study of marine gas hydrate on the northern cascadia ...€¦ · study of marine gas hydrate on the...

201
Study of Marine Gas Hydrate on the Northern Cascadia Margin: Constraints from Logging and Seismic Interpretation Xuan Wang Department of Earth and Planetary Sciences McGill University, Montreal Submitted: April 09, 2009 A thesis submitted to McGill University in partial fulfillment of the requirements of the degree of Master of Science © Xuan Wang 2009

Upload: hoangtruc

Post on 14-Sep-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

Study of Marine Gas Hydrate on the Northern

Cascadia Margin: Constraints from Logging and

Seismic Interpretation

Xuan Wang

Department of Earth and Planetary Sciences

McGill University, Montreal

Submitted:

April 09, 2009

A thesis submitted to McGill University in partial fulfillment of the

requirements of the degree of Master of Science

© Xuan Wang 2009

i

ABSTRACT

This thesis presents estimates of gas hydrate distribution and saturation

utilizing data from the four IODP Expedition 311 drilling sites, located at the

northern Cascadia Margin. The objectives are to constrain geologic models of

hydrate formation by determining mechanisms controlling magnitude and

distribution of hydrate occurrences, finding pathways enabling fluid

migration, and examining the effect of hydrate on physical properties of the

sediment. Well log and core data are used to calculate hydrate concentrations,

complemented by pore-fluid geochemical data. Correlation of seismic and

logging data is achieved through synthetic seismograms. Lithology and

faulting appears to control hydrate occurrences, which contradicts established

hydrate formation models. Individual sedimentary layers (e.g. turbidites) and

abundant faults act as migration pathways for fluids and gas explaining the

high hydrate concentrations at shallow depths of less than 100 meters below

seafloor, instead of the previously predicted enrichment near the base of the

hydrate stability zone.

ii

ABSTRAIT

Cette thèse présente des estimations de la distribution et de la

saturation des hydrates de gaz, en utilisant des données provenant des quatres

sites des forages océaniques effectués durant l’expédition 311 de l’IODP,

situés au nord de la marge de Cascadia. Les objectifs de cette étude sont

d’établir des modèles géologiques d’hydrates de gaz en déterminant les

mécanismes contrôlant leurs ampleurs et leurs distributions, les voies

permettant la migration de fluides, en plus d’examiner l’effet des hydrates

sur les propriétés physiques de sédiment. Les concentrations d’hydrates sont

calculées à partir des données provenant de techniques de diagraphie et de

carottage, complétées par des données géochimiques. La calibration de

données sismiques est faite à l’aide d’informations issues des diagraphies de

puits, générant des sismogrammes synthétiques. Les failles et la lithologie

contrôlent l’occurrence des hydrates, contredisant certains modèles

géologiques déjà établies. Les couches individuelles de sédiments (e.g.

turbidites) et l’abondance de failles deviennent des voies de migration pour

l’accumulation de fluides et de gaz. Ceci expliquerait les hautes

concentrations d’hydrates présent à 100 mètres en-dessous du fond de la mer,

et non comme un enrichissement, prédis précédemment près de la base de la

zone de stabilité des hydrates.

iii

ACKNOWLEDGEMENTS

First of all, I wish to give my earnest thanks to my supervisor Michael

Riedel who has provided me with a most inspiring research environment and

has guided me with the utmost patience by his rich knowledge and

experiences, invaluable suggestions and firm support. Thanks for his valuable

suggestions and all the effort he put into my thesis, I cannot have this

achievement without his help.

I would also like to thank several individuals for assistance and warmly

encouragement during my thesis. Many thanks for my boyfriend Zhongwei

Wang who provided invaluable suggestions on programming and data editing.

Thanks also to Jonathan Menivier, who helped me to translate the abstract into

French. Thanks for Peter Neelands who provided high resolution bathymetry

data for this thesis.

I would like to express my gratitude to all the faculty members in the

Department of Earth & Planetary Sciences, who have helped me firmly

through various hard times.

Thanks to all of my friends, thank you for making my life in Montreal

colorful and memorable.

Finally, thanks for my parents, Qing Wang and Yuerong Sue, I owe all

my success to you.

iv

LIST of ABBREVIATIONS

AI Acoustic Impedance

APC Advanced Piston Corer APCT Piston Corer Temperature AVO Amplitude-Versus-Offset

BGHSZ Base of Gas Hydrate Stability Zone BSR Bottom Simulating Reflector CDPs Common Depth Points

CORKs Instrumented Borehole Seals DIT Dual Induction Tool

DSDP Deep Sea Drilling Project DSI Dipole Sonic Imager

DTAGS Deep-Tow Acoustics Geophysics System DVTP Davis-Villinger Temperature Probe

DVTPP Davis-Villinger Temperature-Pressure Probe GHOZ Gas Hydrate Occurrence Zone HNGS Hostile Environment Gamma Ray Sonde IODP Integrated Ocean Drilling Program LWD Logging while Drilling MAD Moisture and Density mbsl meters below sea floor MCS Multi-Channel Seismic MWD Measurement while drilling

NEPTUNE Northeast Pacific Time-Series Undersea Networked Experiment OBS Ocean Bottom Seismographs ODP Ocean Drilling Program RC Refelection Coefficient

ROPOS Remotely Operated Platform for Ocean Sciences SGT Scintillation Gamma Ray Tool TAP Temperature/ Acceleration/ Pressure TWT Two-Way-Traveltime VSP Vertical Seismic Profile XCB Extended Core Barrel

v

TABLE of CONTENTS

CHAPTER 1 INTRODUCTION.............................................................................................1 1.1 OVERVIEW AND OBJECTIVES OF THE STUDY.....................................................................1 1.2 INTRODUCTION TO STUDY AREA .......................................................................................2 1.3 GAS HYDRATE DEFINITION, OCCURRENCE AND DISTRIBUTION .........................................3

1.3.1 Significance of gas hydrate: Potential future fuel resource.....................................4 1.3.2 Significance of gas hydrate: Implication in geohazard issues .................................5 1.3.3 Significance of gas hydrate: A source of green-house gas?.....................................6

1.4 POSSIBLE GAS HYDRATE FORMATION SCENARIOS .............................................................6 CHAPTER 2 OVERVIEW OF PREVIOUS GAS HYDRATE STUDIES OFFSHORE VANCOUVER ISLAND..........................................................................................................9

2.1 SUMMARY OF PREVIOUS SEISMIC WORK AND GAS HYDRATE CONCENTRATION ESTIMATES...............................................................................................................................................9 2.2 SUMMARY OF ODP LEG 146 WORK AND ACHIEVEMENTS ...............................................14 2.3 SUMMARY OF IODP EXPEDITION 311.............................................................................15

2.3.1 Introduction ...........................................................................................................15 2.3.2 Site descriptions.....................................................................................................16

2.4 LOG DATA INTERPRETATION FOR GAS HYDRATE ASSESSMENT........................................17 CHAPTER 3 GAS HYDRATE SATURATION FROM WELL LOGGING CONSTRAINTS .....................................................................................................................19

3.1. GAS HYDRATE OCCURRENCES IN NATURE AND LOGGING APPROACHES.........................19 3.2 ELECTRICAL RESISTIVITY MEASUREMENTS AND ARCHIE’S LAW.....................................20

3.2.1 Archie’s law ...........................................................................................................20 3.2.2 Determing the empirical parameters .....................................................................22 3.2.3 Bulk Density-porosity calculation..........................................................................24 3.2.4 Log resistivity.........................................................................................................24

3.3 ARCHIE ANALYSES ESTIMATES FOR THE IODP EXPEDITION 311 TRANSECT DRILL SITES 25 3.3.1 Site U1325..............................................................................................................25 3.3.2 Site U 1326.............................................................................................................28 3.3.3 Site U1327..............................................................................................................29 3.3.4 Site U1329..............................................................................................................30

3.4 GAS HYDRATE AMOUNT ESTIMATED FROM ACOUSTIC TRANSIT TIME LOGS.....................31 3.4.1 Introduction ...........................................................................................................31 3.4.2 Principle of using acoustic logs to estimate gas hydrate saturation......................35 3.4.3 Methodologies........................................................................................................36 3.4.4 Determining the Empirical weighting factor W .....................................................38 3.4.5 Determining the Gas hydrate saturation................................................................39

3.5 VELOCITY ANALYSES ESTIMATES FOR THE IODP EXPEDITION 311 TRANSECT DRILL SITES.............................................................................................................................................40

3.5.1 U1325 ....................................................................................................................40 3.5.2 U1326 ....................................................................................................................41 3.5.3 U1327 ....................................................................................................................41 3.5.4 U1329 ....................................................................................................................42

CHAPTER 4 REGIONAL SEISMIC ANALYSES AND SYNTHETIC SEISMOGRAMS..................................................................................................................................................43

4.1 SITE U1325.....................................................................................................................44 4.1.1 General Seismic description ..................................................................................44 4.1.2 Lithostratigraphy at Site U1325 ............................................................................48 4.1.3 Synthetic Seismogram generation and Log-to-seismic correlation........................49

4.2 SITE U1326.....................................................................................................................52 4.2.1 General description ...............................................................................................52 4.2.2 Lithostratigraphy at U1326 ...................................................................................55 4.2.3 Synthetic Seismogram generation and Log-to-seismic correlation........................56

4.3 SITE U1327....................................................................................................................57 4.3.1 General Description...............................................................................................57

vi

4.3.2 Lithostratigraphy at U1327 ...................................................................................60 4.3.3 Synthetic seismogram generation and Log-to-seismic correlation ........................61

4.4 SITE U1329.....................................................................................................................64 4.4.1 General description ...............................................................................................64 4.4.2 Lithostratigraphy at U1329 ...................................................................................65 4.4.3 Synthetic seismogram generation and Log-to-seismic correlation ........................66

CHAPTER 5 SUMMARY, UNCERTAINTIES, AND CONCLUSIONS..........................68 5.1 COMPARISON OF RESULTS: RESISTIVITY, VELOCITY AND SYNTHETIC SEISMOGRAM......68

5.1.1 Gas Hydrate Saturation Estimates.........................................................................68 5.1.2 Comparison with previous interpretation ..............................................................69

5.2 UNCERTAINTIES..............................................................................................................70 5.3 CONCLUSION ..................................................................................................................72

REFERENCES:......................................................................................................................74 FIGURES ................................................................................................................................85

TABLES ................................................................................................................................183 APPENDIX Ι : PROGRAMMING OF GAS HYDRATE SATURATION ESTIMATES FROM VELOCITY DATA .................................................................................................194

1

Chapter 1 Introduction

1.1 Overview and Objectives of the study

In this thesis, the main objectives are to characterize marine gas hydrate

occurrences and distribution along a margin-wide transect of drilling sites

located on the Northern Cascadia Margin, offshore Vancouver Island, Canada

by utilizing downhole well logging and seismic data, including Multi-channel

seismic (MCS), Single-Channel seismic (SCS) and Vertical Seismic Profile

(VSP) data.

The drilling data used in this thesis mainly come from Integrated Ocean

Drilling Program (IODP) Expedition 311 (Riedel et al., 2006a), which was

carried out during September – October 2005, and an additional cruise carried

out in August 2008. A transect of four deep drilling sites (U1325, U1326,

U1327, and U1329), which represent different stages in the evolution of gas

hydrate across the entire margin, was chosen during IODP Expedition 311 to

study the occurrences and formation of gas hydrate in accretionary complexes.

IODP Expedition 311 provides new data to the many previous gas hydrate

studies conducted on the Cascadia margin especially Ocean Drilling Program

(ODP) Legs 146 and 204, by extending the aperture of the transect sampled

and by introducing new tools to systematically quantify the gas hydrate

content of the sediments. It was the first part of a multi-expedition study of the

gas hydrate formation system on the northern Cascadia accretionary complex.

Further borehole monitoring experiments tied to the NEPTUNE (Northeast

Pacific Time-Series Undersea Networked Experiment) Canada cable

observatory, would complement these data. Geophysical well log data,

including Logging While Drilling (LWD), measurement while drilling

(MWD), and additional wire-line (WL) logs were carried out along each site

of the transect.

In this thesis, the LWD electrical resistivity and velocity data, wire-line

and MCS velocity data are analyzed combined with other log and core sample

data to calculate gas hydrate concentrations. Correlation of seismic data and

2

logging data is then achieved by generating synthetic seismograms to define

the detailed gas hydrate occurrence at each of the four transect drilling sites.

Results from the geophysical analyses are complemented by geochemical pore

fluid data to further constrain gas hydrate formation characteristics. A more

recent cruise was carried out at the Northern Cascadia Margin during August,

2008. Further seismic and core data were collected and analyzed with a range

of geophysical, geochemical and sedimentological techniques by the scientific

party of 2008-007-PGC Cruise to determine the role of natural gas hydrate in

the mechanics of slope collapse and the mechanism of natural gas movement

around the cold vent area. The data from this new expedition yield a better

understanding of the geologic controls, the evolution, and provide new insight

into the role of gas hydrate in slope stability.

A two-dimensional (2D) MCS line, acquired in 1989 prior to ODP Leg

146 connects all of the four drilling sites, providing a complementary measure

of seismic velocity at each drilling site and adjacent areas. The VSP data from

Hole U1327E are utilized to estimate the depth of the bottom simulating

reflector (BSR), which is usually considered as the interface between gas

hydrate (above BSR) and free gas (below the BSR). Chlorinity data from core

samples provided an independent estimate of gas hydrate concentration where

samples of fresher chlorinity values than the assumed background trend

indicate the presence of dissociated gas hydrate in the recovered core.

1.2 Introduction to study area

The Northern Cascadia margin is part of the Cascadia subduction zone,

which is a convergent plate boundary that stretches from northern Vancouver

Island to northern California. The deformation front separates the Juan de

Fuca, Explorer, and the Northern American Plate. The Oceanic crust subducts

beneath the continent at a rate of ~40mm/yr (Hyndman et al. 2001). The study

area of IODP Expedition 311 is located on the continental slope off Vancouver

Island. Four sites transect southwest-northeast across the margin (Fig. 1.1).

The young Juan de Fuca plate (6-8Ma) is continuously subducted beneath the

North American plate with the consequence that the sediments/rocks all along

the edges of the incoming oceanic plate are compressed (squeezed) and

3

uplifted. The incoming pre-Pleistocene hemipelagic sediments, which are

overlain by ~2-3km thick rapidly deposited Pleistocene turbidite sediments,

are scraped off the oceanic plate during the subduction at the deformation

front, and are accreted so that a large clastic sedimentary prism has formed

since the Eocene. Since the young hot oceanic plate has a relatively low

density and the incoming sediment section is thick, a very shallow trench with

a depth of only ~2500m has been formed at the deformation front. The first

elongated anticline ridge (where Site U1326 located) of the accretionary

wedge is then ~700 m higher compared to the adjacent basins further inland

from the deformation front. The nearest adjacent basin relative to the first

anticlinal ridge developed to the east and is covered by thick (~350m at least),

course-grained sand layers within clay-rich interlayers. The characters of the

sediment within this slope basin indicate that the sediments were transported

parallel to the uplifted bounding ridges (Riedel et al, 2006; Westbrook et al.,

1994).

1.3 Gas Hydrate definition, occurrence and

distribution

Gas hydrates are ice-like solids, which combine water molecules with one

or more hydrocarbon or-non-hydrocarbon gas molecules. The gas molecules

are trapped inside “cages” of hydrogen bonded water. Gas hydrate can only

form under suitable high pressure and low temperature conditions, which

defines the gas hydrate stability zone (Fig.1.3). Thus, natural gas hydrate can

only be found in permafrost regions or within sediments of the deep ocean

(Fig 1.4).

Due to its distinct geophysical and geochemical characteristics, the

presence of gas hydrate within sediments can be identified with well-logging,

seismic (and other remote-sensing geophysical techniques such as controlled-

source electromagnetic surveying) and core-geochemistry data. For example,

gas hydrates can substantially increase sediment electrical resistivity and

seismic velocity; the occurrence of BSRs in seismic profiles, which is usually

considered as the base of gas hydrate stability zone, confine the depth range

4

within which gas hydrates could appear (Hyndman, 1992). BSRs are observed

in reflection seismic data from continental margin worldwide, especially in

subduction zone accretionary prisms. In addition, gas hydrate can also be

indicated from local chlorinity data of recovered core samples. When gas

hydrate dissociates in sediments, it releases the gas and fresh water, with the

fresh water diluting the in situ pore fluid salinity (and chlorinity), thus

resulting in the fresher chlorinity values.

1.3.1 Significance of gas hydrate: Potential future fuel

resource

A widely quoted estimate of global methane gas in gas hydrate published

by the U.S. Geological Survey (Kvenvolden, 1993) predicts that there is twice

as much organic carbon in gas hydrate than in all recoverable and

unrecoverable conventional fossil fuel resources, including natural gas, coal

and oil, combined. Much of this large reservoir of carbon has been thought to

be located on continental slopes in close proximity to major energy-consuming

nations (Fig.1.4) However, the amount of hydrate-bound gas has been

repeatedly quantitatively estimated over the years (Fig. 1.5), though large

uncertainties remain (Milkov, 2004; Klauda et al., 2005).

Arctic gas hydrate reservoirs are potentially the first economically

exploitable reservoirs for the generation of natural gas from hydrate. Some

important arctic hydrate accumulations (Collett et al., 2008) have high

porosity and gas hydrate concentrations, within high intrinsic permeability

sand layers. Such parameters can be considered as the perfect condition for gas

hydrate production.

Compared to permafrost gas hydrate production, marine gas hydrates

have not been considered to represent a recoverable reservoir, and the

complexity of the production techniques is even more enhanced in the marine

environment because of the unconsolidated nature of the host sediments.

Originally estimation of marine gas hydrate reservoir is ~10,000 times larger

than the global conventional gas endowment. However, recent estimation is

5

smaller, although large uncertainties exist (Milkov A.V., 2004). Fig.1.5

compares historic and current estimates of total hydrate bound gas to proved

reserves of conventional gas.

1.3.2 Significance of gas hydrate: Implication in

geohazard issues

Due to the properties of gas hydrate, it can cause natural geohazards.

The decomposition of gas hydrate is speculated to generate a weakness in

continental margin sediments that may help explain some of the observed

patterns of continental margin slope instability. The processes of gas hydrate

formation expel the in situ pore water and take place of the porosity inside the

sediment. Both the growth and decomposition processes inside the sediments

may deteriorate the structure and strength of the sediments. If gas hydrates is

present over a large surface area, the weakened sediments may form potential

planes of weakness, and the potential for sediment failure and slope collapse is

increased. While the relationship between slope failures and gas hydrate

decomposition has not yet been proven, a number of empirical observations

support their potential connection (Paull, 2003). If correct, gas hydrate in near

seafloor sediments may play a significant role as a geohazard, especially in

earthquake-prone areas such as subduction zones.

In addition, new concerns about gas hydrate production emerge. For

example, ocean-based oil-drilling operations sometimes intersect methane

hydrate reservoirs above the oil deposits. As a drill penetrate through the

hydrate layers, the process can cause the hydrate to dissociate, the free gas

may explode, causing the drilling crew to lose control of the well. Another

concern is that unstable hydrate layers could give way beneath oil platforms or,

on a larger scale, even cause tsunamis. When gas hydrate layers are detected

above the reservoir of interest, the production process may require expensive

deviated wells to circumvent the gas hydrate occurrence.

6

1.3.3 Significance of gas hydrate: A source of green‐

house gas?

Although highly disputed, marine gas hydrate is a possible source for

green-house gases, not only because of the globally enormous inventory of gas

hydrate deposits, but also because the stability of gas hydrate deposits can

easily be disturbed by climate changes. Furthermore, dissociated gas hydrate

releases the free gas, which can move through the seabed and reach into the

ocean and/ or atmosphere. Fig. 1.6 shows the snapshot of a gas plume on the

mid-slope (near Site U1327) on Cascadia margin which was recorded by 18Hz

echo-sounders during the 2008-007-PGC Cruise. Several active gas plumes

have been observed on during and prior to the 2008 cruise. Although current

theories of how natural gas can penetrate the gas hydrate stability zone and

exit at the seabed through deep-water gas vents are un-tested, the plumes are

strong evidences to prove that the transportation of methane from deeper

geologic reservoirs to the oceanic and/or atmospheric systems can be quite

effective.

Gas-hydrate-bound methane contained in the sediments may also be

released by large-scale slumping adding methane to the atmosphere. The

speculated mechanism is to transport methane directly into the atmosphere

while bypassing processes, such as methane oxidation and dissolving in the

ocean. During the recent 2008-007-PGC Cruise, large slumps with significant

volumes of solid gas hydrates underneath, were investigated. Additionally,

sites with active release of free gas (indicated as gas plumes on acoustic

profiler data) were studied.

1.4 Possible gas hydrate formation scenarios

At the Northern Cascadia margin, the accretionary sedimentary prism has

been considered the most common host environment for marine gas hydrate

accumulation. In the past few decades, various observations, theories and

models have been brought forward to explain the origin and mechanism of the

free gas and the formation, distribution and stability of gas hydrate, and the

7

impact of the gas hydrate to surrounding host sediment (see Chapter 2 for a

review of the most important studies and findings). Gas hydrate forms if the

concentration of methane exceeds the critical concentration close to the local

solubility threshold. Hydrate will dissolve when the methane saturation falls

below the critical concentration. Gas hydrate formation is also controlled by

the characteristic of the host sediment (e.g. porosity, permeability) the rate of

sedimentation at the local area, the quantity and quality of the organic material,

and the vigor of biological productivity (Matthew et al. 2001). Other factors

like the concentration of gases other than methane, the salinity of the pore

water and the composition of the host sediment also play an important role in

gas hydrate stability zone (e.g. Clennell et al., 1999).

The BSR is believed to correspond to the deepest level at which methane

hydrate is stable; that means below the BSRs, methane may be present as

dissolved or free gas, but not as hydrate. The appearance of BSRs in seismic

sections may largely depend on whether there is sufficient upward pore fluid

expulsion through a thick clastic sedimentary sequence on accretionary

sedimentary prisms (Hyndman and Davis, 1992). The (mainly) biogenic

methane forms far deeper below the base of gas hydrate stability zone, which

is carried upwards by the expulsive fluid flow. When the methane-rich but not

necessarily saturated fluids pass into the gas hydrate stability zone, methane is

removed from the fluid to form the gas hydrate and the hydrate zone begins to

build upward (Fig.1.7). The model explained the origin of the large amounts of

methane and the lack of large quantity of free gas below the stability field.

