., 1991). 37
Figure 4-8. a. Comparison of undrained shear moduli from triaxial tests with theoretical predictions (Wheeler
and Gardner, 1989). b. Comparison of undrained shear moduli from resonant column tests with theoretical
predictions (Duffy et al., 1994). 38
Figure 4-9. Schematic view of the development of plastic zones around oblate spheroidal gas bubbles with
increasing operative stress (σ-uw) during consolidation (After Thomas, 1987). a. Initial stage where there is
sufficient load to produce a small plastic zone around each bubble, without any settlement of the soil matrix. b.
Critical loading stage when the plastic zones are just connecting, still without settlement. c. Connection of plastic
zones producing a mechanism which then deform plastically to allow load to be transferred to the gas resulting
in an overall settlement of the soil. 38
Figure 4-10. Undrained shear strength (a); failure strain (b); degree of saturation at failure (c). 41

91 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-11. Stress ratio at failure against degree of saturation at start of shearing for different combination of
operative stress and back pressure (Wheeler, 1986). 42
Figure 4-12. Experimental values of undrained shear strength (Wheeler 1986, 1988a). 42
Figure 4-13. Undrained triaxial tests results on Ottawa sand with initial void ratio of 0.76 and initial degree of
saturation varying between 0.87 and 1 (Pietruszczak et al., 2003). a. Effective stress paths. b. stress- strain
curves. c. Pore pressure-strain curves. 43
Figure 4-14. Undrained triaxial tests results on loose samples of gassy Ottawa sand with initial void ratio (e 0)
varying between 0.852 and 0.878 and initial degree of saturation (S w0) varying between 0.816 and 0.912 (Grozic
et al., 1999). a. Stress-strain curves. b. Effective stress paths. c. Degree of saturation-strain curves. 44
Figure 4-15. Results of triaxial tests on dense Baskarp sand specimens (Rad et al., 1994). a. Stress-strain
curves. b. pore pressure-strain curves. c. volumetric against axial strains curves. 47
Figure 4-16. Results of undrained triaxial compression test on gassy Ottawa sand and gassy mixtures of
Ottawa sand and Penticton silt Haththotuwa and Grozic (2010). a. Effective stress paths. b. Stress-strain
curves. 48
Figure 4-17. Void ratio versus mean effective stress plot for gassy sand, gassy silt sand and fully saturated silt
sand Haththotuwa and Grozic (2010). 48
Figure 4-18. Drained triaxial tests on a. fully saturated silty clay and silty clay containing methane (Wheeler et
al., 1991); and b. fully saturated silty clay and silty clay containing C0 2 (Vanhoudeusden et al., 2004). 49
Figure 4-19. a. Scanning electron microscope (SEM) micrograph showing a clay fabric around a oval-shape
depression created by a large gas bubble in a sampled maintained under in situ pressures. b. SEM micrograph
showing well-oriented clay fabric from the wall of the gas bubble depression in a. (Chiou et al., 1991). 49
Figure 4-20. SEM micrograph showing the clay fabric of a depressurised sample (Chiou et al., 1991). Somewhat
oriented clay arrangement is shown in the upper portion of the micrograph whereas randomly arranged clay
flakes associated with channel-like features meandering through the central portion of the micrograph. 50
Figure 4-21. Effects of gas exsolution on oedometer tests on samples of Nile Delta Clay. (Hight et al., 2002). 51
Figure 4-22. Results of oedometer tests conducted on gassy (pressurised), degassed (depressurised) and
remoulded clayey sediments of the Mississippi Delta (Johns et al., 1985). a. Comparative plots of oedometer
results in a e-log(σ-uw) graph. b. Values of the coefficient of consolidation, cv, calculated with the Log time and
Square root time methods versus the vertical operative stress. 52
Figure 4-23. Effective preconsolidation pressure (a.) and compression indices (b.) versus depth for samples
collected and tested at in situ pressure (pressurised) and for degassed samples (depressurised). (Johns et al.,
1985). 53
Figure 4-24. Preconsolidation pressure versus degree of gas saturation for Lierstranda and Bothkennar CRS
tests (Lunne et al., 2001). 53
Figure 4-25. Results of CAUC triaxial tests on Lierstranda clay with different degrees of gas saturation (Lunne
et al., 2001). a. Stress-strain curves. b. Effective stress paths. 54
Figure 4-26. Results of CAUC triaxial tests on Bothkennar clay with different degrees of gas saturation (Lunne
et al., 2001). a. Stress-strain curves. b. Effective stress paths. 54
Figure 4-27. Results of CIU triaxial tests on Nile Delta clay (Hight et al., 2002). The red segment highlights the
discontinuity in the effective stress path of samples damaged by gas exsolution. 55
Figure 4-28. Observed undrained pore pressure response (Sobkowicz and Morgenstern, 1984). 56
Figure 4-29. Effective stress path for q-ramping and q-constant unloading of Ottawa sand specimens saturated
with CO2 saturated water (Amaratunga and Grozic, 2009). 56
Figure 4-30. Comparison of collapse lines derived from unloading and shearing of Ottawa sand specimens.
(Amaratunga and Grozic, 2009) 57
Figure 4-31. Comparison of measured and predicted pore pressures for an undrained unloading test on dense
triaxial samples of Ottawa sand (Sobkowicz and Morgenstern, 1984). 60
Figure 4-32. a) Theoretical stress path in q-p’-pc space and b) theoretical void ratio changes versus effective
stress. 60
Figure 4-33. Comparison of predictions of a Cam Clay model modified by Grozic et al. (2005) with experimental
tests results of Wheeler (1988a). 63
Figure 4-34. a. Schematic representation of the structure of a soil containing large bubbles. b. Idealised model
of a soil containing large bubbles (Wheeler 1988b). 63