However, most observations from IODP Expedition 311 contradict this model

in that the highest concentrations of gas hydrate do not necessarily occur

around the BSR.

From combined observations of IODP Expedition 311, gas hydrates are

present at different depths within the sediment and can be found significantly

above the BSR (especially at Site U1326 and U1327). The base of the local

gas hydrate occurrence can be shallower than the base of the

thermodynamically defined stability zone.

According to the model by Xu and Ruppel (1999), the base of gas hydrate

stability may also not coincide with the depth of the BSR. Methane gas

8

hydrate can also occur within this model in a meta-stable form below the base

of hydrate stability zone (and BSR).

An alternate free-gas model has been brought forward by Minshull et al.

(1994). The preferential development of the BSR in structures that would tend

to intercept fluid flow or migrating gas and the presence of free gas beneath

the BSR indicate a mechanism of BSR formation in which free methane gas

bubbles migrate upward into the hydrate stability field or is carried there in

advecting pore water. The gas hydrate zone will form downwards as more gas

migrates upward. However, this model still needs to be verified because gas in

vapor phase is thermodynamically strongly unstable in the hydrate stability

zone (e.g. Buffett, 2000).

The composition of the gas sources, which contribute to gas hydrate

formation, can be classified into two categories: microbial gas, which is

dominated by methane with 13Cδ signatures ranging from -90‰ to -60‰, and

thermogenic gas that contains higher concentrations of 2C + hydrocarbons and

methane with 13Cδ signatures range from -50‰ to -30‰ (Whiticar, 1999). It

is regarded that the majority of gas hydrate near the Earth’s surface is mainly

the result of biogenic conversion of organic matter into methane gas

(Hyndman et al, 1992). However, under certain circumstance, preferential

consumption of 122CO during 2CO reduction has been shown to enrich, or

fractionate, 132CO to the residual 2CO pool to the extent that methane

generated via this process acquires a 13Cδ signature that might be mistaken as

a mix or pure thermogenic source (Claypool et al., 1985; Kvenvolden and

Kastner, 1990; Whiticar et al., 1995).

Paull et al. (1994) postulated a third model of hydrate formation, where it

is produced from in situ local organic carbon. Microbial methane production

takes place below the depth of sulfate reduction. When methane saturates in

the sediment, additional methane production will from either gas hydrate or

free gas, which depends on local temperature and pressure conditions.

However, this model may have some limitations in explaining larger observed

gas hydrate accumulations at the Cascadia Margin, which may require a larger

influx of methane into the stability zone for hydrate formation since the

sediment only contains 0.5% organic carbon.

9

Chapter 2 Overview of Previous Gas

Hydrate Studies Offshore Vancouver

Island

Extensive geophysical studies have been carried out since 1985 to

constraint the occurrence, distribution, and concentration of gas hydrate and

underlying free gas on the Northern Cascadia margin, off Vancouver Island.

This chapter will present an overview of the previous studies and major results,

and linkages to the existing seismic data for more detailed interpretation and

regional extrapolation of gas hydrate occurrences will also be discussed.

2.1 Summary of Previous Seismic work and gas

hydrate concentration estimates

Natural occurrences of gas hydrates have been reported in the

geosciences literature since the early 1970s (e.g. Trofimuk et al., 1973, 1977;

Shipley et al., 1979). In 1985 a BSR was identified in seismic data and the

presence of gas hydrate was inferred on that basis on the margin off

Vancouver Island (Davis and Hyndman, 1989). A wide variety of geophysical

surveys have been carried out to investigate the characters of the gas hydrate

and underlying free gas on Northern Cascadia margin (Fig. 2.1).

The different seismic surveys and data-types include:

(1) Conventional multichannel seismic (MCS) survey lines, which have

been presented and discussed by Spence et al. (1991a, b). The seismic lines

were shot by Digicon Geophysical Corporation in 1989. The survey was

collected using a DSS-240 recording system. The airgun array source was a

tuned array with a total volume of 125L (7820 3in ). Shots were recorded by a

3600m streamer with 144 hydrophones, towed 183m behind the vessel. The

10

MCS survey provided amplitude data as a function of source-receiver offset,

which can be utilized in quantitative AVO analysis (e.g. Yuan et al., 1996,

1999; Chen, 2006);

(2) A pseudo 3-D high resolution MCS survey in the area of ODP Leg

146. The location of this 3-D seismic experiment (shot in 1999) was chosen in

the vicinity of the ODP Sites 889 and 890, which are located at the mid-slope

plateau. The survey location is between two topographic highs and also

covers the blank zones area. The survey was conducted in a NW-SE direction

with a single 40 in3 airgun and a 1.2 km long streamer (details see Riedel,

2001).

(3) Two single-channel seismic (SCS) surveys across known cold vents,

seismically characterized by strong amplitude reduction (thus the name ‘blank

zone”. The first survey carried out in 1999 showed prominent linear features in

seismic amplitude time-slices created from interference-patterns at the edges

of the cold vents (Riedel et al., 2002). The SCS survey carried out in 2000 was

located at the SE corner of the 1999 SC seismic survey, with NE-SW direction,

focusing on Bullseye Vent only (Fig. 2.2).

(4) A very high resolution Deep-Tow Acoustics Geophysics System

(DTAGS) multichannel system that is towed near the seafloor. This survey

was deployed in the area of investigation around the ODP Site in 1997

(Chapman et al., 2002). The system consists of a Helmholtz transducer source

emitting a chirp-like sweep signal of frequencies from 250-650 Hz and a 600m

long hydrophone streamer with two subarrays each containing 24 hydrophone

groups (Gettrust and Ross 1990: Gettrust et al., 1999; Chapman et al. 2002).

(5) Detailed regional SCS and MCS reflection mapping with several

different frequencies (Fink and Spence, 1999; Mi, 1996; Ganguly et al., 2000).

(6) Seismic surveys using Ocean Bottom Seismometers (OBS) for 2-D

and 3-D inversions (e.g. Hobro et al., 2005; Zykov and Chapman, 2004; Lopez,

2008; Dash, 2008).

Beside these seismic methods, additional geophysical techniques were

used: surface heat-probe measurements (Davis et al. 1990; Riedel et al.,

2006b), piston coring with physical property measurements, geochemical

analyses (Novosel, 2002; Riedel et al., 2002; Solem et al., 2002), seafloor

compliance (Willoughby and Edwards, 2000), controlled source electro-

11

magnetics (Yuan and Edwards, 2000, Schwalenberg et al., 2005), and surveys

with the Remotely Operated Platform for Ocean Sciences (ROPOS) of the

Canadian Scientific Submersible Facilities (Riedel et al., 2002) were carried

out in the study area as the complementary for seismic surveys.

The regional occurrence and distribution of BSR, which is considered as

an important indicator of marine gas hydrate appearance, can be mapped from

the various seismic surveys. The BSR across the continental slope of the

Northern Cascadia margin has been defined by data from non multichannel

seismic lines (Hyndman et al., 1994). The BSR can be easily identified almost

over more than half of the region, which is in contrast to the deep sea Cascadia

basin, where no BSR appears. The results can be further constrained and

improved by other data mentioned above. The BSR reflection amplitude

represents a decrease in seismic impendence of 20-30% (Hyndman et al. 2001).

No continuous reflection above the BSR was observed that may indicate the

top of a high-velocity hydrate layer and no continuous reflection below the

BSR was seen that may represent the base of a low-velocity free gas layer. The

surveys and analyses of the seismic data that are described below include

mapping the BSR and estimating the BSR reflection coefficient (Spence et al.,

1991a; 1991b; 1995; Hyndman et al., 1994; Yuan et al., 1996, 1999).

Mapping the BSR is used to be considered a method for estimating the

distribution of gas hydrate over a certain region. Gas hydrate concentration are

semi-quantitatively estimated by assuming that the BSR reflection coefficient

is related to hydrate concentration above the BSR (Fink and Spence, 1999),

though large uncertainties exist due to the reflection coefficient is sensitive to

the velocity increase due to the hydrate concentration.

IODP Expedition 311 was based on about 20 years of geophysical

research and field studies of gas hydrate on the northern Cascadia margin

offshore Vancouver Island. Ocean Drilling Program (ODP) Leg 146 provided

some deep drilling data at two sites (Sites 889 and 890) offshore Vancouver

Island at a mid-slope location within the accretionary prism (Westbrook et al.,

1994). A third site was established in the Cascadia Basin (Site 888) where no

gas hydrate is present, providing important data for comparison-purposes to

observations made at the two Sites 889 and 890 within the general zone of gas

hydrate occurrence.

12

Early analyses of the well-log data pointed towards a large accumulation

of gas hydrate near the base of the gas hydrate stability field (e.g. Hyndman et

al., 1999, Yuan et al., 1999; Hyndman et al., 2001), but was later revised

(Riedel et al., 2005) and then re-visited during the IODP Expedition 311

drilling campaign (Riedel et al., 2006a). Regional geophysical analyses

focused for a long time on the seismic occurrence of the bottom-simulating

reflector (BSR). Analysis of seismic data demonstrated a characteristic

reflector, which, in general, coincides with the predicted base of methane

hydrates stability field and mimics the pattern of the sea-floor. The BSR also

has a reversed polarity compared to the seafloor reflection, which indicates a

strong transition from high impedance above the BSR to the low impedance

below the BSR.

There are a variety of approaches to modeling the BSR reflection

waveform (e.g. Hyndman and Spence, 1992; Fink and Spence, 1999). The

seafloor reflection coefficient has been estimated from the relative amplitude

of the primary seafloor reflection and its multiple (Warner, 1990), and from

the velocity and density of the sediments from piston core data (Mi, 1998).

When calculating from conventional multichannel seismic data, the BSR

reflection coefficients yield ~0.1 compared to typical seafloor reflection

coefficients of 0.18-0.24. (Hyndman and Spence, 1992; Fink and Spence, 1999;

Yuan et al., 1999; Yuan, 1996; Mi, 1998; MacKay et al., 1994).

The other characteristic of the BSR is a frequency dependence of the

reflection strength (Chapman et al., 2002; Riedel et al., 2002). The BSR

reflection coefficient has been observed to decrease markedly with increasing

frequencies of different methods of seismic data collection (Spence et al., 1995;

Zühlsdorff et al., 2001; Gettrust et al., 1999; Chapman et al., 2002). This can

be explained by a BSR that represents a gradational velocity contrast over a

depth interval of 6-10m (Chapman et al., 2002). The thickness of the BSR will

appear as a sharp interface in low-frequency data because it is less thick than

one wavelength for these low frequency waves; and it is thicker or about the

same thickness than the wavelength for higher frequency waves, so that the

BSRs will appear as a transition zone similar to a thin-bed.

Detailed seismic interval velocity studies provide the main quantitative

constraint on gas hydrate and free gas concentrations from remote seismic data;

13

Velocities determined to greater depth where no gas hydrate and free gas

appears provide an estimate of the reference velocity-depth profile for the

homogeneous sediment section, which can be used to estimate the regional gas

hydrate saturation for those sediments within the gas hydrate stability zone

(e.g. Yuan et al., 1996). The interval velocity data mainly were determined

from the conventional multichannel seismic data that were acquired in 1985

and 1989 as pre-site survey for ODP Leg 146. The deep towed DTAGS

multichannel system give similar results but with higher spatial (vertical)

resolution (Chapman et al., 2002).

The other methods that may provide constraint on the concentration of gas

hydrate and mainly free-gas from suitable multichannel seismic data is the

study of BSR amplitude variation with offset (AVO), which determines the

change in reflection amplitude as a function of receiver offset (Hyndman and

Spence, 1992; Yuan et al., 1999; Riedel et al., 2001, Chen, 2006). In today’s

oil and gas industry, AVO analysis has been considered as a method

commonly used for hydrocarbon detection. However, the model from Yuan et

al. (1999) has indicated qualitatively that AVO behavior of the hydrate BSR

may not be as useful for determining the amount of free gas below the BSR as

was previously hoped. This was further investigated by Chen et al., (2006)

who performed a complete non-linear inversion combined with probability

determinations of the complete AVO problem and showed that it is almost

impossible to derive meaningful estimates of the physical properties of

sediments above and below the BSR.

Full-waveform inversions of seismic data from the northern Cascadia

margin have also been carried out to find a velocity-depth profile for further

constraints on gas hydrate saturations. The full-wave form inversion works

such that the synthetic data produced by the model fits amplitude, waveform

and phase of the data sample by sample over all offsets (Singh and Minshull,

1994; Yuan et al., 1996, 1999). However, the determined P-wave velocity

profiles from this type of inversion did not yield any additional constraints that

changed interpretations already performed with all other techniques.

Further seismic studies were carried out using ocean-bottom seismometers

(OBS) on a more regional scale (e.g. Hobro et al., 2005) or for more site-

specific purposes (e.g. Zykov and Chapman, 2004; Lopez, 2008; Dash, 2008).

14

A new cruise was carried out on August, 2008, with a multi-disciplinary

effort to determine the relationship with the natural marine gas hydrate and the

slope collapse at the frontal ridges of the Vancouver Island accretionary wedge

(Lopez, 2008), as well as gas migration through the seabed to the ocean at cold

vents. This expedition has visited 25 core sites distributed over the frontal

ridge, mid-slope, and Barkley Canyon areas of the Vancouver Island

accretionary wedge. Core processing and sampling, physical properties

measurements, geochemistry analysis, seismic activities were carried on the

ship for achieving the main objectives (Haacke et al., 2008).

2.2 Summary of ODP Leg 146 work and

achievements

ODP Leg 146 was directed to investigate the fluid flow and sediment

deformation within the accretionary wedge that forms the Cascadia Margin,

and tries to find the cause of BSRs and their relationship to the occurrence of

gas hydrate and free gas. ODP Leg 146 includes 5 drilling sites. Sites 889

and 890 are located on the mid-slope off Vancouver Island and were drilled

over a strong BSR. Sites 889/890 provided significant guidance for the IODP

Expedition 311. Site 888 was drilled in the deep Cascadia Basin and is near

the proposed location of Site CAS-04B (yet not drilled) and is used as a

reference “non-gas-hydrate” environment.

The other two Sites 891 and 892 are located on the Central Oregon

Margin, close to the deformation front where the sediment structures are

controlled by a series of well-defined folds and thrust faults in the lower

continental slope (MacKay et al., 1992) (Fig. 2.3a and b). These two sites

were drilled to specifically investigate the flow through fault zones. Site 891

examined the frontal thrust fault that connects to the master décollement that

lies close to the igneous basement of the subducting oceanic lithosphere. Site

892 shows that the BSR at this Site forms toward the surface of a

hydrologically active fault. Studies show that this fluid flow controls the

composition, distribution, and concentration of the gas hydrate at this site.

15

Site 889 was drilled through bedded, slope-basin sediment on the mid

slope plateau and also penetrated into the underlying, deformed sediments of

the accretionary wedge. Site 890 was cored to only 50mbsf to sample the

near surface sediments, about 8 km further west of Site 889. The major

objective at these two sites was the investigation of a well-developed BSR at

~225 mbsf and characterization of diffuse fluid flow from the accretionary

wedge. An instrumented borehole seal (CORK-Circulation Obviation

Retrofit Kit) was emplaced in Hole 889C (another CORK was also placed at

Site 892) to provide long-term observation of the thermal, chemical, and

hydrogeological conditions associated with the gas hydrate zone (Davis et al.,

1995; Westbrook et al., 1994). The CORK at Site 889C will be connected to

the NEPTUNE cable observatory in the summer of 2009 to provide

continuous recording of pressure and temperature.

According to the combined observations from laboratory data and core

sampling, and logging data analyses, the BSR occurs at this area at the base

of the hydrate stability, as typically seen on subduction zones (e.g. Shipley et

al., 1979; Kvenvolden and Barnard, 1983). The temperature and pressure

conditions at the BSR agree with the maximum P-T condition at which

hydrate is stable.

2.3 Summary of IODP Expedition 311

2.3.1 Introduction

IODP Expedition 311 successfully established a transect of four drill sites

across the Northern Cascadia margin. At each site several different boreholes

were drilled, with a variety of different measuring methods. Details on the

measurement methodology can be found in the IODP Proceedings Volume for

Expedition 311 (Riedel et al., 2006a).

16

2.3.2 Site descriptions

Site U1325 is located near the southwestern end of the drilling transect

and is situated within a slope basin that developed eastward of the

deformation front (Fig 1.3 and Fig. 1.4). Four individual boreholes were

occupied at Site U1325. The depth of the penetration at each hole can be

found in Table 2.1. Hole U1325A was dedicated to the acquisition of

Logging-while Drilling (LWD) data with a total depth of 350mbsf.

Advanced piston corer (APC) system, extended core barrel (XCB) and

pressure coring were applied in Hole U1325B; high quality temperature

data using the Davis-Villinger Temperature Probe (DVTP) were acquired

to define the geothermal gradient. Two separate logging runs in Hole

U1325C were done. The first run included the Dual Induction Tool (DIT)

and the Hostile Environment Gamma Ray Sonde (HNGS). The second run

included the Dipole Sonic Imager (DSI), the Scintillation Gamma Ray Tool

(SGT), and the Temperature/ Acceleration/ Pressure (TAP) tool. The

second tool string (sonic without the Formation MicroScanner [FMS]) was

deployed and logged successfully.

Site U1326 is located on the first uplifted ridge closet to the

deformation front, where a BSR is present underneath most of the ridge (Fig

1.3). There are three individual boreholes at Site U1326. Hole U1326A was

drilled for LWD logging to a total depth of 300m; Hole U1326B was failed to

establish a mudline, and U1326C resumed spudding on the same location of

U1326B, which serves as a continuous core hole with conventional APC and

XCB coring. Three pressure cores were deployed in this hole. U1326D was

drilled ~30m southwest from U1326C.

Site U1327 is located near two topographic highs on the mid-slope of the

high plateau over a clearly defined BSR. The major scientific problem at this

site is to determine a reliable geochemical reference profile as well as well-

logging data. These data are also of utmost important to calibrate remote

sensing techniques, such as reflection seismic and controlled-source

electromagnetic (CSEM) surveys. Five holes were occupied at Site U1327.

Hole U1327A was dedicated to the LWD measurements. U1327B recovered

17

only one APC core to a depth of 9.5 mbsf. However, Hole U1327C was

continuously cored to 300mbsf, with 10 APC cores, 22 XCB cores, and 3

pressure cores. In addition, four advanced piston corer temperature (APCT)

tool measurements were made, and three additional deployments using the

DVTP and Davis-Villinger Temperature-Pressure Probe were completed.

Logging data including a Vertical Seismic Profile (VSP) were acquired in

Holes U1327D and U1327E.

U1329 is located at the landward end of the transect, and is believed to be

at the eastern-most limit of gas hydrate occurrence along the northern

Cascadia margin. No apparent gas hydrate layers occur at this site according to

well logging and geochemical data. However, a BSR appears at this site, but at

a much shallower depth of 125 mbsf compared to all the other sites. Five

boreholes were occupied at Site U1329. LWD measurements for Hole

U1329A to a total depth of 220mbsf, Hole U1329B was occupied for a ~10m

long APC core, Hole U1329C was continuously cored to 189.5mbsf with 17

APC, 5 XCB and 3 PCS cores, and Hole U1329D was wireline logged with

the triple combo and FMS-sonic tool strings. Hole U1329E was a special tool

hole with 5 APC cores.

2.4 Log data interpretation for gas hydrate

assessment

To quantitatively determine the amount of the gas hydrate from

geophysical, geochemical, and geological data is challenging as all techniques

involved have to rely on the comparison to an assumed gas hydrate free

reference, which is not always easily defined. The gas hydrate saturations at

the new IODP sites have been previously determined through combining pore

water chemistry and in situ downhole log measurements (e.g. Malinverno et

al., 2008; Chen et al., 2008). The pore water salinity methods are based on the

process that chlorinity decreases due to fresh water generated by gas hydrate

dissociation upon core recovery (Ussler and Paull, 2001). Well logging

resistivity methods utilize the empirical Archie’s relationship (Archie, 1942).

Pore water data analyses provide only dispersed data highly dependent on core

18

recovery rates, but the interpretation of the chlorinity data is straightforward if

a continuous background-trend in chlorinity can be determined. Downhole log

analyses provide continuous geophysical data downhole, but reference values

for resistivity and P-wave velocity have to be determined from other (regional)

data. The interpretation of both of these data types have many uncertainties

that mainly come from the inaccurately determined in situ formation

temperatures, baseline chlorinity (salinity), density, and resistivity. A recent

review of ODP Site 889 log data showed that the gas hydrate concentration

reaches as high as 0.3-0.4 or as low as 0.05-0.1, depending on the no gas

hydrate reference salinity used in the interpretation (Riedel et al, 2005).

The average gas hydrate saturations within the GHOZ across the entire

margin are likely between 0.04-0.1 (Malinverno et al., 2008). This value is

potentially underestimated (by probably about 10%) considering that thin sand

layers were not recovered or are below the resolution limit of the various

logging tools employed. This average is much lower than previous estimates

of 0.2-0.3 in the Cascadia accretionary wedge (Hyndman and Spence, 1992;

Hyndman et al., 1999).

Considering the estimates from density porosity to be the most accurate

(Riedel et al., 2006a), gas hydrate saturations averaged over a 10 m window

show distributed gas hydrate occurrence in many intervals (Sh = ~0.09 ± 0.07

at Site U1326 [170–200 mbsf]; Sh = ~0.10 ± 0.07 at Site U1325 [190–230

mbsf]; Sh = ~0.11 ± 0.07 at Site U1327 [140–225 mbsf]), with average

concentrations of 5%–15% of the pore space, over depth intervals of 20–100

m (Chen. et al, 2008).

In comparison, distributed gas hydrate environments around Southern

Hydrate Ridge (Cascadia Margin, offshore Oregon, ODP Leg 204) show low

average gas hydrate saturation (0.02-0.08 (Tréhu et al., 2004).

In this thesis, the various log data from LWD and wire-line tools were

used to estimate gas hydrate saturations. Additionally a complete time-depth

conversion was carried out for the first time using the vertical seismic profile

and MCS velocity data to accurately convert seismic measurement in two way

travel time (TWT) to drilling-depth. Through this approach, a linkage is build

between the seismic data and log data for a more careful interpretation and

regional extrapolation (details will be introduced in Chapter 4).

19

Chapter 3 Gas hydrate saturation from

well logging constraints

3.1. Gas Hydrate occurrences in nature and logging

approaches

There are different types of natural gas hydrate formations, and different

gas hydrate formations cause different responses on well logging data.

Information about the different nature and texture of gas-hydrate occurrences

is needed to assess the response of well-logging devices to the presence of gas

hydrates. Most of the time, the occurrence of gas hydrates can be detected

locally when undergoing drilling and coring, but its large scale distribution is

difficult to assess due to the lack of further well control. Therefore, the

existing and proposed quantitative well-log evaluation techniques assume

uniform distribution of gas hydrates as interstitial “suspending” or “cement”.

Under these conditions, gas-hydrate-bearing reservoirs can be evaluated using

“standard” well-log evaluation techniques developed for a mixed

multicomponent rock matrix, water, gas, ice, and/or gas hydrate systems

(Collett, 2000). There are a series of reservoir models for the occurrence of

gas hydrates in nature; however, the most suitable model of the study area in

this thesis is that designed to represent a clay-rich marine gas hydrate

reservoir. It assumes the clay content of the sediment and associated volume

of bound-water is higher in most marine gas-hydrate reservoir (Collett, 2000).

Combined drilling and seismic observations suggest strong lithologic control

of the gas hydrate occurrence with preferred gas hydrate formation in sand-

rich sedimentary sections. However, the occurrence of various clays in

sediments has numerous significant effects on well-log responses, and the

responses will change due to different clay content of the sediments, such as

neutron spectroscopy or electrical resistivity well logging. Such parameters

are also considered in this study.

20

The technology of well logging (velocity, porosity, resistivity) and how

well the drilling tools are able to capture the nature of hydrate occurrences,

like the limitation from the resolution and disseminated issues is given a full

account in Proc. of the Integrated Ocean Drilling Program, Volume 311

(Riedel et al., 2006a).

In the following section two different methods of calculating gas hydrate

concentrations are reviewed from: (a) Archie’s law by using log-derived

electrical resistivity, porosity (neutron and density-porosity) and core-derived

data, under the condition that in situ pore fluid resistivity (i.e., salinity) is

known, and (b) using three-component modified compressional-wave (Vp)

acoustic equation by calculating log-derived and seismic-derived acoustic data

(sonic logs, VSP velocity data and multichannel seismic interval velocities).

3.2 Electrical resistivity measurements and Archie’s

law

3.2.1 Archie’s law

Archie’s law (Archie, 1942) is a purely empirical relation that can be used

to describe the relationships between gas hydrate saturation and water-

saturated sediment resistivity. The critical problem with this method is to

estimate the in situ pore fluid resistivity, and to determine appropriate values

for the empirical parameters in the equations involved. Archie has established

the relationship that the resistivity of a fully water saturated sediment ( 0R ) is

proportional to the resistivity of the pore fluid ( wR ):

0 wR FR= (3.2.1)

While the proportionality constant F is called the formation factor.

After examining core samples from different formations, Archie (1942) further

established an exponential empirical relationship between F and the porosity

(ϕ ):

21

0 m

w

RFR

ϕ−= = (3.2.2)

The exponent m is called as the formation parameter. Winsauer and

Shearin (1952) modified Archie’s original equation by including a coefficient

a in the relation. The modified equation relates the water content (saturation)

to the in situ electrical conductivity of water-saturated sediments, of varying

inter-granular porosity:

m n

t w wR aR Sφ− −= (3.2.3)

Here, φ denotes the porosity; tR represents the formation electrical

resistivity (from log), wR is the electrical resistivity of formation water, m is

the cementation exponent of the rock (usually in the range 1.8–2.0), and n is

the saturation exponent (usually close to 2).

For most situations, the laboratory core-derived pore fluid resistivity or

salinity can be used. However, these values are strongly affected by the

dilution effect due to the dissociation of gas hydrate upon core recovery.

Fortunately, this problem can be solved by integrating the logging resistivity

data with core salinity and core porosity data, again using Archie’s law to

estimate both the hydrate concentrations and the in situ salinities (Hyndman et

al., 1999).

The cementation exponent m models how much the pore network

increases the resistivity, as the rock itself is assumed to be non-conductive.

Theoretically, if the pore network were to be modeled as a set of parallel

capillary tubes, a cross-section area average of the rock's resistivity would

yield porosity dependence equivalent to a cementation exponent of 1.

However, the tortuosity of the rock increases this value to larger than 1. This

relates the cementation exponent to the permeability of the rock; increasing

permeability decreases the cementation exponent. Usually, for unconsolidated

sands, the exponent m has been observed to be near 1.3, and it is believed to

increase with cementation. Common values for this cementation exponent for

consolidated sandstones are 1.8 < m < 2.0.

22

The formation factor is sometimes modified to a / mφ , where the constant

a means to correct for variation in compaction, pore structure and grain size.

The constant is sometimes denoted turtuosity factor or cementation intercept,

and usually falls between 0.6 and 1. The cementation exponent is usually

assumed not to be dependent on temperature.

The saturation exponent n can be estimated by averaging values for different

lithologies (Pearson et al., 1983). However, modeling showed that the

parameter n is not constant and is dependent on the grain size distribution and

gas hydrate concentration itself (Spangenberg, 2001). Thus, there is no fixed

value for n to use for the entire sediment column under investigation.

However, the most widely accepted approach is to set the value of n equal to

m (e.g. Hyndman et al., 1999), so that equation (3.2.3) becomes:

( ) mt w wR aR Sφ −= (3.2.5)

3.2.2 Determining the empirical parameters

The only reliable source for the numerical values of both exponents is

from experiments on sand plugs from cored wells. The brine conductivity can

be measured directly on produced water samples. However, the dissociation of

gas hydrate will release fresh water, thus largely increase the conductivity

value. If this value is used in determining the empirical parameters in

Archie’s law, the results will be significantly in error. In this thesis the

approach taken was to calculate the in situ formation electrical resistivity

using the equation of state of sea water (Fofonoff, 1985). The equation

calculates the in situ water resistivity as function of salinity (in PPT),

temperature (ºC) and pressure (deci-bar). The brine conductivity and the

cementation exponent can also be inferred from downhole electrical

conductivity measurements across free gas/gas hydrate intervals. For such

intervals (Sw = 1) Archie's law can be written as:

mt wR aR φ−= (3.2.4)

Parameters a and m in equation (3.2.3) are determined from purely

water-saturated sediments. Usually they are determined by a log-log

23

crossplot of formation factor (FF) as a function of porosity, known as

“Pickett plot” (Serra, 1984). A straight-line relationship is expected (or a

power line relationship when measured porosity and conductivity were

plotted). The slope of this linear trend equals the cementation exponent m

and the intercept equals the logarithm of the parameter a.

There are two sources of data to generate the FF-porosity crossplot, (a)

data from discrete core samples, or (b) from LWD downhole logging data.

Core data are only measured at discrete intervals and the generally small

numbers of data points limit the accuracy of the results in a statistical sense.

Furthermore, core data can be largely disturbed from the actual in situ

sediment conditions. A correction for temperature and pore-water salinity can

be made; but the sediment texture is often heavily altered. In most cases, the

formation of the gas hydrate in the sediment may crack the poriferous

sediment and disrupt sediment fabric. Estimates of the parameter a and m

from discrete core samples must therefore be taken with caution.

The LWD geoVISION resistivity-at-the-bit (RAB) tool is considered the

most reliable downhole resistivity measurement. The RAB tool is connected

directly above the drill bit and uses two transmitter coils and several

electrodes to obtain different measurements of resistivity. It has three depths

of investigation; in this thesis, the deepest investigation depth was used, that

the signal can penetrate ~0.4-0.6m deep into the surrounding borehole wall.

LWD downhole log data are less affected by poor borehole conditions (e.g.

borehole enlargements in washout zones); compared to other data set acquired

by LWD logging tools. Neutron logs are measuring the amount of hydrogen

within the pore-space of the rock sequence. When gas hydrate forms inside

the pore space, it will largely affect the neutron log measurement. Due to the

principle of the measuring method, neutron porosity measures the pore space

excluding the gas hydrate and gives larger results when the gas hydrate

formation appears in the sediment. Measurement for depths well below the

BSR may represent the best representation of a no-hydrate/no-gas reference

(Riedel et al., 2005).

It should be noted that estimates of the Archie coefficients may differ for

near surface sediments and deeper sediment sections (Novosel, 2002).

24

Typically the upper 10-20mbsf show considerably higher porosities than

predicted by an upward extrapolation of a deeper trend (e.g. by Athy’s law)

and result in a much smaller value for the parameter m (Riedel et al., 2005). In

most of the cases in this thesis, the first 10-20mbsf is usually excluded from

the calculations.

3.2.3 Bulk Density‐porosity calculation

In conventional analysis, density logs are primarily used to assess in situ

sediment porosities. The porosities can be derived from the standard relation:

m b

m w

ρ ρφρ ρ

−=

− (3.2.6)

Where mρ is the known matrix density (g/ 3cm ), which is 2.70g/ 3cm ,

wρ is the fluid (water) density (g/ 3cm ), which equals to 1.05 g/ 3cm , and bρ

is the log measured formation bulk-density (g/ 3cm ), which is the gas hydrate

density of 0.9g/ 3cm (from Collett, 2000).

3.2.4 Log resistivity

Figure 3.1 shows the resistivity profiles at the four Expedition 311 site

wells. At each site, the seismically and log inferred base of gas hydrate

stability zone or BSR depth is shown. These resistivity logs qualitatively

indicate certain zones of gas hydrate occurrence by intervals of high values

relative to a lower assumed background value. High-porosity unconsolidated

marine sediments in the study area generally have resistivities on the order of

1 Ωm (Riedel et al, 2006a). Slightly elevated values are seen around the depth

of the BSRs only at Site U1326 (255–261 mbsf) and at Site U1329 (120–124

mbsf), probably related to some small amount of gas hydrate. However,

certain zones above the inferred BSR exhibit much higher resistivities and are

therefore interpreted to be gas hydrate bearing, especially at Site U1326 at 73–

94 meters below seafloor (mbsf) and 252–261 mbsf, at Site U1325 in thin

layers between 195 and 240 mbsf, and at Site U1327 at 120–138 mbsf. The

25

high-resistivity zone below 176 mbsf at Site U1329 is interpreted not to be gas

hydrate or free gas related but rather to an unconformity, below which much

older, low-porosity, lithified Miocene (>6.7 Ma) sediments occur.

In certain well log and core porosity measurements, the gas hydrate is

considered to be a part of pore space because the properties of gas hydrate,

which is measured by the tools, is similar to pore water. Available porosity

measurements are from the density and neutron logs and from IODP

shipboard core moisture and density (MAD) analyses after gas hydrate has

dissociated. In this thesis, LWD density values were utilized to calculate the

density-porosity value.

In the following sections, the empirical value a and m were derived from

a “Pickett Plot” using two data source: (a) core-derived density porosity with

FF (Rt/Rw); (b) LWD density porosity with FF (Rt/Rw). According to Chen et

al. (2008), using all density porosity data and formation factor values from gas

hydrate-free zones, a cementation factor m of 1.751 and a of 1.394 is

determined. This result is very close to the earlier estimates by Westbrook et al.

(1994) and Hyndman et al. (1999). The empirical value m is fixed to 1.76 for

this part of the Cascadia Margin for all calculations in this thesis.

3.3 Archie analyses estimates for the IODP

Expedition 311 transect drill sites

3.3.1 Site U1325

Through the equation of state of sea water (Fofonoff, 1985), the in situ

water resistivity values were calculated from temperature and salinity data.

The in situ resistivity is roughly decreasing from 0.34 Ω m at the seafloor to

~0.2 Ω m at the bottom at ~300mbsf. However, the down-hole profile can be

divided into three parts. The first part is the segment of 0-36 mbsf, starting

from 0.34 Ω m and then decreasing with a power trend. The second part is

from 36 mbsf to 171.5mbsf; the values decrease from 0.29 Ω m to 0.24 Ω m

26

almost linearly. In the third segment from 171.5mbsf to 263mbsf resistivity

raises from 0.241 Ω m to 0.243 Ω m at 183mbsf and then decrease again to

0.21 Ω m at 263mbsf. For this part, polynomial fitting is used to get the best

curve fitting result (Fig. 3.2).

To determine both of the empirical parameters a and m, the density

porosity values are plotted against the FF data. A power line is then fit through

the data points to give the best fit value. The results from this method are not

satisfactory as they provide values that are outside the expected range from

previous analyses (Hyndman et al., 1999; Chen et al., 2008). Based on the

LWD data from Hole U1325A, the curve-fitting yields a equals 3.48 and m

equals to 0.43 (Fig 3.3). When calculating the parameters from the core data

alone, a is 3.23 and m is 0.46. Referring to the results from Hyndman et al

(1999), parameters a and m at Site 889/890 are 1.41 and 1.76, respectively. A

very similar result was obtained at the deep basin Site 888 which was

considered as the reference site where no gas hydrate appear at this site

(a=1.39, m=1.76). This result was calculated from discrete core sample

measurements.

Other substantially different values were found by Collett (2000) who

used data from log-derived neutron and core porosities and the electrical

resistivity logs at Site 889 (a=1, m=2.8). Another possible data set for

comparison can be derived from drilling data at the southern Cascadia margin,

ODP Leg 204 (Tréhu et al., 2003). The results from Leg 204 within a very

similar tectonic and sedimentological environment than Expedition 311 are

mainly determined from slope sediments. Drilling and logging during Leg 204

only partially penetrated into accreted sediments, which is in contrast to Site

889 (and U1327), where large gas hydrate concentrations were estimated for

the accreted mélange sediments below 128mbsf and lower concentrations

within the upper slope basin sediments (Riedel et al., 2005).

The values determined at Site U1325 (m=0.57 from log-derived

calculation and m=0.46 from core-derived calculation) are much lower

compared to the results of Collett (2000) and Hyndman et al (1999). It is

speculated that the different empirical parameters may be an effect of quite

different site locations. Site U1325 is located near the southwestern end of the

27

accreted prism and is at the northeast side within a major slope basin. The

differences of sediment structures between the three drilling sites result in

large discrepancies of the porosity and pore connectivity at each site, thus

causing the different empirical values. The sediment-setting of Site U1325 and

Leg 204 has more similarities (in general terms).

As mentioned above, the empirical value m=1.76 from Hyndman et al

(1999) and Chen et al., (2008) is utilized in the following calculation to

determine the empirical value a. When m is fixed, we can get a data set of

value a by substituting the density –porosity and FF value at each depth into

the Archie relationship, the value a can be determined as it is the only

unknown parameters in the equation. Then, the average value of a is obtained

from the whole data set. The parameter a is calculated three times according to

the different type of porosity measures utilized in this thesis.

The result of this constraint in the determination of a for the three

different available porosity data sets is the following:

(a) Using the core porosities only, a = 1.64; (b) using density-porosity

data, a =1.17, and (c) using neutron porosities, a =2.24. It should be noted that

due to enlarged hole diameter, the neutron porosity log data are significantly

degraded, especially in the upper 25mbsf and below 250mbsf. Fig. 3.4 shows

the different gas hydrate saturations values calculated using the three different

a values and the fixed m value.

From Fig. 3.4 (a), the gas hydrate saturation curves from the density-

porosity and core-derived porosity roughly follow the same trend, but the

core-derived values are slightly lower than the density-porosity estimates. The

average saturation estimate given by log derived porosity is ~15% and

estimates from core-derived porosity are ~20%. From Fig. 3.4 b, the average

gas hydrate saturation from m=0.43 and a=3.48 is10-20%. Pore-water

saturation estimates below 225 mbsf, which is the base of the gas hydrate

stability zone, may indicate the presence of free gas, although no independent

confirmation from pressure-core analyses was derived (Riedel et al., 2006a).

28

3.3.2 Site U 1326

From Fig. 3.5 (a), we can see that the core-derived salinity value roughly

follow the baseline of 33-34 ppt, except for two segments between 50-

110mbsf and 190-225mbsf. The salinity increases a little bit in the top-most

shallow part, i.e. from 32 ppt at 6mbsf to 33 ppt at 19mbsf. Below that,

salinities decreases slightly to ~31.7 and stays around this value for the next

70m thick segment. Intermittently layers with much fresher values are

encountered, corresponding to gas hydrate bearing sections (mainly sand).

For the deeper segment with fresher values appearing at 190-225mbsf, it is

speculated that the BSR may appear at the bottom of this segment (i.e. at 225

mbsf), and the dissolved gas hydrate near the BSR caused the fresher salinity

values in the recovered cores. However, complete post-cruise analyses

indicated the BSR to be at ~260 mbsf (corresponding to the lower-most two

fresher salinity values).

To avoid interruption from the most-abundant gas hydrate zone (shown

by the fresher pore-water salinity values) when utilizing the equation of state

of sea water (Fofonoff, 1985), the values from the 50-180mbsf segment are

not considered, according to the anomalous fresher values in the chlorinity

data set. To determine the in-situ temperature, the temperature values are

interpolated from the in-situ geothermal gradient of 0.054 ºC/m and a

starting temperature of 3.5 ºC. Fig 3.5 (b) shows the final result of the in situ

water resistivity.

Using the standard Picket plot technique to determine the empirical

Archie parameters yields the values 4.65 and 0.27 for a and m respectively

(Figure 3.6).The segment of 50 – 180mbsf was excluded from the analyses

because gas hydrate appears to be very abundant. When m is again fixed at

1.76, the result of the liner regression using density-porosity yield a=1.57; the

core data yield a value for a =2.3; and the neutron porosity data yield a value

of a=2.65.

In Fig 3.7 (a), the red line represents gas hydrate saturation calculated

from density-porosity, and the blue line is calculated from core derived

29

porosity. Large amounts of gas hydrate appear to be present in the interval

from 70mbsf to 100mbsf. Gas hydrate saturation reaches to ~80% of the pore

space. Generally, using density-porosity yields larger values of gas hydrate

saturation than using core-derived porosity. For example, during the first

75mbsf segment, the average saturation derived from density-porosity data is

50% higher than the core-derived porosity estimation. Also, from 100mbsf to

the bottom of the borehole, the log-derived saturation yield higher value than

the core-derived saturation, especially at 200-250mbsf, the log-derived

saturation value is much higher than the core-derived saturation value, about

~4 times higher. However, in the segment with maximum hydrate content, the

saturation results from core-derived porosity are a little higher than that from

density-porosity. Below 240mbsf, core-derived gas hydrate saturation still

have an increasing trend, however, log-derived saturation is decreasing.

In Fig3.7 (b), the results were all calculated from density-porosity values;

the yellow line stands for gas hydrate saturation using the empirical value

m=0.46, a =4.05, which is obtained through linear fitting of the scattered

logarithm porosity data in terms of logarithm FF data. The discrepancy

between the two results calculated from m=1.76 and the fixed m value also

appears at the bottom 20 meters.

3.3.3 Site U1327

Figure 3.8(a) shows that salinity values from the recovered core are

continuously decreasing downhole from the seafloor to ~100 mbsf. Below this

depth, the salinity depicting the assumed background trend (i.e. in situ

condition) remains constant. However, the 100-225mbsf segment shows

several outliers of fresher salinity values, where gas hydrate may have

appeared in situ, and dissociation upon recovery has added fresh water to the

recovered sediment. When calculating the Rw value, the geothermal gradient

of 0.059ºC/m was used (with a seafloor temperature of 3.4ºC). The depth

interval where gas hydrate is most abundant was excluded from the calculation,

and the polynomial fitting method was used to best fit the scattered resistivity

values (Fig. 3.8 (b)).

30

Using the standard Picket plot technique (Figure 3.9 (a)) to determine the

empirical Archie parameters yields the empirical values 3.74 and 0.26 for a

and m respectively for the LWD density porosity data (with the segment of

115 – 145mbsf excluded from the analyses because gas hydrate appears to be

very abundant). The core-derived density porosity data yields the results of

a=3.74, m=0.21 (Fig. 3.9 (b)). With a fixed value m=1.76, the empirical value

a was calculated from density porosity data to 1.27, which is the most

proximate to the results from Hyndman et al (1999) of a=1.41. The result

calculated from core-derived data is a=1.99, and by using neutron porosity, the

empirical value is a=2.11.

Fig 3.10 (a) shows the gas hydrate saturations calculated from all porosity

data and different Archie parameters. Gas hydrates largely exist within the

120-145mbsf interval. In the upper 145m sediment, the gas hydrate saturation

estimation with empirical parameters a & m equals to 3.74 and 0.26 is higher

than the saturation value calculated with m=1.76 and a=1.27. However, during

the lower part from 145mbsf to the bottom of the borehole, the saturation

value with empirical parameters m=1.76 and a=1.27 is higher than the one

with empirical parameters m=0.26 and a=3.74. From Fig. 3.10 (b), the red line

represents the gas hydrate saturation estimation using empirical parameters

m=1.76 and a=1.27, and the yellow line is the gas hydrate saturation value

obtained using the empirical parameters m=1.76, a=1.99. The values of the red

line, which is from density-porosity, are higher than the results using the core-

derived porosity. Below the inferred depth of the BSR (~240mbsf), the

saturation values from log-derived data increase from 10% to 30%, which may

indicate that free gas existed in these segment as free gas may increase

resistivity values.

3.3.4 Site U1329

From Fig. 3.11, it can be seen that the core-derived salinity values jump

down from a value ~34ppt to 32ppt within first 10mbsf. The salinity remains

constant for another 15meters, and then drops to 30.5 at ~50mbsf. Below that

depth, the salinity values remain around 30-31 ppt. However, at ~175mbsf,

31

due to a different lithology of tertiary sediments below an unconformity,

abnormal fresh salinity values appear.

The assumed BSR depth is at ~125mbsf but no freshening from recovered

gas hydrate is observed. However, the chlorinity values obtained from the

pore-water do show a slight freshening, which was interpreted to represent

small amounts of gas hydrate of less than 5% of the pore-space (Riedel et al.,

2006a).

The portion of the pore-water salinity data below the unconformity at

~175 mbsf was excluded from the final calculation of estimated in situ pore-

water resistivity. To calculate the in situ water resistivity, a local geothermal

gradient of 0.072ºC/m was used with a seafloor temperature of 3.34ºC.

Using the log derived density porosity data, the empirical values of m and

a are determined to 0.32 and 2.94 (Fig. 3.12), respectively; the results from

the core-derived density porosity yield m of 3.02 and a of 0.26. With fixed

m=1.76, the value for a is 1.18. Using the core-derived porosity data only and

fixing m to 1.76, the results for a is 1.43. The log derived neutron porosity

data yield the empirical value of a=1.44 with m=1.76.