92 / 94
FUGRO – CITEPH GASSY SOILS – FINAL REPORT

Figure 4-35. Theoretical lower bound curves for undrained strength for different soil types and overconsolidation
ratios (Wheeler et al., 1991). 65
Figure 4-36. Experimental and theoretical values of undrained shear strength. a. CH 4 charged Combwich mud
(Wheeler 1988a). b. He charged Kaolin (Sham 1989). 66
Figure 5-1. Effect of recent stress history on stiffness (Stallebrass and Taylor, 1997) 68
Figure 5-2. Schematic diagrams of MCC and 3-SKH yield surfaces. 69
Figure 5-3. Sensitivity framework (Cotecchia and Chandler, 2000). 70
Figure 5-4. Schematic diagram of S3-SKH model (Baudet and Stallebrass, 2004). 70
Figure 5-5. Definition of S3-SKH model parameters describing structure: initial and ultimate sensitivities, rate of
degradation. 71
Figure 5-6. Ratio of influence of plastic shear to volumetric strain in damage strains used in different models for
structured soils. 71
Figure 5-7. Ratio of influence of plastic shear to volumetric strain in damage strains used in different 73
Figure 5-8. Evolution of strength sensitivity as a function of total strain and effects of strain rate changes on
London clay (experimental data from Sorensen et al. (2010)). 77
Figure 5-9. Evolution of strength sensitivity as a function of total strain and effects of strain rate changes on
Grand Baleine clay (Québec) (experimental data from Lefebvre and LeBoeuf (1987)). 77
Figure 5-10. Evolution of strength sensitivity as a function of total strain and effects of strain rate changes on
GoG sediments (experimental data from Sendir-Torisu 78
Figure 5-11. Coupled effects of structure and strain rate effects. 79
Figure 5-12. RASTRA formulation: influence of strain rate on void ratio–effective stress response. 80
Figure 5-13. Illustration of structure effects and its degradation: influence of initial degree of structure on void
ratio–effective stress response. 80
Figure 5-14. Water retention curve used to simulate gas effects on the mechanical behaviour of a soil. 81
Figure 5-15. Undrained unloading of a gassy clay: evolution of the degree of saturation. 81
Figure 5-16. Mechanical isotropic compression after undrained unloading of a gassy soil. 82

93 / 94

Вам также может понравиться