Fig. 3.13 (a) shows gas hydrate saturation levels derived from log density-

porosity data using two sets of Archie parameters. The red line depicts gas

hydrate saturations generated using the empirical data a=1.18 and m=1.76,

which reaches 20%-30% where gas hydrate may appear near the BSR. The

blue line is the saturation estimate with the empirical parameters a=2.94,

m=0.31. Most of the gas hydrates concentrate estimate falls below 10% of the

pore space. In Fig. 3.13(b), the yellow line represents gas hydrate saturation

from core-derived data.

3.4 Gas hydrate amount estimated from acoustic

transit time logs

3.4.1 Introduction

Acoustic transit time logs measure the propagation of elastic vibrations

32

through the sediments. There are mainly two types of elastic waves:

compressional, and shear waves. The distinct physical properties of gas

hydrate bearing sediments have strong impacts on the velocity and attenuation

of compressional and shear waves. Gas hydrates exhibit relatively high elastic

velocities, compared to the pore filling fluids. P-wave velocity in marine

sediments typically increases with depth as porosity decreases because of

compaction. S-wave velocities may be altered more strongly if gas hydrates

forms in such a way that the overall matrix is stiffened, for example, gas

hydrates acting as grain cement. The compressional acoustic velocities in

methane hydrates usually yield 3300-3800 m/sec, and the shear wave

velocities are usually 1800-1900m/sec. Downhole acoustic data were acquired

in low-velocity; hydrate-bearing formations at five sites drilled on the

Cascadia Margin during the IODP Expedition 311, including wire-line sonic

logs at Sites U1325, U1326, U1327, and U1329, and LWD acoustic logs at all

four transect sites.

Additionally Normal move-out (NMO) velocities from MCS line 89-08,

along the transect of wells were utilized. The amount of in situ gas hydrates is

estimated at the transect sites using all these velocity information combined.

Gas hydrate saturation is derived using published models and the best estimate

of the background non-hydrate Vp depth-profile (e.g. Yuan et al., 1999) at

these sites and compare results with independent resistivity-derived saturations.

At all these sites, the background Vp increases from ~1550 m/s at the

seafloor to ~2000 m/s at ~300 mbsf. Gas hydrate-bearing intervals appear as

high-velocity anomalies over this trend because solid hydrates stiffen the

sediment and effectively reduce porosity. Hydrate saturation was derived

using published models and the best estimate of Vp at these sites and compare

results with independent resistivity-derived saturations.

(a) LWD acoustic logs

The Schlumberger SonicVision LWD tool was deployed at all sites

during IODP Expedition 311. The principle of the SonicVISION tool is

similar to that of the wire-line sonic tools (Schlumberger, 1989), which

records monopole acoustic waveforms downhole. The acoustic slowness (or

33

velocity) is then obtained by processing the recorded waveforms. The

monopole source produces a ~13 kHz pulse that travels into the formation and

refracts back into the borehole. More than 1400 m of LWD sonic waveform

log data were successfully recorded (Goldberg et al., 2008, Riedel et al., 2006)

describe preliminary compressional velocity (Vp) logs computed from the

leaky-P wave mode in these data. We produce new estimates of gas hydrate

saturations using the wire-line and LWD sonic waveforms and compare the

results with the saturation calculated from MCS data. The results of the sonic

measurements made while drilling are highly affected by drilling noise.

Logging-while-drilling (LWD) sonic technology, however, is challenged to

recover accurate P-wave velocity in shallow sediments where velocities are

low and approach the fluid velocity. Low formation Vp make the analysis of

LWD sonic data difficult because of the strong effects of leaky-P wave modes,

which typically have high amplitudes and are dispersive. The frequency

dispersion of borehole leaky-P modes was examined and a minimum depth

(approx 50-100 m) below the seafloor at each site where Vp can be accurately

estimated using LWD data was established (Table 3.2). Below this depth, Vp

estimates from LWD sonic data compare well with wire-line sonic logs and

VSP interval velocities in nearby holes, but differ in detail due to local

heterogeneity (Goldberg et al. 2008).

(b) Sonic wire-line log measurements

Wire-line logging with the dipole sonic imager (DSI) tool was utilized to

measure the P-wave velocity at all sites during IODP Expedition 311. The DSI

measures P- and S- wave transit times between a sonic transmitter and an array

of eight receiver groups at 15cm spacing (Riedel et al. 2006a). P- and S- wave

transit times (slowness) are used to compute P- and S-wave velocity. The

vertical resolution is ~107 cm, and the depth of investigation is ~10cm (Chen,

2006).

The quality of the data collected by sonic log is highly dependent on

borehole conditions, and requires good contact between the tool and the

borehole wall. Because the caliper arm was accidentally broken off during the

expedition, the borehole radius information is not available. However, the log

data seem to be generally of good quality and in agreement with other velocity

34

data sets (Riedel et al., 2006a).

(c) Vertical Seismic Profile (VSP) measurements

VSPs provide a direct time vs. depth relationship from the direct arrival

time of the compressional wave at each receiver depth. The well seismic tool

(WST) was used at Site U1327 to produce a VSP. A geophone on the WST

recorded the full waveform generated by a seismic source (a 1.68 L (105 3in )

Generator-Injector air gun) just below the sea surface. The WST was pressed

against the borehole wall at 5m intervals, and the airgun was fired between 7

and 15 times at each geophone depth (station). The measured VSP had 16

stations between 181 and 276m below the seafloor. The root mean square

(RMS) velocity picks were converted to interval velocities using the Dix

equation (Dix, 1955) without further constraints.

(d) Multi-channel seismic (MCS) interval velocity measurements

A detailed velocity analysis was conducted as two MCS line 8908 and

8910 at the proposed drilling locations (Spence et al., 1991a and 1991b). The

lines were shot by Digicon Geophysical Corporation in 1989. The survey was

collected using a DSS-240 recording system. The airgun array source was a

tuned array with a total volume of 125L (7820 3in ). Shots were recorded by a

3600m streamer with 144 hydrophones, towed 183m behind the vessel. The

shot point interval was 50m, for a bin spacing of 12.5m, and yielded 36

common-depth-points (CDP) fold. The record length was 14s at a 4 ms

sampling interval.

There is generally good agreement between log, VSP and MCS velocity

measurements at the well sites, with a few exceptions (Fig.3.14):

(1) At Site the MCS velocity values are higher than the LWD and WL

velocity data. Also the WL velocity data were only recovered over 108 meters,

from 73mbsf to 181mbsf and show signs that the measurements were affected

by deteriorated borehole conditions. LWD measurements below 140mbsf are

also affected by borehole enlargement (Riedel et al., 2006a)

(2) WL velocity data are consistently higher than LWD velocity at Site

U1326, with a dramatic ~250 m/s discrepancy right below the BGHSZ. The

35

distance of 23 meters between the LWD (Hole U1326A) and WL (Hole

U1326D) boreholes may introduce lateral variability. A high velocity layer

found in the U1326 sonic log at 60-90mbsf was not observed in the MCS

velocities, and may be explained by the offset between Site U1326 and MCS

line 89-08 (Chen et al., 2006).

(3) At Site U1327 the WL velocity data remain relatively constant at

~1550m/s over the first 50meters measured, whereas the LWD data show an

increasing trend over the top 120 meters. However, the top-most 71mbsf in

LWD data are regarded unreliable (Table 3.2) [Goldberg et al. 2008].

Significant differences between the LWD and the two wireline logs in the

various boreholes at Site U1327 are mainly due to structural control on the gas

hydrate occurrence and the lateral distance of > 80 m between the boreholes.

3.4.2 Principle of using acoustic logs to estimate gas

hydrate saturation

Until now, acoustic modeling of gas- hydrate accumulations has been

taken to two general directions dealing with either the development of

empirical acoustic relations or the application of multiphase wave theory and

grain contact models to directly calculate gas-hydrate concentrations for

various geological conditions (e.g. Lee et al., 1993, 1996). Numerous velocity

models and equations have been published to describe the effect of gas

hydrate on the acoustic velocities. Earlier studies of the velocity model for

gas hydrate bearing sediments used various forms of time average equations

(Miller et al., 1991; Bangs et al., 1993; Wood et al., 1994). Furthermore, a

new velocity model which is derived by Hyndman and Spence (1992),

predicts the velocity of hydrate-bearing sediment by assuming the effect of

porosity reduction by gas hydrate saturation is identical to the normal

compaction porosity reduction with depth. Lee et al. (1996) also introduced

and applied the multiphase wave scattering theory developed by Kuster and

Toksoz (1974), to assess the acoustic properties of gas hydrate saturation. An

alternative gas hydrate acoustic model has been developed by Dvorkin et al.

(1991) and Dvorkin and Nur (1993, 1996), which is based on estimating

36

acoustic properties of cemented gas-hydrate-bearing sediments from grain

contact theory. Most recently, an additional acoustic model for gas-hydrate –

bearing sediments, based on effective medium theory, has been proposed and

tested by Helgerud et al (2000).

The variety of the proposed gas hydrate acoustic models has been

examined and compared in series of recent publications (Guerin et al., 1999;

Helgerud et al., 2000; Chen, 2006; Goldberg et al, 2008). This published

series of analyses affords us with the opportunity to test and compare the

results of the proposed models and theories. The detailed examination can be

found in the references mentioned above and are therefore not included in this

thesis.

The purpose of this chapter is to compare observed velocities of gas

hydrate bearing sediments from the empirical weighted average equation,

based on Wood (1941), Timur (1968), and Lee et al. (1993). The data used in

this study are a suite of logs including LWD acoustic logs, WL logs, the

velocity data from MCS profile and the VSP data at Site U1327.

3.4.3 Methodologies

(1)Timur equation

Timur modified Wyllie’s time-average equation and developed the first

three component time average equation that can be used to directly calculate

the volume of gas hydrate within a sediment section (Timur, 1968). Timur’s

time average equation is as follows:

(1 )1 1h h

b w h m

S SV V V V

φ φ φ− −= + + (3.4.1)

With φ = Porosity (fractional %), hS =Hydrate saturation (fractional %),

bV =Log-measured bulk compressional velocity (km/s), wV =water

compressional velocity (km/s), hV =Hydrate compressional velocity (km/s),

mV =Matrix compressional velocity (km/s).

37

The following parameters will strongly affect the observed velocity

behavior of some sedimentary rocks (Lee et al., 1993). They are (1) high clay

content of the sediments, (2) the degree of rock consolidation, (3) presence of

organic matter in the sediments, and (4) secondary porosity.

A modified version of the Wood equation (Wood, 1941), which is

approximately valid for particles in suspension, could be used to overcome the

problems with the time-average equation. The modified Wood equation can

be written as:

2 2 2 2

(1 )1 1h h

b b w w h h m m

S SV V V V

φ φ φρ ρ ρ ρ

− −= + + (3.4.2)

The parametersφ , hS , bV , wV , hV , mV have the same meaning as in the

Timur equation, bρ =Bulk-density ( 3/g cm ), wρ =Water density ( 3/g cm ),

hρ =Hydrate density ( 3/g cm ), mρ =Matrix density ( 3/g cm ).

Bulk density can be calculated from the following equation:

(1 ) (1 )b m h w h hS Sρ φ ρ φρ φρ= − + − + (3.4.3)

Based on the results of Timur and Wood, Lee et al. (1993) proposed an

acoustic equation for interval velocities within marine gas-hydrate-bearing

sediments that uses weighted means of the Timur and modified Wood

equations, that is:

(1 ) 1 (1 )1 [ ] [ ]r r

h h

b Wood Timur

W S W SV V V

φ φ− − −= + (3.4.4)

r = a constant simulating the rate of lithification with hydrate

concentration, when there is no gas hydrate saturation, r =1. W is the

weighting factor, WoodV is the results of the Wood equation (km/s), TimurV is the

results of the Timur equation (km/s).

The log derived density-porosity is calculated through the following

equation:

= b m

b w

ρ ρφρ ρ

−−

(3.4.5)

38

In this equation bρ is the bulk density ( 3/g cm ), wρ is the in situ water density

( 3/g cm ).

The values employed for the constant parameters were adapted from Collett

(2000) and are listed in Table. 3.3. The gas hydrate structure in this study area

is assumed to be all structure- Ι , thus, the compressional wave velocity of

methane gas hydrate is ~3.35km/s (Table 3.3).

3.4.4 Determining the Empirical weighting factor W

The most important component of using empirical weighted average

equation for the P-wave velocity is to determine the various empirical

parameters. To achieve this goal, a reference velocity depth profile for no-gas-

hydrate and no-free-gas-bearing sediments is necessary. Such a reference can

either be theoretically calculated, determined from regional seismic velocity

studies where it is concluded that there is no hydrate or gas (e.g. Yuan et al.,

1996), or it can be determined by using downhole porosity logs and

empirically derived functions that relate porosity and log velocity.

The weighting factor W is critical in predicting gas hydrate saturations

from velocity- and porosity logs. It is best determined for those intervals of

the velocity-logs where no gas hydrate is supposed to occur, which is defined

from comparison to the resistivity log data. The following text describes the

search for the best W-value to predict the observed velocity logs using the

various equations given above.

Suppose the porosity data range between 0 and 1. The bulk density can be

predicted for the values defined in Table 3.2 and the generic porosities simply

from equation (3.4.5):

( )*b m m wρ ρ ρ ρ φ= − − (3.4.5)

Given no gas hydrate is present within the sediment, the composite Lee-

weighted P-wave velocity values can be calculated from Equations 3.4.1 -

3.4.3. A series of W values was substituted into the equation 3.4.3 (assuming

39

Sh=0) to get different velocity values by using the generic porosity data. The

cross-plot of porosity and Lee P-wave velocity yields simple best-fit equations

for each chosen W-value to quickly predict the Lee weighted velocity without

the need to go through the Timur and Wood functions every time. Once the

simple best-fit equations are determined for a variety of W values, the LWD

porosity and bulk density values are used to define P-wave velocities for the

same set of W-values. The different velocity data sets are then compared with

the directly measured wire-line velocity data (the gas hydrate bearing intervals

were excluded from this analysis) and that velocity profile that is closest to the

real measurements defines the best-fit W-value.

3.4.5 Determining the Gas hydrate saturation

By substituting the known values for parameters as listed in Table 3.3 into

the equation 3.4.1 and 3.4.2, the results of the Wood and Timur velocities

were calculated using Mathematica. These results were then substituted into

Lee’s equation and the final result of gas hydrate saturation at each site is

calculated by a Matlab code (Appendix I).

Mathematica yields the result as the equation below:

φφ **10*23.3*10*85.310*01.210*78.8

444

4

hTimur S

V−+

= (3.4.6)

)***10*12.1**10*4.1*10*44.6(10*82.5

212119

10

φρφρρ bhbb

woodS

V−+

−= (3.4.7)

These equations can be solved to yield hydrate saturation (Sh) for the known

log values of bulk density (ρb), porosity (φ).

40

3.5 Velocity analyses estimates for the IODP

Expedition 311 transect drill sites

3.5.1 U1325

Based on the chlorinity data, the sections showing fresher values have

been deleted from the process to estimate the best-fit weighting factor W.

Thus, the rest of the section can be considered as the gas-hydrate free

reference. At Site U1325, the fresher values in the chlorinity data are seen

from 70 mbsf until the depth of the estimated base of GHSZ, which is ~240.5

mbsf, also corresponding to the high resistivity and high velocity values

throughout this interval. Excluding the hydrate-occurrence zone, the data yield

a most preferable W value of 1.3.

Using this empirical value W=1.3, the gas hydrate saturation can be

calculated from the LWD velocity, WL velocity and MCS velocity (Fig. 3.15).

Gas hydrate-bearing intervals correspond to intervals of anomalously high

measured resistivity and P-wave velocity. The upper 50m show no gas hydrate

appearance from the LWD data (as expected). Throughout the next 50 m thick

section (50 – 100 mbsf), several thin layers with high saturation values are

present though these remain speculative if indeed related to gas hydrate due to

lack of corresponding chlorinity freshening in this interval. The results from

the WL data yield ~5% higher saturations than the LWD approach, but overall,

the results from these two data sets follow about the same trend. The MCS

velocity data from Site U1325 indicate a gradual increase in gas hydrate

saturation from near 0 at the seafloor to 30% at the BSR. Large amounts of gas

hydrate appear during the interval from 100mbsf to the BSR depth, yielding an

average gas hydrate saturation value of 10%, with the maximum value

reaching 22% (all percentages are of pore-space). LWD data show high

saturation values below the BSR depth of ~250mbsf, almost reaching 25%,

which may be caused by the effect of the free gas, but is likely to be an effect

of erroneous velocity measurements.

41

3.5.2 U1326

At Site U1326, most of the fresher chlorinity values appear within the top

100 m and across an interval of 150-250 mbsf, corresponding to the extremely

high resistivity and high velocity values in the logging data. By excluding this

part, the velocity data from remaining sections yield the preferred W value of

1.2.

Using the empirical value W=1.2, the gas hydrate saturation can be

calculated from the LWD velocity, WL velocity and MCS velocity (Fig. 3.16).

Gas hydrate-bearing intervals correspond to intervals of anomalously high

measured resistivity and high measured velocity. The first 50meters show

several layers with saturation value reaching 10%, but are likely not related to

gas hydrate, but measurements of poor quality. WL data show extremely high

saturation across the section of 70-100 mbsf reaching > 40%. However,

similar saturations are not shown in the LWD data. The gas hydrate saturations

from LWD data are on average ~10%, with the maximum saturation of only

15%.

The WL data yield high saturations across the interval of 150-250mbsf,

with a gas hydrate saturation of ~20%-30%. Again, the LWD data do show too

much gas hydrate saturation, except for some thin layers with saturation of

10%-15%. The BSR is located at ~260 mbsf, and below this depth, high

concentration values may indicate free gas, but could also be related to noise

in the porosity and velocity data. For the most part, the MCS data yield similar

gas hydrate estimates as the result from WL data, except the higher

concentrations in the shallow section especially (~50 mbsf). This may be

caused by the effect from lateral heterogeneity, since the MCS line 8908

located 1.3 km SE away from Site U1326.

3.5.3 U1327

At Site U1327, most of the fresher chlorinity values are within a section at

110 mbsf to 150 mbsf and again below 150 mbsf to the BSR depth (230 mbsf).

These sections are also deleted for the W-factor calculations because of the

42

high resistivity and high velocity log values. By deleting this part, the

remaining data yield the preferred W value of 1.3.

Using the empirical value W=1.3, the gas hydrate saturation can be

calculated from the LWD velocity, WL velocity and MCS velocity (Fig. 3.17).

There is no gas hydrate in the top-most 50 mbsf according to the LWD data

estimate (correlating well with all other observations from chlorinity and core

infra-red imaging). MCS data yield a higher estimate than the LWD data and

reaches 5%-10% on average. Within the next 50meters, gas hydrate-bearing

intervals correspond to intervals of LWD and WL velocity, yielding the

consistent estimate between the two well log data of ~5% gas hydrate

saturation. The most prominent feature of U1327 is between 120-150mbsf,

marked by extremely high gas hydrate concentration estimate reaching >40%

according to LWD data and >20% according to WL data. Relatively high gas

hydrate concentration still remains present to the depth of BSR, with average

values of ~10%-20%. Below the BSR, the saturation is up to almost 20%

again, possibly an effect of erroneous calculations due to poor data of porosity

and velocity.

3.5.4 U1329

At Site U1329, less gas hydrate was determined to be present than at any

other site. However, a section around the depth of the BSR (124 mbsf) is

deleted since it has been proposed from the core data (mainly chlorinity) that

gas hydrate may be present just above the BSR. After deleting this part, the

data from rest section yield preferred W value of 1.5.

Using the empirical value W=1.5, the gas hydrate saturation can be

calculated from the LWD velocity, WL velocity and MCS velocity (Fig. 3.18).

There is almost no gas hydrate present within the top 50 mbsf. LWD data yield

an average estimate of less than 5% during the section of 50-100 mbsf. Below

100 mbsf, the average concentration rise a little and have values around 10%.

Estimates from WL data give the same result within this section. The

concentration reaches a maximum of 20% just above the BSR depth at

120mbsf, and decreases to 0 below the BSR.

43

Chapter 4 Regional Seismic Analyses and

Synthetic Seismograms

The major objectives of this chapter are to define through seismic analyses

(a) the regional tectonic setting associated with the formation of gas hydrate,

(b) the geologic structures such as those that control fluid flow, (c) the

properties of the BSR and (d) the potential linkages between gas hydrate

occurrence and the BSR. To tie the reflectors with specific layers and

individual formations, which are already known from the information

provided by the well logs and cores, synthetic seismograms are generated.

These synthetic seismograms relate the depth based on log data with the time-

based seismic data. In other words, a time-depth chart, or a velocity-depth

function is generated.

A common problem is that the synthetic seismogram does not provide a

good tie to the acquired regional seismic data. This could be solved through

use of additional data and preparation of a suite of different synthetics from a

range of parameters. Some compromises may be made for inadequate log data,

e.g. when the acquired wire-line (WL) P-wave velocity log is not covering the

entire borehole section, such as at Site U1325, or when the borehole diameter

(caliper) is too large to acquire high-quality logs. Also, added features, such as

AVO analyses, could be considered as complementary methods to achieve

preferable results. The efforts to include the AVO method has been previously

published by e.g. Riedel (2001) and Chen et al. (2007). However, no such

methods are included in this thesis.

The BSR with its bright reflection amplitude and the typical reflection

pattern (e.g. cross-cutting other layers) yields one way to define the base of

gas hydrate stability zone (BGHSZ). A first estimate of the BGSHZ was

determined seismically using a constant velocity, and was then refined using

log and chlorinity data. In this thesis, the seismic imaging and synthetic

seismograms have provided further ways to constrain the BGHSZ. The best

estimates for the BGHSZ are summarized in Table 4.1.

44

4.1 Site U1325

4.1.1 General Seismic description

Site U1325 is located near the southwestern end of the margin-

perpendicular transect and is within a major slope basin that developed

eastward of the deformation front behind a steep ridge of accreted sediments

(where Site U1326 was drilled). At Site U1325 a total of 6 seismic lines are

used to describe the tectonic setting including MCS line 89-08, 89-11, MCS

lines ODP7, ODP3, ODP2, as well as line CAS02B_line05_04. The MCS

lines from two surveys with different seismic sources of different frequency

content and additional SCS and high-resolution 3.5 kHz data are compared.

The multibeam bathymetry map (Fig. 4.1) along the transect line across

the accretionary prism offshore Vancouver Island shows that the seafloor in

the western part of this slope basin is relatively flat, but at the eastern part the

seafloor (~1200 m water depth) becomes gradually shallower before it rises

rapidly where it ultimately forms a plateau of the second main accreted ridge.

From the bathymetry data (Fig. 4.1), it can be seen that the south-western

portion of the basin floor is relatively flat, and within this part of the basin the

sedimentary layers are well developed and almost parallel to the seafloor

(Fig.1.2, 4.2), indicating an evenly sedimentation process inside the basin.

The seafloor at Site U1325 is covered by a thin interval of Holocene

sediments seen as a highly-transparent layer in the 3.5 kHz profile (Fig. 4.3).

The uniform Holocene sediment cover indicates regular pelagic sedimentation,

without influx of new turbidite sediments, which would otherwise disrupt the

high-porosity, low-density layer (Novosel, 2005) (Fig. 4.3).

The landward edge of the basin (North-East) where Site U1325 is located

is disrupted and deformed. Three parallel scars (small-scale canyons) are

located ~2.5km southeast from the Site U1325, which may be linked to the

tectonic uplift of a buried ridge best seen in Fig. 4.4. The seismic line ODP2

(parallel to MCS line 89-08) cuts through these seafloor-scars but at a ~2.5km

offset to the SE (Figure 4.5). However, the high frequency seismic data cannot

provide any definitive information about the BSR occurrence.

45

The seismic profile MCS line 89-08 (with a main frequency of ~40Hz)

runs in a southwest-northeast direction. This seismic line shows that Site

U1325 is located at the northeastern end of the basin at the onset of highly

disturbed and deformed sediments (Fig. 1.2, 4.4) associated with an uplifting

ridge of accreted sediments. The seismic data suggest that the sediments cored

at Site U1325 are similar to those seen in the seafloor-parallel package of

sediments in the southwest side of the slope basin (between CDPs 1150 –

1300), but have been deformed and tilted as a result of tectonic movements

associated with that buried ridge of accreted sediments (seen between CDPs

1320 and 1440). The seismic profile of MCS line 89-08 around Site U1325

shows a prominent BSR which extends ~ 1km through the buried ridge.

Starting from the most eastern side of the well-bedded basin, the BSR only

reaches the eastern limit of the buried ridge and the continuity of the BSRs is

abruptly truncated where the dipping sediments cover the western side of the

ridge. However, the BSR is strong and continuous within the buried ridge, and

roughly follows the pattern of the seafloor. It weakens again on the eastern

side of the buried ridge and is lost where the seafloor steepens to rise to the

plateau of the second major accreted anticline along the transect.

In contrast, the sediments to the western side of the basin are nearly

undeformed and seafloor-parallel along line MCS 89-08.There is no BSR

visible in this almost un-deformed part of the basin (compare to observations

along line 89-11, Figure 4.2). The gas hydrate occurrence in the well-bedded

slope basin sediments is still unclear. On one side, gas hydrate may not appear

in the sediments because the permeability of the well-bedded sediments is

lower than the deformed unconsolidated sediments, thus, the upward fluid flux

carrying free gas is inhibited and there is only little or no gas hydrate formed

in these sediments and no gas accumulated at the BGHSZ (e.g. Zühlsdorff et

al., 2001). On the other side, the progressive tectonic subsidence and ongoing

sediment deposition in the slope basin will cause the BGHSZ to always move

downward, so any gas layer that accumulates at the BGHSZ will turn to be a

gas hydrate layer and the BSR would be weakened (e.g. Von Huene and

Pecher, 1999). This model would suggest hydrate to be present within these

sediments and the lack of a BSR is simply the result of no free gas at the

46

BGHSZ. Without any new deep core penetrating the entire GHSZ, the

different models cannot be verified.

Several faults cut the sediments in the immediate vicinity of Site U1325

and some of them cut through the top of the ridge, even distributing the

sediments draped across the ridge. Some of the faults reach the BSR (Fig 4.4),

which may be considered as possible pathways for upward migrating fluids.

2.5km NW of MCS line 89-08, two higher frequency MCS seismic lines

ODP3 and ODP2 provide more information about the local tectonic structure

inside the slope-basin. Fig. 4.5 shows the image from MCS line ODP3, where

the buried ridge has already pierced through the seafloor. The discontinuous

seismic reflectivity within the ridge indicates the highly deformed sediments

underneath. A BSR can be traced in patches along the SW side of the buried

ridge. It can be found on the edge of the basin but fades away inside it. The

basin sediments layers have undergone progressive deformation, and the

sediment layers are folded and uplifted severely at the edges of the basin.

Fig.4.6 shows the seismic profile of MCS line ODP2, the boundary of the

buried ridge is harder to distinguish than in line MCS 89-08. However, the

bedded reflectors show an abrupt truncation, which is considered as the

western boundary of the buried ridge. A BSR is difficult to identify in this line,

but the progressive deformation of the basin sediments layers is marked by the

highly disrupted sediment sequences on the NE side of the basin.

The higher frequency seismic MCS line ODP7 (Fig. 4.7) is coincident to

MCS line 89-08. The unconformity on top of the buried ridge is clearly seen as

a higher amplitude seismic reflector. The high frequency seismic data give a

much clearer view of the dome-like structure inside the buried ridge. The BSR

is weak in this seismic profile and can be barely traced on the southwestern

side of the ridge as expected and described above. Comparison to seismic data

from MCS Line 89-08, which has half the frequency content than line ODP7,

shows a much weaker BSR in the higher frequency data. This weakening of

the BSR reflection strength is likely an effect of the higher frequency content

in the MCS data of line ODP7 (and other lines from that 1999 survey). The

BSR is a transition zone of several meter thickness, in which the velocity

gradually decreases, instead of a single sharp interface as explained in the

model by Chapman et al. (2002).

47

Large amplitude but discontinuous seismic reflectors are detected above

the depth of BSR in the centre of the buried ridge. The reflectors have

different dips on either side of the buried ridge and extend deeper into the

sediment to below the BSR on the northeastern corner of the ridge. Dense

faults cutting through the buried ridge may act as the pathways for upward gas

flux and result in gas hydrate formation where higher porosity sediment

appears in gas hydrate occurrence zone (Fig.4.7). The reflection amplitudes of

these truncated layers just above the BSR in the centre of the ridge are about

twice as large as the surrounding reflectors. Although it is difficult to discern

their polarity, it is unlikely these bright spots are free-gas related, as they are

well within the gas hydrate stability zone (above the BSR). More likely, they

could reflect higher hydrate concentrations from their proximity to the more

densely faulted area. The highest flux is expected inside the centre of the

buried ridge and thus more gas hydrate could form there, compared with

outside the ridge, although the faults do reach to shallower depths.

A series of five parallel SCS lines were acquired around Site U1325 with a

perpendicular direction of MCS line 8908 in 2005 in response to safety

concerns raised in preparation of the IODP drilling. These seismic lines were

acquired to explore the eastern portion of the slope basin. The most landward

line CAS02B_line 05_04 passes through Hole U1325A, and shows that Site

U1325 is located in a topographic low and the sediments on both sides of this

trough rise up rapidly by 0.11-0.16s TWT. The seismic reflection events are

generally difficult to discern due to interference patters with side-echo

reflections from the steep rise in the seafloor to the east (the side-echoes

appear as steeply dipping reflectors as shown in Fig. 4.8. There is an

uncertainty about the location of the BSR from this seismic image. The

position was therefore predicted from comparison with the other seismic

profiles. However, a series of high amplitude seismic reflectors show up

around the BSR level (Fig. 4.8). The sediment layers around the BSR depth

appear tilted upwards in a similar pattern as the seafloor.

48

4.1.2 Lithostratigraphy at Site U1325

The recovered core was described and analyzed onboard during

Expedition 311 to define lithostratigraphic units and sedimentary

environments. Here the results of these findings are summarized, but a

complete description can be found in Riedel et al. (2006a).

The first lithostratigraphic subunit (0-24 mbsf) is characteristic by very

abundant, thick, coarse-grained sand layers within fine-grained (clay and silty

clay) detrital interlayers. The thickness of the sand layers suggests the

depositional environment of a distributary channel within the slope basin with

mass transport parallel to the uplifted bounding ridges.

Lithostratigraphic Subunit IB (24 – 52 mbsf) is composed of fine-grained

(clay to silty clay) detrital sediments with few thin silty/sandy interlayers from

turbidites. Hemipelagic sedimentation dominated this part of sediment. The

boundary between lithostratigraphic Subunits IA and IB is an unconformity

marked by a seismic horizon that can be traced for ~1.5 km around Site U1325.

The division from subunit 1A and 1B also separates undeformed seafloor-

parallel slope sediments from underlying more deformed and slightly faulted

sediments. There is no drastic change in the physical properties as seen in the

log data, but a rather gradually increasing (from 0.75-1 Ωm) resistivity and an

increase of the pore-water chlorinity values (~550-600 mM). Lithostratigraphic Unit II (52–102 mbsf) is characterized by fine-grained

(clay to silty clay) detrital sediments with intervals of silty/sandy interlayers,

interpreted to represent turbidite deposits. Their frequent occurrence might

indicate times of more active tectonism on the margin. The lithostratigraphic Units 1 and 2 are characterized by a well-defined

gradual increase in density with depth and a corresponding decrease in

porosity. This increase in density is matched by a corresponding increase of

resistivity with depth, from ~1 m near the seafloor to ~1.5 m at 122 mbsf.

Lithostratigraphic Unit III (102–198 mbsf) is characterized by fine-

grained (clay to silty clay) detrital sediments. The presence of unlithified

authigenic carbonate cement suggests that diagenetic processes are active in

lithostratigraphic Unit III. Soupy and mousselike sediment textures related to

49

the presence of gas hydrate are present in lithostratigraphic Unit III within the

depth interval from ~121 to ~145 mbsf.

The boundary between lithostratigraphic Units III and IV is marked by the

sudden increase in resistivity data. LWD-derived resistivity-at-the-bit (RAB)

images from Hole U1325A suggest that gas hydrate was concentrated in thin

sand layers within the interval between 173 and 240 mbsf. The LWD porosity

and resistivity logs from Hole U1325A further show that it is a very

heterogeneous gas hydrate–bearing section composed of alternating layers of

gas hydrate–saturated sands and clay-rich layers with little to no gas hydrate.

This interpretation is in general agreement with the marked freshening of the

interstitial waters observed in sampled sand layers.

Lithostratigraphic Unit IV (198 mbsf–TD) is characterized by fine-

grained (clay to silty clay) detrital sediments with few silty/sandy interlayers

from turbiditic deposits. A BSR appears at the depth of ~240.5mbsf, the

resistivity and the chlorinity data show an increase (‘bulge’) just above the

BSR from 190mbsf to 240mbsf. The resistivity data decrease to 1Ωm at BSR,

and there is a small increase just below the BSR, indicting the possibility of

free gas.

4.1.3 Synthetic Seismogram generation and Log‐to‐

seismic correlation

A synthetic seismogram is generated by convolving the reflectivity

derived from sonic and density logs with the wavelet derived from the seismic

data. By comparing marker horizons or other correlation points picked on well

logs with major reflections on the seismic section, interpretations of the data

can be improved. The quality of the match between a synthetic seismogram

depends on well log quality, seismic data processing quality, and the ability to

extract a representative wavelet from seismic data, among other factors.

Generation of the synthetic seismograms was performed using the

Kingdom Suite "SynPAK" module. In creating a synthetic seismogram, the

interpreter is able to tie time data (the seismic data) to depth data (the well-log

data) by integrating over the velocity profile. Starting and total depth of well-

50

log data together with the corresponding times (in TWT) from the seismic

data were put into the Time-depth Charts as the preliminary information to

estimate the time-depth conversion function. At Site U1325, the seafloor

depth is 2201.1m; the starting time is read from the seismic data, which is

TWT 2.94s. The end depth is 2559.16m, with a predicted time of TWT 3.31s

(using a uniform velocity). A complete time depth chart is then generated by

integrating the log P-wave velocity information or coincident MCS velocity

data (Table 4.2). The starting velocity is finally adjusted to achieve the best

result for the seafloor pick. The starting velocity at the seafloor at Site U1325

is set to 1485.8m/s, which is significantly smaller compared with the uniform

velocity estimate of 1676m/s used to generate the two starting points of the

TD function.

An impedance log and reflection coefficient is generated from the

velocity and density log data. The velocity dataset at Site U1325 is generated

from the wire-line velocity log but also from MCS interval velocities

determined for line 89-08 (Chen, 2006) since the wire-line log is only

covering a relatively short depth interval of 73 – 181 mbsf. The density data is

form the LWD Hole U1325A. The reflection coefficients are then convolved

with a seismic wavelet to produce a synthetic seismic trace. At this site the

seismic wavelet is obtained using a wavelet extraction from MCS line 89-08.

The wavelet is generated by choosing traces within a distance of 500 m

around Hole U1325A, with a starting time of 2.9 s TWT to a time of 3.5s

TWT. A total of 94 traces were selected to generate the wavelet (Fig. 4.9a).

The computed spectrum shows a frequency range of 0 – 64 Hz. Little noise

appears from 0 – 32 Hz, but from 32-64Hz, which is the main frequency

content of the seismic data, the noise is much increased (Fig. 4.9b). The

synthetic seismogram is then compared with the actual seismic traces at the

drill site. The trace at the drill site was compared with adjacent traces to

assure that it was representative of that part of the seismic section. Fig. 4.10

illustrates the relationship between the impedance logs, reflection coefficients,

MCS traces, and synthetic traces for Sites U1325.

In the following a comparison of the synthetic seismograms with the

various seismic lines, well logging information and data from core-derived

51

chlorinity data is described. Also, the gas hydrate saturations calculated in

Chapter 3 are incorporated.

In lithologic subunit 1A, which is composed of slope-basin, well-bedded

sediments, no prominent synthetic seismogram features are generated and no

gas hydrate is present.

In subunit 1B, chlorinity data follow the baseline value of ~600mM, and

the resistivity value is not changing during this interval. The MCS interval

velocity value has a sharp jump at the depth ~32mbsf, from ~1610 to ~1680

m/s but no prominent change in density and porosity values are observed.

However, the seismic profile shows a strong reflection at this depth, which

corresponds to the unconformity between the uppermost undeformed slope-

sediments to those sediments that have been tilted and deformed in response to

the tectonic uplift of the buried ridge (see Fig. 4.11). In lithologic Unit 2,

which is composed of the sand layers intersected with clay layers, several

sections of high resistivity values appear and one data point of fresher

chlorinity data appears at ~80mbsf, indicating the top of gas hydrates at this

site. The computed gas hydrate saturation shows several zones of possible

hydrate occurrence, with a maximum concentration of ~20% gas. However,

the average concentration for he entire Unit 2 is less than 5% of the pore space.

Most of the gas hydrate occurs within Unit 3 and the uppermost part of Unit 4.

Within these sections several fresher chlorinity data points are seen as well as

elevated LWD resistivity data. Average gas hydrate saturation yields ~20% for

this entire interval, although maximum concentrations reach up to 60% of the

pore space.

Resistivity and chlorinity data show a prominent “bulge” from a depth of

190 mbsf to 240 mbsf, although no equivalent seismic event is observed nor

stronger synthetic reflection coefficients calculated for this interval. This bulge

of elevated resistivities (and estimated concentrations) may reflect an overall

increase in gas hydrate concentration, but could also be an effect of unusual in

situ pore-fluid composition, although no indication for that was found in the

core data (see Chapter 3).

There is a strong BSR reflection, cross-cutting the otherwise steeply

dipping sediment reflections and there are also several stronger reflections just

below the BSR, which may indicate the presence of a free gas below the BSR.

52

MCS line ODP7 provides a more detailed view of the sediment structure

around Site U1325; for example there are strong reflectivities at the boundary

of the subunit 1A and 1B.

4.2 Site U1326

4.2.1 General description

Site U1326 is located on the first uplifted ridge east of the deformation

front, and is at the southwest (seaward) end of the drilling transect (Fig.4.13).

As seen on the multibeam bathymetry map (Fig. 4.13) along the drilling

transect across the accretionary prism offshore Vancouver Island, the frontal

Cascadia basin (abyssal plain) shows a relatively smooth seafloor surface

indicating undisturbed sedimentation characteristics. The frontal basin is much

deeper than the first interior slope basin where Site U1325 is located, ~TWT

0.445s deeper (~330 m). An anticlinal frontal ridge rises up to the east of the

Cascadia basin and has a mainly northwest-southeast strike-direction. The

ridge is asymmetric with the northwestern end being narrower, and the

southwestern part being much wider.

The most prominent feature of the frontal ridge is a collapse structure

(slope failure) located ~ 1.9km southeast from Site U1326, as well as many

steep linear fault-outcrops concentrated in the central portion of the frontal

ridge. The slump covers an area of ~2.4km by 6.5km and originates almost at

the summit of the frontal ridge. The sediments from the upper-most part are

transported down-slope along the failure surface and are deposited at the foot

of the ridge. The slope failure is very prominent in the multibeam bathymetry

map (Figure 4.13) and it exposes the pre-collapse gas hydrate stability zone to

a considerable depth (up to 200 m), consequently changing the pressure and

temperature conditions in the slide scar. Previous work in the frontal-ridge

area (Lopez 2008; Spence et al., 2008) suggested that the large (1.5 km in

width) sediment slump appears to have failed at a horizon near the depth of a

gas hydrate-related BSR. During the recent cruise 2008-007-PGC, the possible

role of the presence of natural gas hydrate in the frontal ridges and the stability

of the ridge sediment was investigated (Haacke et al., 2008). Possible trigger

53

mechanisms for the slump were examined, which included earthquake activity

and climate change related eustatic sea-level changes. However, preliminary

sedimentological descriptions and analyses, combined with pore-water sulfate

gradients and physical property data, suggest the slump occurrences are not

related to the last mega-thrust earthquake that occurred at the Northern

Cascadia subduction zone in January 1700 (e.g. Satake, et al., 1996). It is

speculated that the slump could have been triggered by earlier earthquakes.

Thus, further analyses and age determinations are underway to assess the

possible linkages between these slumps and mega-thrust earthquakes and

possible other trigger mechanisms.

A series of 2D seismic surveys, which cover the frontal ridge area, are

investigated in this thesis, including SCS lines CAS03B_inline01-inline33,

and CAS03B_Xline 01-10, as well as MCS line CAS3MCS99. Going through

the various seismic surveys from the deformation front eastward (landwards),

the prominent topography changes of the frontal ridge are clearly seen. The

western-most line CAS03B_inline 01 shows a small ridge profile with a

topographic high between Shot Point (SP) 290-400, reaching ~2.45s TWT (Fig.

4.14). Northwest of the topography high, with the seafloor drops rapidly but

well-bedded sediments can be identified. The BSR is overall weak and

discontinuous but can be identified extending from SP 320 to 420 on the

seismic profile. The area of the slump-failure can be seen as a canyon-like

feature between SP 180-290 penetrating down to 2.59s TWT. There is no BSR

observed beneath this slump area.

From the perpendicular NE-SW oriented SCS line CAS03B_Xline01-10,

the vertical structure of the frontal ridge can be defined. In CAS03B_Xline01,

which is the northwestern onset of the frontal ridge, the anticline structure with

bedded sediment layers is clearly seen between SP 1-190, and the maximum

height of the frontal ridge along this seismic profile reaches ~2.53s TWT

(Fig.4.15). Very weak and discontinuous BSRs are observed along this seismic

profile. SCS line CAS03B Xline03, which is 54m southeast from Hole

U1326A, shows similar structures and the ridge reaches to a level of

~2.5TWTs (SP 150-210). The BSR is easier to pick in this seismic profile,

which extend from SP 140 to SP 200 (Fig. 4.16). The slope of the frontal ridge

54

reaches 7 degrees which may explain the lateral heterogeneity between the

different drilling holes at this site (see discussion below).

MCS line 89-08 is located 1.3km SE from Site U1326; it is too far to

project Site U1326 onto that seismic line. Thus, the location of MCS line

CAS3MCS99, cutting through Site U1326, has been marked as the red vertical

line in MCS line89-08 profile, to show the rough location of Site U1326

(Figure 4.17). The well bedded layers of sediments west of the ridge are

marked by slight uplifting and deformation. In contrast, the sediments within

the ridge are highly deformed. The SW side of the ridge shows high

reflectivity and some sediment layers can be seen, but the seismic reflectivity

towards the NE side of the ridges appears to be transparent with very few

layers present. A BSR can be traced across the ridge from CDP 960-1030.

Fig. 4.18 shows the seismic section from MCS Line CAS03MCS99 from

the 1999 survey exactly crossing over Site U1326. This line is oriented in a

southwest-northeast direction, parallel to the two CAS03B SCS surveys shown

before (Fig. 4.14 and Fig. 4.15). The BSR is clearly visible extending from

CDP 2300-2780. The BSR is defined at a depth of 264 mbsf at Hole U1326A

(see Table 4.1), and can be traced by ~1km across the entire drill site. Some

apparent side echoes appear at this seismic line from the steep topography of

the frontal ridge, but can be easily identified and excluded from any

interpretation. The side-echoes form linear features steeply dipping relative to

the regular sediment reflectivity.

Several large scale faults cut through the frontal sediments of the ridge. A

relatively smooth slope can be seen at the northwest end of the frontal ridge.

The slope keeps rising up to the plateau of the ridge where the local geology is

mainly controlled by a dense population of faults. Most of the faults crop out

at the seafloor, causing a heavily disrupted seafloor surface with some outcrop

reaching ~0.04s TWT higher (~25m higher) than the surrounding seafloor.

The fault-controlled area extends across the central portion of the frontal ridge.

The many discontinuous layers are offset along the faults, but form relatively

strong reflection packages. These layers (reflections) are assumed to be the

host-formation for gas hydrate with the deep faults considered as the conduits

for the transport of the fluid also carrying methane gas from deeper sediments

from below the BSR.

55

Strong short dipping reflections can also be detected below the BSR, also

along the faults cutting through the BSRs, which may be considered as the

evidence that the free gas and the pore fluid transport mechanisms with the

regional faults.

4.2.2 Lithostratigraphy at U1326

The 271.40 m thick Quaternary sedimentary section cored at Site U1326

was divided into three lithostratigraphic units. Lithostratigraphic Unit I (0–24

mbsf) is characterized by fine grained (clay to silty clay) detrital sediments

with thin silty/sandy interlayered turbidite sequences. Within this unit, the

pore-water chlorinity maintains a constant value of ~560mM (See Chapter 3).

There are no remarkable changes in resistivity values except some abnormal

data points within the top 5m related to borehole enlargements and disruption

of the drilling process from instable sediments.

Lithostratigraphic Unit II (24–146.3 mbsf) has the same sedimentary

properties with Unit I, but diatoms are absent and other biogenic components

are rare. The frequent occurrence of turbidites might indicate times of active

tectonism. The most prominent segment is a 40m thick sediment section (60-

100mbsf) characterized by extremely high resistivity and P-wave velocity

values. The LWD electrical resistivity data reaches >20 Ωm, corresponding to

sharply increasing WL velocity data, which reaches >2300 m/s. However, the

LWD velocity log shows a slightly thinner layer of high velocity (~25m), and

the largest value of the LWD velocity log only reaches 2100 m/s. In contrast,

the MCS velocity data do not show such a sharply increase. The difference

between the WL and LWD log-velocity data sets may be caused by the lateral

heterogeneity of the different drill-holes due to the special stratigraphic

character on the frontal ridge and its strong fault-control. The presence of high

resistivity and velocity values is matched by abundant fresher chlorinity data

points, suggesting significant amounts of gas hydrate appearing within this

interval (60-100 mbsf).

Lithostratigraphic Unit III (146.3-271.4 mbsf) is characterized by fine

grained (clay to silty clay) detrital sediments with few, thin silty/sandy

turbidite interlayers. This interval was interpreted as mixed hemipelagic-

56

turbiditic deposition. Soupy sediment textures related to the presence of gas

hydrates are also present in this section. Chlorinity data remain at the baseline

value of ~550, with few fresher data points distributed across this section.

Resistivity data almost stay on the baseline value of ~1.74 Ωm through the

whole segment, except for two layers with high resistivity values (9.5 Ωm and

13 Ωm, respectively) around the depth of BSR (~260 mbsf). These two layers

could either correspond to some gas hydrate above the BSR (coincident to the

two fresher chlorinity data points at equivalent depths) or correspond to the

bright reflections seen below the BSR, which thus may indicate the occurrence

of free-gas.

Neither of the two interpretations can be ruled out due to the absence of a

high-resolution VSP data set resolving the time-depth conversion issue. The

time-depth (TD) curve is generated from two fixed points, typically BSR and

seafloor, and a pair of coinciding times (TWT) and log-depth are used to

create a first approximate TD-function (further explanations below).

4.2.3 Synthetic Seismogram generation and Log‐to‐

seismic correlation

The synthetic seismogram is generated to provide a linkage between the

seismic profile and well log data. At Site U1326, the seafloor depth is 1828.1

m, and the seafloor arrival time is defined from the seismic data to 2.481s

TWT. The bottom depth of the borehole is 2128.1m, with a predicted time of

2.853s TWT (using a uniform velocity of 1610 m/s). A complete time depth

chart is then generated by integrating the WL P-wave velocity information.

The starting velocity is finally adjusted to achieve the best result for the

seafloor pick. The starting velocity at the seafloor at Site U1326 is set to

1465.5m/s (Table 4.3).

The density data used in the synthetic seismogram generation is from the

LWD Hole U1326A. The reflection coefficients are then convolved with a

seismic wavelet to produce a synthetic seismic trace. At this site, the seismic

wavelet is obtained using a wavelet extraction from seismic section of MCS

57

line CAS03MCS99. The wavelet is generated by choosing traces within a

distance of 300m from Hole U1325A, with a starting time of 2.5 TWT s to a

time of 2.95 TWT s. A total of 247 traces were selected to generate the

wavelet (Fig. 4.19 a). The computed spectrum shows a frequency range of 0 –

192 Hz. The main frequency is around 100Hz, which is the main frequency

content of the seismic data, the S/N ratio is increased around 100Hz to ~15%

(Fig. 4.19 b). The synthetic seismogram is then compared with the actual

seismic traces at the drill site. The trace at the drill site was compared with

adjacent traces to assure that it was representative of that part of the seismic

section. Fig. 4.20 illustrates the relationship between the impedance logs,

reflection coefficients, MCS traces from line CAS3MCS99, and synthetic

traces for Sites U1326. Synthetic seismograms do not show meaningful results

when using any of the lines crossing the ridge in the transect-direction (SW-

NE) due to the steep nature of the structure and related poor imaging (and high

noise levels).

Lithologic Unit 1 is considered a no gas hydrate zone. A relatively good

correlation exists between the well log data and the synthetic seismogram for

this portion. Where the resistivity and velocity logs show extremely high

values within the section of 60-100mbsf, the synthetics do show corresponding

high reflectivity. However, the seismic reflection from line CAS03MCS99 is

lacking corresponding reflectivity.

As mentioned above, in lithologic Unit 3, the two strong spikes in the

resistivity log data could fit either the BSR or the horizontal layers just below

the BSR, respectively (Fig. 4.21). Without additional VSP data or strong

correlations to the seismic data at other depths, it is not justifiable to shift data

artificially to just create a better match, so no final decision can be made

regarding which of these two options were preferred.

4.3 SITE U1327

4.3.1 General Description

Site U1327 is located near Sites 889/890, where the largest amounts of

gas hydrate had previously been inferred during ODP Leg 146 (Fig. 4.22).

58

Landward from Site U1325 (located in a confined slope basin) the seafloor

rises rapidly to a water depth of 1400-1500m where there is a bathymetric

bench. The bathymetry map shows that Site U1327 is located at the NW side

of a local topographic high on top of the larger bathymetric bench. An area of

12km by 8km has been the focus of detailed seismic investigation, mainly

located near the two topographic highs, which rise ~200m over the

surrounding seafloor. The area between the topographic highs forms a 350m

deep trough filled with slope basin sediments; however, the entire basin is now

uplifted relative to the next landward slope basin, and thus is not receiving any

new sedimentation from turbidites any longer. The landward (eastern)

topographic high rises smoothly at SW side but falls down rapidly at the other

side and is cut off by a thrust fault which can be detected by the linear outcrop

on the seafloor (Figure 4.22). Site U1327 was drilled at a water depth of ~1322

meter, approximately at the midslope of the accretionary prism. There were

five holes (U1327A-U1327E) occupied at Site U1327. Drilling began in thin

bedded, slope-basin sediment and extended downward into the underlying,

deformed sediments of the accretionary wedge. The potential presence of gas

hydrate has been inferred seismically around this area by the occurrence of a

BSR (e.g. Hyndman et al., 2001).

The main objectives of drilling at this site include determining the

reference profile of well logging data of geochemical and geophysical

properties, as well as modeling the mechanism of fluid flow and linkages with

gas hydrate formation.

The SW-NE trending MCS line 89-08 (Fig.4.23), which passes Site

U1327 ~2.5km on the NW side, shows complicated deformation structures

with a mixture of erosional unconformities, faults and onlap sediments,

indicting the sedimentation at this area has undergone deformation during the

period that the basin sediments were deposited. Within the basin located

between the two topographic highs, the sedimentary sequence is well bedded

but shows some tilting and deformation; reflectivity is strong and continuous.

In contrast, the deformed, accreted sediments are generally seismically

transparent (Riedel, 2001; Westbrook et al., 1994). To the eastern side of Site

U1327, a slope basin has developed with up to ~0.6s TWT thick sediment fill;

it is worth noticing the absence of a BSR within this basin (similar to the

59

observations made at Site U1325). Different sedimentary layers can be clearly

distinguished between the ridges of accreted sediment. The BSRs around Site

U1327 can be clearly traced from the mid-slope before rising up to the plateau

where U1327 located, which is located at ~ CDP1590, towards the western

edge of the slope basin with the fault outcrop near ~CDP 2330.

Seismic line MCS99_Inline38 (from the 3D grid) almost cuts through Site

U1327A in a NW-SE direction. From the seismic profile, the sediment

structure of the topographic high and the basin located on the two sides of the

ridge are clearly seen (Fig.4.24). The basin located at the NW side of the ridge

can be interpreted as slope-sedimentation while the ridge is uplifted (syn-

deposition sedimentation) so that the sediments are progressively deformed

while being deposited. The individual layers are folded and uplifted between

CDPs 130-200. Near CDP 200, the seafloor is marked by a small depression,

associated with small scale faults. The western basin boundary can be traced

from CDP 220 to CDP 310. The seismic reflectivity within the ridge is mainly

discontinuous typical of accreted sediments that have a long history of

deformation, during which all initial reflectivity is lost. A BSR can be clearly

seen below the ridge for ~6.1km from CDP 190 to CDP 810. No clear BSR

has been identified within the basin sediment. On the SE side of the ridge, a V-

shaped basin has been filled with well bedded sediment layers.

Since none of the MCS data (Inline 38 and line MCS 89-08) are exactly

over the boreholes of Site U1327 (which were drilled in a SW-NE transect of

~60 m length), new seismic data were acquired during cruise 2008-007-PGC

in August 2008. The survey consisted of a high-frequency Huntec sparker

system (frequency of ~ 3.5 kHz) and an airgun source consisting of a pair of

10 cubic-inch sleeve guns towed at 1.5 m depth, and these produced an

impulsive signal with significant power up to about 350 Hz. The Huntec

system performed well in the mid-slope area with water depths of 1300-1500

m and provided images of the subsurface in the top ~20mbsf in the areas with

best signal penetration. The vertical resolution was typically < 1 m and

horizontal resolution is ~4-5 m. The combination of Huntec and airgun

systems provided excellent imaging of targets in the midslope area (Haacke et

al., 2008). Seismic line 2008007PGC-DTSEXT060 is located 7m SE away

from Site U1327A. It shows a relatively smooth seafloor with no clear

60

indication of faulting. To produce a better image, the Huntec data was

processed to use the seismic envelope (instantaneous amplitude) attribute. The

envelope-image shows that the Huntec signal has penetrated the top ~23 m

(estimated using a velocity of 1500 m/s). A succession of sedimentary layers

can be distinguished within the top 23 mbsf. At the intersection point with line

MCS Inline38 the seafloor was picked at 0.292 s TWT (a deep-water delay of

1.465 s was used to record the Huntec data). The thickness of the first layer is

~ 1.5 m, which is almost transparent in the seismic profile. A second layer was

identified ~4.5 m below the first layer (Fig. 4.25).

The seismic line PGC007-2008-line60 is shown in Figure 4.26. The

sediment layers are relatively well developed and show little sign of

deformation and mainly follow the pattern of the seafloor. A BSR can be

identified throughout the entire seismic profile, which appears at 0.283s TWT

below the seafloor.

The Huntec and airgun data of seismic line 2008007PGCDTSEX058 are

shown in Fig.4.27 and Fig.4.28. The seismic line is parallel to line PGC007-

2008line60 but at a 26m offset NW of Hole U1327A. The seismic features do

not differ much from the seismic line 2008007PGCDTSEX060.

4.3.2 Lithostratigraphy at U1327

At Site U1327, sediments in the upper ~170 m below the seafloor are

mostly silty clays and clayey silts interbedded with fine sand turbidites. This

sequence was interpreted as little-deformed slope basin sediments, deposited

in place (Westbrook et al., 1994). The sediment below this sequence is more

deformed, compacted and cemented, and was interpreted as accreted Cascadia

Basin sediments.

Lithostratigraphic unit Ι (0-90mbsf) is characterized by fine grained

detrital sediments like clay and silty clay, interbedded with abundant coarse

grained sand and gravel layers. The unit Ι is characterized by turbiditic

deposits, lithified carbonate nodules and rocks of various lithologies are

probably dropstones from rafting icebergs. The chlorinity values in this section

decrease strongly from near-seawater values at 550 mM, which is the normal

baseline value at other sites, to 400 mM at the bottom of this section. All

61

resistivity log datasets also have a smooth increasing trend from the seafloor at

~1.0 Ωm to 1.44 Ωm at the bottom of the section. The LWD velocity also

shows an increase of ~120 m/s across this interval.

Unit Ι and Unit ΙΙ (90 – 170.4 mbsf) are divided by a seismically

traceable boundary, which was inferred as the boundary between slope-basin

type and accreted sediments. From a lithology point of view, the boundary is

marked by a sharp decrease in sand and silt layers and the onset of more

diatom-rich sediments. The most prominent feature in this section is the 30 m

thick layer with high LWD resistivity, high LWD velocity and low chlorinity

values (compared with the chlorinity baseline at that section). However, this

phenomenon is not present in the WL resistivity and velocity data, which were

collected from Hole U1327E, located 40m NE of Hole U1327A. The intra-site

variability is striking at this site (Riedel et al., 2006a).

Lithologic Unit ΙΙΙ (170.4-TD) was interpreted as an abyssal plain, with

sediment transport and deposition dominated by low-energy turbidity currents.

The relatively lower sedimentary rate is indicative of the dominant basin-plain

setting within this section. At a depth of 240 mbsf a sharp decrease in P-wave

velocity is seen. Below 240 mbsf, P-wave velocities drops to very low values

near water velocities (~1500 m/s), suggesting the presence of small amounts of

free gas. This section also displays a small drop in resistivity compared to Unit

2. Although resistivity tends to be just above 2 Ωm in Unit 2, it is just below 2

Ωm in Unit 3. The chlorinity data shows a small decrease that drops from the

beginning 375 mM to 350 mM at the bottom of the hole.

4.3.3 Synthetic seismogram generation and Log‐to‐

seismic correlation

At Site U1327, the seafloor depth is 1395.6 m; the starting time is defined

from the seismic data to 1.87s TWT. The total depth of this site is 1622.1m,

with a predicted time of 2.15s TWT (using a uniform velocity). A complete

time depth chart is then generated by integrating the log P-wave velocity

information (Table 4.4). The starting velocity is finally adjusted to achieve the

62

best result for the seafloor pick. The starting velocity at the seafloor at Site

U1327 is set to 1495.3m/s.

An acoustic impedance and reflection coefficient log is generated from

the velocity and density log data. The velocity dataset at Site U1327 is

generated from the LWD velocity log. The density data is form the LWD Hole

U1327A. The reflection coefficients are then convolved with a seismic

wavelet to produce a synthetic seismic trace. At this site, the seismic wavelet

is obtained using a wavelet extraction from MCS_Inline38 data. The wavelet

is generated by choosing traces within a distance of 200m from Hole U1327A,

with a starting time of 1.8 s TWT to a time of 2.15s TWT. A total of 40 traces

were selected to generate the wavelet (Fig. 4.29a). The computed spectrum

shows a frequency range of 0 – 192 Hz With a relatively high noise level (Fig.

4.29b). The synthetic seismogram is then compared with the actual seismic

traces at the drill site. The trace at the drill site was compared with adjacent

traces to assure that it was representative of that part of the seismic section.

Fig. 4.30 illustrates the relationship between the impedance logs, reflection

coefficients, MCS traces, and synthetic traces for Sites U1327.

The most prominent features in Site U1327 are the abnormally low

baseline value for the chlorinity data as well as the high resistivity and velocity

interval at a mid-depth range (~120 mbsf). However, the nearly transparent

seismic reflectivity from seismic MCS99_Inline 38 around this site does not

provide any information about this hydrate-rich zone. The strong reflections in

synthetic seismogram are coincident with the high resistivity values and the

fresher chlorinity value. The most striking feature is at ~120-150 mbsf, where

the strongest reflectivity in the synthetic seismogram is seen. High reflectivity

also appears in the synthetic seismogram exactly at the BSR depth. However,

a poor correlation (less than 20%) was achieved between synthetic and real

seismic trace. The origin of this poor match is unknown, but could be related

to the fact that most of the seismic trace is within accreted sediment, which is

not showing coherent reflectivity.

The fact that none of the seismic data (MCS Inline 38, lines 58 and 60

from the 2008 survey) show a prominent reflection from the high-

concentration interval is puzzling. The five boreholes at Site U1327 are spread

63

by 60 m and over this distance the high-concentration zone is very

heterogeneous, dipping towards the NE and thins from 25 m thickness in Hole

U1327A to 0 m in Hole U1325E (Riedel et al., 2006a). In comparison, none of

the three boreholes of Site 889 drilled by ODP Leg 146 (about 500 m further

to the NW of Site u1327) showed any hint of a high-concentration zone.

Figure 4.32 shows the various logs acquired in the five boreholes of Site

U1327. This strong heterogeneity (thinning and dipping towards the east) may

be the reason that no seismic reflection data was able to image this lens-like

body. It likely diffracts energy so strongly that it cannot be coherently

recorded, especially not with a single-channel seismic streamer. The MCS

inline 38 recorded data perpendicular to the apparent dip of the hydrate-lens

and thus may not have recorded the scattered energy coming off the hydrate-

lens.

In an attempt to achieve a better match between synthetics and seismic

data, the seismic wavelet is regenerated using MCS line 89-08 (lower

frequency content than all other lines). The wavelet is generated by choosing

traces within a distance of 500 m from Hole U1327A, with a starting time of

1.7 s TWT to a time of 2.2 s TWT (Fig. 4.33a). The computed spectrum

shows a frequency range of 0 – 64 Hz with little noise (Fig. 4.33b). The

synthetic seismogram is then compared with the actual seismic traces at the

drill site (projected). The trace at the drill site was compared with adjacent

traces to assure that it was representative of that part of the seismic section.

Fig. 4.34 illustrates the relationship between the impedance logs, reflection

coefficients, MCS traces, and synthetic traces for Sites U1327.

By tying the seismic survey and synthetic seismogram together, a better

match was achieved. The negative reflectivity in the synthetic seismogram at

~240 mbsf perfectly corresponds to the negative reflectivity in seismic survey

which indicates the depth of BSR. The gas hydrate enrichment zone cannot

be expected to have a corresponding reflectivity due to the long-distant

projection of the borehole.

The Airgun PGC007-2008line60 seismic profile with log and synthetic

seismogram overlay has been shown in Fig. 4.36. The same synthetic

seismogram which was used in MCSInline38 has been superimposed on the

64

Airgun PGC007-2008line60 seismic survey because of the difficulty of

generating a wavelet from Airgun PGC007-2008line60 seismic data. The BSR

in the seismic data corresponds to a negative reflectivity in the synthetic

seismogram. Although the synthetic seismogram shows strong reflectivity in

the gas hydrate enrichment zone, this interval still correspond to a blank

seismic reflectivity zone in this section.

4.4 Site U1329

4.4.1 General description

Site U1329 is located at the eastern end of the SW-NE– trending margin-

perpendicular transect of sites occupied during this expedition. The seafloor

depth at this site is ~950 mbsf. Site U1329 is believed to be the landward end

of gas hydrates occurrence on the N. Cascadia margin. From the multibeam

bathymetry map, U1329 is located at the foot of a relatively steep slope. At the

foot of the rise the relatively flat sediments are highly deformed by a series

irregular small scale slumps southeast of Site U1329. At the southeastern end

of this area, a long, deep fault surrounding the lifted slope can be seen (Fig.

4.37).

A few seismic lines cover this area (location see Fig. 4.37). The SCS

profile CAS05C_line03 (oriented northwest-southeast) shows a slight

topographic change from a plateau in the south-east to a deeper basin part on

the north-west end. A BSR was observed in the seismic profile and can be

traced for ~2.5km in line CAS05C_03 (Fig. 4.38).

The MCS line 89-08 shows a clear change in reflection character at about

CDP point 3500 at the foot of the steep rise towards the NE (Fig. 4.39). At the

foot of the rise, a small basin with well bedded sediment-fill can be seen and a

BSR can be identified at the SW edge of that basin. However, at Site U1329

no clear BSR can be seen. High-amplitude and low-frequency reflections

characterize the section underneath Site U1329 coincident with the lithologic

changes associated with unit II and III (see details below).

High frequency seismic line ODP1mc99 parallels seismic line 89-08 19km

to the NW. The different patterns of sedimentary strata can be clearly

65

distinguished within the basin, which is located west of the slope where Site

U1329 is located. Possible unconformity boundaries between the layers with

different trending are highlighted in figure 4.40. The BSR can still be

identified at the SW edge of the basin at ~1.91 s TWT from CDP 4200-4450.

Discontinuous BSRs are also found out around Site U1329, located at ~1.47 s

TWT from CDP6050-6350. Further upslope, deep seismic reflectivity can be

seen, marking a strong contrast to the sediment section underneath Site U1329,

which is almost barren of any reflectivity (Fig. 4.40).

4.4.2 Lithostratigraphy at U1329

Site U1329 is located near the foot of a relatively steep slope, and

sedimentation at this site is clearly dominated by slope processes. The

stratigraphy at Site U1329 can be divided into three lithostratigraphic units.

Unit I (0 – 37.0 mbsf) is characterized by fine grained detrital sediments (clay

and silty clay), locally interbedded with coarse grained sediments. The

sedimentation rate in Unit I appears to be relatively rapid at about 10-12 cm/ka.

The pore-fluid chlorinity data follow the seawater baseline of ~550 mM.

Lithostratigraphic Unit II (37.0 – 135.60 mbsf) is characterized by a high

abundance of biogenic silica (mainly diatoms). The sedimentation rate within

Unit II is between 4 to 10 cm/ka. Pore-fluid chlorinity data slowly decrease to

~525 mM between 80 mbsf to 124 mbsf. Correspondingly, log-resistivity data

show higher values at this depth interval. The depth of the base of the GHSZ is

estimated at 124 mbsf.

Lithostratigraphic Unit III (135.6 mbsf–TD) is characterized by fine

grained (clay to silty clay) detrital sediments with only a few coarser

interlayers from turbiditic deposits. The sediments in lithostratigraphic Unit III

were deposited at a low sedimentation of rate of ~0.8–2.8 cm/k.y. The bottom

of Unit III is characterized by extremely high resistivity and velocity data and

the pore-water chlorinity drastically decreases down to ~470 mM.

Lithostratigraphic Unit III is basically a conglomerate structure composed

of partly lithified and lithified rounded clasts supported by a silty clay matrix.

This feature is marked by the most prominent seismic reflector located at

~1.46 s TWT in the low-frequency seismic MCS line 89-08. This

66

conglomerate corresponds to an unconformity between upper Miocene and

Pleistocene sediments (no sediments preserved from 2.0 – 6.7 Ma); however,

the unconformity is not clearly seen in the higher-frequency SCS line

CAS05C_04.

4.4.3 Synthetic seismogram generation and Log‐to‐

seismic correlation

At Site U1329, the seafloor is at 959.6 m with a corresponding travel-

time of 1.27 s TWT. The total depth of the borehole is 1216.4 m, with a

predicted time of 1.58 s TWT (using a uniform velocity of 1656 m/s). A

complete time depth chart is then generated by integrating the WL log P-

wave velocity information (Table 4.5). The starting velocity is finally

adjusted to achieve the best result for the seafloor pick. The starting velocity

at the seafloor at Site U1325 is set to 1469.2 m/s.

An impedance log and reflection coefficients are generated from the

velocity and density log data. The velocity dataset at Site U1329 is generated

from the wire-line velocity log and LWD velocity log. The density data is

from the LWD at Hole U1329A. The reflection coefficients are then

convolved with a seismic wavelet to produce a synthetic seismic trace. At this

site, the seismic wavelet is obtained using a wavelet extraction from MCS line

89-08 data and SCS line CAS05C_line04 (Figure 4.41 and 4.44). The wavelet

is generated by choosing traces within a distance of 500m from Hole U1329A,

with a starting time of 1.25 s TWT to a time of 1.55 s TWT. A total of 41

traces were selected to generate the wavelet (Fig. 4.41a). The computed

spectrum shows a frequency range of 0 – 64 Hz. Two frequency peaks were

arrived at ~24 Hz and 36 Hz. The S/N ratio almost equals to 1 (Fig. 4.41b).

The synthetic seismogram is then compared with the actual seismic traces at

the drill site. Fig. 4.42 and Fig 4.45 illustrate the relationship between the

impedance logs, reflection coefficients, MCS and SCS traces, respectively,

and synthetic traces for Sites U1329.

67

The Synthetic seismogram for the MCS 89-08 data shows a good match

between synthetic and real seismic reflectors, especially at a depth of ~125

mbsf where BSR is located (Fig. 4.42).

A synthetic seismogram is also generated from seismic line

CAS05B_line03, Fig. 4.44 (a) and (b) show the extracted wavelet and

computed spectrum from seismic line CAS05B_line03 respectively. Fig. 4.45

illustrates the relationship between the impedance logs, reflection coefficients,

MCS traces, and synthetic traces for Sites U1329.

The synthetic seismogram combined with the SCS line and well logging

data is shown in Fig. 4.46. The synthetic trace generally matches the seismic

data well. The seafloor has presented strong positive reflectivity both in

synthetic and seismic data. The first 50 m do not show any indications of gas

hydrate appearance, according to the chlorinity data and the relatively weak

reflectivity in seismic and synthetic data. The BSR is located at ~1.466 s

TWT in the seismic below Site U1329, and the synthetic seismogram shows a

corresponding strong negative seismic reflectivity.

A synthetic seismogram has also been generated from seismic line

ODP1MCS99. Fig. 4.47 shows the extracted wavelet and computed spectrum,

respectively. Fig. 4.48 illustrates the relationship between the impedance

logs, reflection coefficients, MCS traces, and synthetic traces for Site U1329.

The synthetic seismogram including the well log information has been

superimposed on the seismic profile of line ODP1MCS99 (Fig. 4.49). A

relatively good match between synthetic and real seismic trace was achieved

at the BSR. Since most of the seismic data of line ODP1MCS99 between

seafloor and BSR is semi-transparent, it is not surprising to see little

correlation with the synthetic data.

68

Chapter 5 Summary, Uncertainties, and

Conclusions

5.1 Comparison of Results: Resistivity, Velocity and

Synthetic seismogram

5.1.1 Gas Hydrate Saturation Estimates

The different gas hydrate estimates at all Sites of IODP Expedition 311

have been calculated from different empirical relationship using electrical

resistivity logs and P-wave velocity data (from LWD and wire-lie logging).

Additionally, data derived from core pore-fluid chlorinity, and MCS velocities

were used to further constrain gas hydrate saturation estimates. Gas hydrate

saturation estimates from the resistivity and velocity log data are compared in

Fig. 5.1. In this comparison, the resistivity –based results shown are those

from LWD high resolution resistivity and density-porosity logs, and the

velocity-based results are those calculated from LWD velocity and density-

porosity logs.

Gas hydrate saturation estimates from the different methods are generally

in reasonably good agreement with each other. The gas hydrate saturation

estimates from LWD-derived data reveal numerous high-value intervals within

the stratigraphic section above the depth of the BSR. The results based on

velocity measurements generally give lower gas hydrate saturation estimates

than those calculated from resistivity logs. Gas hydrate saturations average

~5% ± 5% at Site U1326, ~10% ± 5% at Site 1325, and 5% at upper section

above 120mbsf and 15% ± 5% below 140mbsf of Site U1327 with the

maximum value reaching 40% at the interval of 120-140mbsf of Site U1327.

No significant gas hydrate is inferred at Site U1329 over the entire range of

the gas hydrate stability zone with the exception where potentially small

amounts may occur just above the BSR, as inferred from a slight freshening in

the chlorinity data.

69

The downhole logging–inferred gas hydrate occurrences appear to be

observed at a minimum depth of ~70–100 mbsf. At most sites we see evidence

for gas hydrates to about the base of the predicted methane hydrate stability

zone. Positive values of apparent estimates of hydrate below the BSR cannot

be related to hydrate, but possibly indicate the occurrence of free gas. The

downhole logging data from especially Sites U1327 and U1326 reveal

dominated gas hydrate reservoir by the appearance of thick high-resistivity

intervals at relatively shallow depths. Archie-based estimates yield high gas

hydrate saturation of ~>40% of the pore space in the intervals of Site U1326

(70-100mbsf) and U1327 (120-140mbsf). The LWD velocity estimates do not

yield equivalent high estimate in the 70-100mbsf section in Site U1326. The

gas hydrate occurrence at Site U1325, however, appears to be more evenly

distributed throughout the entire cored section.

The synthetic seismograms provide a chance to tie the seismic and well

logging data together. The synthetic seismograms match generally well at each

seismic survey, including the depth of the BSR and the gas hydrate enrichment

zones of Site U1326 and U1327.

5.1.2 Comparison with previous interpretation

The gas hydrate saturation estimates from this more comprehensive study

at Site U1327 differ significantly from the previous downhole resistivity study

at nearby Site 889 of Hyndman et al. (1999) and the velocity analysis at the

same study area by Goldberg et al. (2008). Hyndman et al. (1999) estimated a

gas hydrate saturation of 25%–30% within the 100 m interval just above the

BSR, compared to the 10%–20% estimated in this study. This discrepancy is

mainly attributed to the different choice of a pore water salinity baseline used

in this analysis, the new, higher-resolution porosity measurements, and

associated formulation of Archie's equation used to estimate the empirical

parameters.

In an earlier study, Chen et al (2008) calculated gas hydrate saturations

generated from LWD resistivity data with average values of ~9% ± 7% at Site

U1326, ~10% ± 7% at Site 1325, and 11% ± 7% at Site U1327 over the entire

70

range of gas hydrate occurrence, which is quiet close to the results of this

thesis.

The gas hydrate concentration estimates by Goldberg et al. (2008) using

the LWD velocity show quite different values, especially for Site U1326 over

the interval of 70-100mbsf. Goldberg et al. (2008) estimated gas hydrate

saturations reaching 50% of pore space or more. However, in this thesis, the

saturation estimate between these intervals only reaches 12% maximum at

depth 80mbsf and almost no gas hydrate appear within the rest of this gas

hydrate -rich interval, although identical LWD velocity data are used. These

results may be contributed to a different theoretical approach in modeling gas

hydrate within the sediment matrix. Goldberg et al. (2008) modeled gas

hydrate from a grain-cementation point of view (Dvorkin and Nur, 1993),

whereas the much lower estimates in this thesis were obtained from a time-

averaging approach using the Lee et al. (1996) weighted equation in which

gas hydrate is treated as pore-filling medium. However, a closer analysis of

the LWD velocity trend and theoretical approach to estimate gas hydrate

concentration is required, but is beyond the scope of this thesis.

5.2 Uncertainties

Many uncertainties in the estimation of gas hydrate saturation exist and

may be introduced by: (1) the data measurement methods (i.e. random noise in

the actual data sets used), (2) the assumed no-hydrate reference for various

physical properties, and (3) the way to approach the calculation of empirical

values needed in the analyses (e.g. Archie parameters a, m, and n, and the

weighting factor W).

The LWD measurements generally provide better quality data than the

WL measurements, due to lesser dependence on good contact between the

logging tool and the borehole wall. Especially due to the missing caliper logs

in most of the WL holes, it is difficult to identify the zones of poor borehole

condition and possible degraded measurements. However, the velocity is

generally in good agreement with the MCS velocities, which means that the

WL velocities are still reasonably reliable. Also, the differences in vertical

71

resolution between the different measurements also introduce some

uncertainties. The LWD velocity tool collects data every 0.5 meter, and the

LWD density tool collects data every 0.15 meter; when used together, the two

groups of data should be sampled at the same depth (only achieved every

1.5m). In that case, the vertical resolution decreases. Furthermore, inter-hole

variability can also be considered as a source of uncertainty.

Significant uncertainties arise in estimating the reference no-hydrate

profiles for resistivity and P-wave velocity, especially since the reference no-

hydrate resistivity profile is heavily dependent on estimating the in situ pore-

fluid salinity and formation temperature. A higher in situ salinity would result

in a lower reference no-hydrate resistivity, and in turn, larger than reference

resistivity values would result in higher gas hydrate concentrations. Also, the

empirical Archie-parameters depend on the no-hydrate reference profile.

Finally, the baseline of the chlorinity data can be difficult to define and is

dependent on core recovery and core quality. This is especially a problem at

Site U1327, which presents abnormally fresher baseline values throughout the

whole section drilled/cored. It was suggested (Chen et al., 2008) that the

decreased salinity results largely from a deeper fresh water source with the

remaining freshening being the result of dissociation of pervasive gas hydrate.

However, further evidence from direct pore water sampling in situ, which is

proposed through Osmo-sampling as part of the second phase of Expedition

311, is needed.

A similar issue in estimating a reliable reference profile exists for P-wave

velocity. In earlier approaches, the reference was defined from upward

projection of a deep-trendline defined by MCS interval velocities. In this

thesis a different approach was used by selecting a no-hydrate velocity

reference by defining the best-fit empirical W value in the Lee et al. (1996)

weighted equation to calculate velocities for those sediments that are hydrate-

free (mainly above 70 mbsf and below the BSR).

A different type of uncertainty arises from the general rock-physics model

assumed for estimating gas hydrate concentrations. In the Archie and time-

averaging analyses, the general assumption is that the gas hydrate is pore-

filling. The in situ pore water resistivity was calculated from the salinity

baseline and extrapolated temperatures from geothermal gradients of the sites.

72

Uncertainties in the density porosity measurement arise mainly from the

statistical uncertainty in the gamma ray count used to calculate the density and

the error induced by using an average grain density in mapping bulk density to

porosity. The way that Archie parameter n was chosen is reasonable in general

according to common literature. However, it is likely not applicable under the

situation when highly saturated gas hydrate or massive, grain-displacing gas

hydrate textures appear in the sediments (Spangenberg, 2001). For velocity-

based estimates, the way the empirical exponent r=1 (governing cementation)

is chosen gives reasonable results when compared to the resistivity-based

estimates. Lee et al., (1996) concluded that an exponent of r=1 adequately

describes the elastic behavior of the frozen sediment.

5.3 Conclusion

The combined observations from the transect sites of IODP Expedition

311 shows strong contradiction to the old model of gas hydrate formation that

was proposed for an accretionary margin by Hyndman and Davis (1992).

Their model has gas hydrate occurring mainly in sediments just above the

BSR. A good example that contradicts this observation is Site U1326 where

most of gas hydrate is present between 70-100mbsf, with local gas hydrate

concentrations as high as 60%. This interval of very high gas hydrate

saturation was interpreted to be a high-porosity sandy turbidite interval in

which large amounts of gas hydrate exists. The unexpected shallower

concentrated gas hydrate in the sediment as well as the lack of high gas

hydrate concentrations near the BSR indicate strong lithologic control on gas

hydrate formation, associated with sediment grain size and associated

formation parameters, such as porosity and permeability. Similarly, the highest

concentration at Site U1327 appear in a thick zone at ~120 mbsf with almost

no gas hydrate found just above the BSR. This high-concentration zone is

furthermore highly localized as it is laterally discontinuous (Riedel et al.,

2006a). The fact that no corresponding seismic reflection event was seen in

any of the seismic data around U1327 suggests that this high-concentration

layer is a local lens, maybe a remnant of some former continuous layer that

73

was heavily deformed and now has a wave-length smaller than the horizontal

resolution of the seismic data.

The model by Hyndman and Davis (1992) also has a strong component of

pervasive upward migration of fluids. However, the seismic data at each site

and the heterogeneous distribution of gas hydrate suggest that pore-fluids do

not move pervasively (i.e. homogenously upward in a 1-dimensional flow) but

rather get focused in fractures and faults. The best example for fault control in

hydrate formation can be seen at Site U1326 and U1325.

In summary, a revised model of gas hydrate formation should be

formulated. However, without detailed pore-water geochemical modeling that

describes the type and source of fluid flow (especially to estimate flow rates)

the model is incomplete. But this thesis provides the structural, large-scale

framework for such a new model and thus is a cornerstone in the evolving

process to find a new, general fluid-flow model for gas hydrate formation in

accretionary prisms. The main contributions to this “global” model from this

thesis are:

(1) Gas hydrate does not form uniformly throughout the gas hydrate

stability zone, but is rather opportunistic in that it concentrates

where lithologic conditions are met that favor gas hydrate

precipitation (i.e. turbidite layers);

(2) Fluid flow is heavily overprinted by fault and fracture control that

focus the pervasive fluid flow into preferred pathways;

(3) The BSR is mainly an effect of small concentrations of free gas

below the base of gas hydrate stability as almost no large

accumulation of gas hydrate just above the BSR was identified

regionally (with an exception of maybe Site U1329);

(4) Careful time-depth conversion is required to match log- and

seismic data to identify zones of gas hydrate accumulation on

seismic images;

(5) Gas hydrate concentrations are relatively low on a regional scale,

but can locally exceed 50% of the pore-volume.

74

References:

Archie, G.E., 1942, the electrical resistivity log as an aid in determining some

reservoir characteristics. Transaction of the Mining and Metallurgical Engineers, 146,

p. 54-62

Bangs, N.L., Sawyer, D.S., and Golovchenko, X., 1993, Free gas at the base of the

gas hydrate zone in the vicinity of the Chile triple junction, Geology, 21, p. 905-908.

Buffett, B.A., 2000, Clathrate Hydrates, Ann. Rev. Earth Planet. Sci, p. 477-507.

Chapman, N.R., Gettrust, J., Walia, R., Hannay, D. Spence, G.D., Wood, W., and

Hyndman, R.D., 2002. High resolution deep-towed multichannel seismic survey of

deep sea gas hydrates off western Canada. Geophysics, Vol. 67, p. 1038-1047.

Chen, M. P., Riedel, M., Spence, G. D., and Hyndman, R. D., 2008, Data report: a

downhole electrical resistivity study of northern Cascadia marine gas hydrate,

Proceedings of the Integrated Ocean drilling Program, Vol. 311,

doi:10.2204/iodp.proc.311.203.

Chen, M. P., Riedel, M., Hyndman, R. D., and Dosso, S. E., 2007, AVO inversion of

BSRs in marine gas hydrate studies, Geophysics, 72 No. 2, doi:10.1190/1.2435604.

Chen, M. P., 2006, Northern Cascadia Marine Gas Hydrate: Constraints from

Resistivity, Velocity, and AVO, M.Sc. Thesis, School of Earth and Ocean Sciences,

University of Victoria.

Claypool, G. E., Threlkeld, C. N., Mankiewicz, P. N., Arthur, M. A., Anderson, T. F.,

1985, Isotopic composition of interstitial fluids and origin of methane in slope

sediment of the Middle America trench, Deep-Sea Drilling Project Leg-84, Initial

Reports of the Deep Sea Drilling Project 84, p. 683-691.

Clennell, M.B., Hovland, M., Booth, J.S., Henry, P., and Winters, W.J., 1999

Formation of natural gas hydrate in marine sediments 1. Conceptual model of gas

hydrate growth conditioned by host sediment properties, J. Geophys. Res., 104 (B10),

p. 22,985-23,003.

75

Collett, T.S., Agena, W.F., Lee, M.W., Zyrianova, M.V., Bird, K.J., Charpentier, T.C.,

Houseknect, D.W., Klett, T.R., Pollastro, R.M., and Schenk, C.J., 2008, Assessment

of gas hydrate resources on the North Slope, Alaska, 2008: U.S. Geological Survey

Fact Sheet 2008-3073, p. 4.

Collett, T.S., 2000. Quantitative well-log analysis of in-situ natural gas hydrates

[Ph.D. dissert.]. Colorado School of Mines, Golden, CO.

Dash, R.K., 2008, Crustal Structure and marine gas hydrate studies near Vancouver

Island using seismic tomography, [PhD. Thesis]: Victoria, University of Victoria, p.

171.

Goldberg, D., Guerin, G., Malinverno, A., and Cook, A. 2008, Velocity Analysis of

LWD and Wireline Sonic Data in Hydrate-bearing sediments on the Cascadia Margin,

Proc. Of the 6th International Conference on Gas Hydrate (ICGH 2009), Lamont-

Doherty Earth Observatory, Columbia University, Palisades, NY.

Davis, E.E., Becker, K., Wang, K., and Carson, B., 1995, Long-term observations of

pressure and temperature in Hole 892B, Cascadia Accretionary Prism, in: Carson, B.,

Westbrook, G.K., Musgrave, R.J., and Suess, E. (Eds), 1995: Proceedings of the

Ocean Drilling Program, Scientific Results, Vol. 146 (Pt 1).

Davis, E.E., Hyndman, R.D., and Villinger, H., 1990, Rates of fluid expulsion across

the northern Cascadia accretionary prism: constraints from new heat flow and

multichannel seismic reflection data. J. Geophys. Res., 95 (B6), p. 8869-8889.

Davis E.E., and Hyndman, R. D., 1989, Accretion and recent deformation of

sediments along the northern Cascadia subduction zone, Geol. Soc. Am. Bull., Vol.

101, p.1465-1480.

Dix, C.H., 1955, Seismic velocities from surface measurements, Geophysics, 20, p.

68-86.

Dvorkin J., and Nur, A., 1996, Elasticity of high-porosity sandstones: Theory for two

North Sea Datasets, Geophysics, 61, p. 1363-1370.

76

Dvorkin, J., and Nur, A., 1993, Rock physics for characterization of gas hydrates, in

Howell, D.G., ed., The Future of Energy Gases. U.S. Geological Survey Professional

Paper, 1570, p. 293-311.

Dvorkin, J., Mavko, G., and Nur, A. 1991, The effect of cementation on the elastic

properties of granular material. Mechanical Materials, Vol. 12, p. 207-217.

Fink, C. R., and Spence, G. D., 1999, Hydrate distribution of Vancouver Island from

multifrequency single-channel seismic reflection data, J. Geophys. Res., 104, 2909-

2922.

Fofonoff, N. P., 1985, Physical Properties of Seawater: A New Salinity Scale and

Equation of State of Seawater, Journal of Geophysical Research, Vol. 90, No. C2, p.

3332-3342.

Ganguly, N., Spence, G.D. , Chapman, N.R., and Hyndman, R.D., 2000, Heat flow

variations from bottom simulating reflectors on the Cascadia margin, Marine Geology,

164, p.53-68.

Gettrust, J., F., Chapman, R., Walia, R., Wood, W., Hannay, D., Lindwall, D., Spence,

G. D., Louden, K., and Hyndman, R. D., 1999, High resolution seismic studies of

deep sea gas hydrate using the DTAGS deep towed multichannel system. EOS,

Transactions, American Geophysical Union, Vol.80, No.38, p.439-440.

Gettrust, J. F. and Ross, J. H., 1990, Development of a low-frequency, deep-towed

geoacoustics system, Proc. Oceans 90, p. 38-40.

Guerin G., Goldberg, D., and Meltser, A., 1999, Characterization of in situ elastic

properties of gas hydrate-bearing sediments on the Blake Ridge, J. Geophy. Res., Vol.

104, p.17781-17795.

Haacke, R., Hamilton, T., Enkin, R., Esteban, L., and Pohlman, J, 2008, Pacific

Geoscience Centre marine gas hydrates cruise to the frontal ridge and mid-slope areas

of the Vancouver Island accretionary wedge, PGC2008007- cruise report, geological

Survey of Canada.

77

Helgerud, M.B., Dvorkin, J., and Nur, A., 2000, Rock physics characterization for gas

hydrate reservoirs, elastic properties. In Holfert, G.D., and Bishnoi, P.R., eds., Gas

Hydrates, Challenges for the Future, Annals of the New York Academy of Sciences,

Vol. 912, p. 116-215.

Hobro, J. W. D., Minshull, T. A., Singh, S. C. and Chand, S., 2005, A three-

dimensional seismic tomography study of the gas hydrate stability zone, offshore

Vancouver Island. J. Geophy. Res., 110, B09102, doi: 10.1029/2004JB003477.

Hyndman R.D. and Davis, E. E., 1992, A mechanism for the formation of methane

hydrate and seafloor bottom-simulating reflectors by vertical fluid expulsion, J.

Geophys. Res., 97, p.7025-7041.

Hyndman, R.D., Spence, G.D., Yuan, T., and Davis, E.E., 1994. Regional geophysics and structural framework of the Vancouver Island margin accretionary prism. In Westbrook, G.K., Carson, B., Musgrave, R.J., et al., Proc. ODP, Init. Repts., 146 (Pt. 1): College Station, TX (Ocean Drilling Program), 399–419.

Hyndman, R.D., Yuan, T., and Moran, K., 1999, The concentration of deep sea gas

hydrates from downhole electrical resistivity logs and laboratory data, Earth Planet.

Sci. Lett., Vol. 172(1–2), p.167–177. doi: 10.1016/S0012-821X(99)00192-2

Hyndman, R.D., G.D., Chapman, R., Riedel, M., and Edwards, R.N., 2001,

Geophysical Studies of Marine Gas Hydrate in Northern Cascadia, AGU Geophys.

Monograph 124, Natural Gas Hydrates: Occurrence, Distribution, and Detection,

p.273-295.

Klauda, J. B., and Sandler, S. I., 2005, Global Distribution of Methane Hydrate in

Ocean Sediment, Energy & Fuels, Vol. 19, p. 459-470.

Kuster, G.T., and Toksőz, M.N., 1974, Velocity and attenuation of seismic waves in

two-phase media, 1, theoretical formulation. Geophysics, Vol. 39, p. 587-606.

Kvenvolden, K.A., and Barnard, L.A., 1983. Hydrates of natural gas in continental

margins, In Watkins, J.S., and Drake, C.L. (Eds.), Studies in Continental Margin

Geology. AAPG, p.631-640.

78

Kvenvolden K. A. and Kastner M., 1990, Gas hydrates of the Peruvian outer

continental margin, In: Suess, E., and von Huene, R. e. al. (Eds.), Proceedings of the

Ocean Drilling Program, Scientific Results, Vol. 112. p. 517-526.

Kvenvolden, K. A., 1993, Gas hydrates - geological perspective and global change,

Reviews of Geophysics, Vol.31, p.173-187.

Lee, M.W., Hutchinson, D.R., Dillon, W.P., Miller, J.J., Agena, W.F., and Swift, B.A.,

1993, Method of estimating the amount of in situ gas hydrates in deep marine

sediments, Marine and Petroleum Geology, Vol.10, p. 493-506.

Lee, M.W., Hutchinson, D.R., Collett, T.S., and Dillon, W.P., 1996. Seismic

velocities for hydrate-bearing sediments using weighted equation. J. Geophy. Res.,

101:20347–20358.

Lopez C. D., Spence, G. D., Hyndman, R.D., and Kelley, D., 2008, Frontal ridge

slope failure at the northern Cascadia margin: margin-normal fault and gas hydrate

control; in preparation for Geology.

Lopez, C. D., 2008, Seismic velocity structure associated with gas hydrate at the

frontal of northern Cascadia Margin [MSc. Thesis]: Victoria, University of Victoria,

p.145.

MacKay, M.E., Moore, G.F., Cochrane, G.R., Moore, J.C., and Kulm, L.D.,1992,

Landward vergence and oblique structural trends in the Oregon Margin accretionary

prism: implications and effect on fluid flow. Earth Planet. Sci. Lett., Vol.109, p.477-

491.

MacKay, M.E., Jarrard, R.D., Westbrook, G.K., Hyndman, R.D., and the Shipboard

Scientific Party of ODP Leg 146, 1994, Origin of bottom simulating reflectors:

Geophysical evidence from the Cascadia accretionary prism, Geology, Vol. 22, p.

459-462.

Malinverno, A., Kastner, M., Torres, M. E. and Wortmann, U.G., 2008, Gas hydrate

occurrence from pore water chlorinity and downhole logs in a transect across the

northern Cascadia margin (Integrated ocean Drilling Program Expedition 311), J.

Geophys. Res. Vol. 113, B08103, doi: 10.1029/2008JB005702.

79

Matthew K. D., and Bruce A. B., 2001, Rock the Foundation Convention, Canadian

Society of Petroleum Geologists.

Mi, Y., 1998, Seafloor sediment coring and multichannel seismic studies of gas

hydrate, offshore Vancouver Island, MSc thesis, University of Victoria.

Milkov A.V., 2004. Global Estimates of Hydrate-Bound Gas in Marine Sediments:

How Much Is Really out There? Earth-Science Reviews Vol. 66 (2004), p.183–197.

Miller, J.J., Lee, M.W., and von Huene, R., 1991, An analysis of a seismic reflection

from the base of a gas hydrate zone, offshore Peru, AAPG Bull., Vol.75, p. 910-924.

Minshull, T.A., Singh, S.C., and Westbrook, G.K., 1994, Seismic velocity structure at

a gas hydrate reactor, offshore western Columbia, from full waveform inversion, J.

Geophys. Res., 99, B3, p.4715-4734.

National Petroleum Council, 2007. Global Oil & Gas Study, Topic Paper #24

Hydrates.

Nicholson, T., Michael G., and Cassidy, J. F., 2003, Structure of the northern

Cascadia subduction zone, Geological Society of America Abstracts with Programs,

Vol. 35(6), p. 308.

Novosel, I., 2005, Physical properties of gas hydrate related sediments offshore

Vancouver Island [M.Sc. thesis]: Victoria, Canada, University of Victoria, p.114.

Paull, C.K., Ussler III, W.U., and Dillon, W. P., 2003, Continental Margin gas

hydrate: Slope stability, geo-hazard, and climate? Geological Society of America

Abstracts with Programs, Vol. 35(6), p. 532.

Paull, C.K., Ussler III, W.U., and Borowski, W.S., 1994, Source of methane to form

marine gas hydrates, in Natural Gas Hydrates (Ed.) E.D. Sloan, J. Happel Jr., M.A.,

Hnatow, 161-176, New York, N.Y. Acad. Sci.

Pearson, C.F., Halleck, P.M., McGuire, P.L., Hermes, R., and Mathews, M., 1983,

Natural gas hydrate deposit: a review of In-situ properties, Journal of Physical

Chemistry, Vol. 87, p.4, 180-4,185.

80

Pohlman, J.W., Kaneko, M., Heuer. V., Plummer, R., Coffin, R.B., and the IODP

Expedition 311 Scientific Party, 2006, Molecular and isotopic characterization of

gases from IODP Expedition 311: source and gas hydrate related controls, Eos, Trans.

Am. Geophys. Union, Vol. 87(52) (Suppl.) : 0S11E-06.

Riedel, M., Collett, T.S., Malone, M.J., and the Expedition 311 Scientists, 2006a,

Proceedings of the Integrated Ocean Drilling Program, Volume 311: Washington, DC

(Integrated Ocean Drilling Program Management International, Inc.).

doi:10.2204/iodp.proc.311.101.2006

Riedel, M., Novosel, I., Spence, G.D., Hyndman, R.D., Chapman, N.R., and Lewis,

T., 2006b, Geophysical and geochemical signatures associated with gas hydrate–

related venting in the northern Cascadia margin. Geol. Soc. Am. Bull., Vol.118, p.23–

38. doi:10.1130/B25720.1.

Riedel, M., Collett, T.S., and Hyndman, R.D., 2005, Gas hydrate concentration

estimates from chlorinity, electrical resistivity, and seismic velocity, Geol. Surv. Can.

Open-File Rep., p.4934.

Riedel, M., Collett, T.S., Malone, M.J., and the Expedition 311 Scientists, 2005,

Integrated Ocean Drilling Program Expedition 311 Preliminary Report, Cascadia

Margin Gas Hydrates, p. 73.

Riedel, M., Hyndman, R.D., Spence, G.D., and Chapman, N.R., 2002, Seismic

Investigations of a Vent Field Associated with Gas Hydrates, Offshore Vancouver

Island, J. of Geophys. Res., JGR Solid Earth, Vol 107(B9), p.2200, doi:10.1029/2001.

Riedel, M., 2001, 3-D Seismic Investigations of Northern Cascadia Marine Gas

Hydrates, University of Victoria, BC, [PhD thesis], p.9, 35-36.

Satake, K, Shimazaki, K., Tsuji, Y., and Ueda, K., 1996, Time and size of a giant

earthquake in Cascadia inferred from Japanese tsunami records of January 1700,

Nature, Vol. 379, p.246 - 249; doi:10.1038/379246a0.

Schlumberger Educational Services, 1989. Log Interpretation Principles/Applications.

Schlumberger Educational Services Publishers, Houston, Texas, p.247.

81

Schwalenberg, K., Willoughby, E. C., Mir, R., and Edwards. R. N., 2005, Marine gas

hydrate signatures in Cascadia and their correlation with seismic blank zones, First

Break, Vol. 23, p. 57-63.

Serra, O., 1984, Fundamentals of Well-Log Interpretation, The Acquisition of

Logging Data, Dev. Pet. Sci., Vol. 1, 15A: Amsterdam (Elsevier).

Shipley, T.H., Houston, R.T., Buffler, R.T., Shaub, F. J., McMillen, K.J., Ladd, J.W.,

and Worzel, J.L., 1979, Seismic evidence for widespread possible gas hydrate

horizons on continental slopes and rises, AAPG Bull., Vol.63, p.2204-2213.

Shipley, T.H., M.H. Houston, R.T. Buffler, F.J. Shaub, K.J. McMillen, J.W.

Sloan, E.D., Clathrate Hydrates of Natural Gas, 2nd ed., Marcel Dekker, New York ,

NY, (1997). Pg 41- 47

Solem, R.C., Spence, G.D., Vukajlovich, D., Hyndman, R.D., Riedel, M., Novosel, I.,

and Kastner, M., 2002. Methane advection and gas hydrate formation within an active

vent field offshore Vancouver Island. Fourth Int. Conf. Gas Hydrates, 19023.

Spangenberg, E., 2001, Modeling of the influence of gas hydrate content on the

electrical properties of porous sediments, J. Geophys. Res., 106(B4), 6535–6548.

Spence, G., He, T., Dash, R., Lopez, C., Hyndman, R., 2008, Seismic structure, gas

hydrate concentrations, and slumping along the IODP X311 transect on the N.

Cascadia margin; Proceedings of the 6th International Conference on Gas Hydrates,

Vancouver, paper 4-C6.

Spence, G.D., Hyndman, R.D. Davis, E.E., and Yorath, C.J., 1991a. Marine

multichannel seismic reflection data across the continental margin of Vancouver

Island, Geological Survey of Canada. Open File Report, 2391.

Spence, G.D., Hyndman, R.D., Davis, E.E., and Yorath, C.J., 1991b. Seismic

Structure of the northern Cascadia accretionary prism: Evidence from new

multichannel seismic reflection data, in R. Meissner, L. Brown, Durbaum, H. –J.,

Franks, W., Fuchs, K., and Seifert, F., eds., Continental Lithosphere: Deep seismic

Reflections, Geodynamics Series, Vol. 22, p.257-263.

82

Spence, G.D., Minshull, T.A., and Fink, C., 1995, Seismic structure of methane gas

hydrate, offshore Vancouver Island, Proc. Ocean Drill. Program, Sci. Results 146,

p.163-174.

Timothy S. C., 2000, Quantitative well-log analysis of in-situ natural gas hydrates

[PhD Thesis], Colorado, Colorado School of Mines, p.115.

Timur, A. 1968, Velocity of compressional waves in porous media at permafrost

temperature, Geophysics, Vol. 33, p.584-595.

Toksőz, M.N, Cheng, C.H., and Timur, A., 1976, Velocity of seismic waves in porous

rocks, Geophysics, Vol. 41, p. 621-645.

Torres, M. E., Tréhu, A. M., Cespedes, N., Kastner, M., Wortmann, U. G., Kim, J.-H.,

Long, P. E., Malinverno, A., Pohlman, J. W., Riedel, M., and Collett, T. S., 2008,

Methane hydrate formation in turbidite sediments of northern Cascadia, IODP

Expedition 311, Earth Planet. Sci. Lett., Vol. 271, p.170– 180,

doi:10.1016/j.epsl.2008.03.061.

Tosaya, C., and Nur, A. 1982, Effects of diagenesis and clays on compressional

velocities in rock. Geophysical Research Letters, Vol.9, p. 5-8.

Tréhu, A. M., et al., 2004, Three-dimensional distribution of gas hydrate beneath

southern Hydrate Ridge: constraints from ODP Leg 204, Earth Planet. Sci. Lett., Vol.

222, p.845– 862.

Tréhu, A.M, Bohrmann, G., Rack, F.R., Torres, M.E., et al., 2003, Proc. ODP, Init.

Repts. 204: College Station, TX (Ocean Drilling Program). doi:

10.2973/odp.proc.ir.204.2003

Trofimuk, A. A., Cherskiy, N. V., and Tsarev. V. P., 1973, Accumulation of natural

gases in zones of hydrate — formation in the hydrosphere (in Russian), Doklady

Akademii Nauk SSSR, Vol. 212, p. 931-934.

Trofimuk, A. A., Cherskiy, N. V., and Tsarev. V. P., 1977, The role of continental

glaciation and hydrate formation on petroleum occurrences, In Meyer, R. F. (Ed.),

Future supply of nature-made petroleum and gas, NY: Pergamon Press, p. 919-26.

83

Ussler III, W., and Paull, C. K., 2001, Ion exclusion associated with marine gas

hydrate deposits, in Natural Gas Hydrates: Occurrence, Distribution, and Detection,

Geophys. Monogr. Ser., Vol. 124, Paull, C. K., and Dillon, W. P., (Ed.), p. 41– 51,

AGU, Washington, D. C.

Von Huene, R., and I. Pecher, 1999, Vertical tectonics and the origins of BSRs along

the Peru margin, Earth Planet. Sci. Lett., Vol.166, p.47-55.

Warner, M., Absolute reflections from deep seismic reflections, Tectonophys. 173, 15-23, 1990.

Westbrook, G. K., Carson, B., Musgrave, R. J., et al., 1994, Proceedings of the

Ocean Drilling Program, Initial Reports, Vol. 146 (Part 1), p.5.

Westbrook, G. K., Carson. B., Musgrave, R.J., et al., 1994, Proceedings of the Ocean

Drilling Program, Initial reports, Vol. 146 (Part 1)

Whalley, E., 1980, Speed of longitudinal sound in clathrate hydrates, J.of Geophys.

Res., Vol. 85(B5), p. 2,539-2,542.

Whiffen, B.L., Kiefte, H., and Clouter, M.J. 1982, Determination of acoustic

velocities in xenon and methane hydrates by Brillouin spectroscopy, Geophysical

Research Letters, Vol.9(6), p. 645-648.

Whiticar M. J., 1999, Carbon and hydrogen isotope systematics of bacterial formation

and oxidation of methane, Chem. Geol, Vol. 161, p.291-314.

Whiticar M. J., Hovland M., Kastner M., Sample J. C., 1995, Organic geochemistry

of gases, fluids, and hydrates at the Cascadia accretionary margin, In: Carson, B.,

Westbrook, G. K., Musgrave, R. J. et al. (Eds.). Proceedings of the Ocean Drilling

Program, Scientific Results. p. 385-397.

Willoughby, E. C., Latychev, K., Edwards, R.N., and Mihajlovic, G., 2000,

Resource Evolution of Marine Gas Hydrate Deposits Using Seafloor Compliance

Methods, Gas Hydrates: Challenges for the Future, Ann. N.Y. Acad. Sci., 912.

Winsauer, W.O., and Shearin, H.M., Jr., 1952, Resistivity of brine-saturated sands in

relation to pore geometry, AAPG Bull., Vol.36, p. 253–277. doi:10.1306/3D9343F4-

16B1- 11D7-8645000102C1865D

84

Wood, A.B., 1941, A text book of sound, MacMillan Publishing, New York, p. 578.

Wood, W.T., Stoffa, P. L., and Shipley, T.H., 1994, Quantitative detection of methane

hydrate through high-resolution seismic velocity analysis, J. Geophys. Res., Vol. 99,

p.9681-9695.

Xu, W. and Ruppel, C., 1999, Predicting the occurrence, distribution and evolution of

methane gas hydrates in porous sediments. Journal of Geophysical Research, v. 104,

p. 5081-5096.

Yuan, J., and Edwards, R. N., 2000, Towed seafloor electromagnetics and assessment

of gas hydrate deposits, Geophysical Res. Lett., Vol.27, p.2397-2400.

Yuan, T., Spence, G.D., Hyndman, R.D., Minshull, T.A., and Singh, S.C., 1999,

Seismic velocity studies of a gas hydrate bottom-simulating reflector on the northern

Cascadia continental margin, Amplitude modeling and full waveform inversion, J.

Geophys. Res., Vol.104, p.1179-1191.

Yuan, T., Hyndman, R.D., Spence, G.D. and Desmos, B., 1996, Seismic velocity

increase and deep-sea gas hydrate concentration above a bottom-simulating reflector

on the northern Cascadia continental slope, J. Geophys. Res., Vol. 101, p.13665-

13671.

Zühlsdorff, L., V. Spiess, C. Huebscher, H. Villinger, and A. Rosenberger, 2001,

Implications for focused fluid transport at the northern Cascadia accretionary prism

from a correlation between BSR occurrence and near seafloor reflectivity anomalies

imaged in a multi-frequency data set, Geologische Rundschau, 88.

Zykov, M.M., and Chapman, N.R., 2004, 3-D velocity model of hydrocarbon vent

site in Cascadia region offshore Vancouver Island, AAPG Hedberg conference “Gas

hydrate; Energy resource potential and associated geologic hazards”(2005),

Vancouver, BC, Canada.

85

Figures

(a)

(b)

86

(c)

(d)

87

Fig. 1.1 a) Northern Cascadia Margin tectonic setting and location of IODP

Expedition 311 (red transect line) (from IODP Expedition 311 preliminary

reports, 2005). The arrows in the inset chart show the direction of overall plate

motion. The gas hydrate occurrence zone is indicated with gray shading area

on the midcontinental slope, where wide-spread BSRs can be detected. Four

drilling sites of the IODP Expedition 311 are distributed around the transect

line. The water depth at the Cascadia basin reaches ~2600m, and the water

becomes shallower landward, as the sediment section grows rapidly in

thickness due to tectonic shortening.

b, c) Bathymetry and location maps for IODP Expedition 311 Sites U1325,

U1326, U1327, U1329. The white line marks the transect MCS line 89-08,

which goes across the Cascadia accretionary wedge drilled during Expedition

311 (bathymetry courtesy of D. Kelley, J. Delaney, and D. Glickson,

University of Washington, and C. Barnes, C. Katnick, NEPTUNE Canada,

University of Victoria; funded by the University of Washington and the W.M.

Keck Foundation)

88

89

Fig. 1.2 Seismic transect line (MCS line 89-08) across the northern Cascadia

margin showing the location of the four IODP Expedition 311 drilling sites

visited (Site U1326, U1325, U1327 and U1329). The cold vent Site U1328 is

3 km off the line to the southeast. The Cascadia basin with turbidites and

hemipelagic sediments has been marked out at the deformation front. The

oceanic crust can be distinguished below. The location of the BSR is also

indicated on the image. The location of the four drilling sites has been marked

on the top with the relative offsets between each other.

90

Dep

th(k

m)

Deep sea gas hydrate stability

Gas Hydrate Phase Boundary

Water temperature

Ocean

Seafloor

Geothermal Gradient

BSR

Hydrated sediment

free gas zonewater saturated sediment

1

2

3

0 10 20 30 40

Temperature (degree Celsius)

Figure 1.3 Generic Methane Hydrate Stability Curve. The “Phase boundary”

line (a pressure/temperature line) divides the conditions under which methane

and water would result in hydrate on the left side and those where methane

would remain as free gas on the right side. The “geothermal gradient” line

indicates the actual temperature of the sediments, which has a linear

relationship with depth below sea floor. Below the zone, free gas can exist.

91

Fig. 1.4 Locations where natural gas hydrate has been recovered or is inferred (modified from Kvenvolden and Rogers, 1993).

92

Figure 1.5 Estimates of global gross volume of hydrate-bound gas (solid dots)

(modified from Milkov, 2004), including the most recent estimate (open circle)

(Klauda and Sandler, 2005). The red line indicates the USGS estimate

(Kvenvolden, 1993), while the orange line indicates global conventional gas

reservoirs. The unit of the vertical axis is by 15 310 m (From National petroleum

council, 2007).

93

Fig.1.6 Gas flare profile snapshot near IODP Site U1327. The data was

collected by echo-sounders of 18 kHz frequencies. The color bar stand for

amplitude, and vertical scale is meter, the horizontal scale is in nautical miles

(nmi). The red flat horizontal layer indicates the seafloor which is located at

~1300 meters below sea level.

94

Fig. 1.7 Model for upward methane migration and gas hydrate formation in

subduction zone accretionary prisms such as off Vancouver Island (From

Hyndman and Davis, 1992)

95

Fig. 2.1 Schematic diagram showing marine gas hydrate studies (From Hyndman et al. 2001)

96

Fig. 2.2 Single channel seismic track lines of the 1999 and 2000 surveys

across known cold vents. The position of inline 27 from the main 3D grid

serves as a reference. Positions of the five OBSs from the 1999 deployment

are shown as circles. OBSs are labeled from A-F. Selected piston-core

locations (dark squares) at blank zone 1 are also shown (From Riedel, 2001).

97

(a) (b)

Fig. 2.3 (a) Location map of Leg 146 drilling sites off Vancouver Island and Oregon. (b) Contour map with positions of drilling Sites 888-890,

along with the positions of drilling sites U1325-U1329 of Expedition 311. The two MCS transect lines 89-08 and 89-04 are also shown on the

map. The depth-contour lines are each 100 meter apart (modified after Westbrook, G. K. et al., 1994).

98

Fig 3.1 Downhole LWD measurements with resistivity logs. Solid line = seismically inferred bottom-simulating reflector (BSR). (From Chen et

al., 2006)

99

(a)

15 20 25 30 35 40

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Interstitial water salinity

U1325

(b)

0 0.1 0.2 0.3 0.4

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Interstitial water resistivity(ohm-m)

U1325

Fig 3.2 (a) Water salinity (in ppt) from recovered sediment at Hole U1325B, U1325C and U1325D (b) estimated in situ water resistivity

calculated from the equation of state of sea water (Fofonoff, 1985).

100

(a) 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.90

2

4

6

8

10U1325 Pickett Plot

LWD density-Porosity

FF(R

t/Rw

)

m=0.26, a=4.65m=1.76, a=1.17

(b)0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4

6

8

10U1325 Pickett Plot

core-derived porosity

FF(R

t/Rw

)

m=0.44, a=3.76m=1.76, a=1.64

Fig. 3.3 (a) Pickett Plot in Site U1325, the blue scattered points are the density-porosity values plotted against the FF data. A red power line is

then fitted through the scattered points, the empirical parameters from the equation of the power line (as showed at bottom of the chart) is a=3.48,

m=0.43, and the blue line stands for the empirical relationship calculated also from LWD density-porosity data, with m fixing to 1.76, parameter

a from empirical fitting equals to 1.17. (b) Pickett plot in Site U 1325 using the core derived density porosity values plotted against the FF data.

A power line is fitted through the scattered point, the empirical parameters is a= 3.76, m=0.44. Blue line is the power relationship between core-

derived porosity and FF with m=1.76, a=1.64.

101

0 0.1 0.2 0.3 0.4 0.5 0.6

0

50

100

150

200

250

300

350

400

Dep

th b

elow

sea

floor

(m)

Gas Hydrate Saturation

U1325

Sh (density-porosity, a=1.17, m=1.76)Sh (core-derived porosity, a=1.64,m=1.76)

0 0.1 0.2 0.3 0.4 0.5 0.6

0

50

100

150

200

250

300

350

400

Dep

th b

elow

sea

floor

(m)

Gas Hydrate Saturation

U1325

Sh (m=0.43, a=3.48)Sh (m=1.76, a=1.17)

Fig. 3.4 (a) the red line is Site U 1325 gas hydrates saturation (Sh) calculated

from LWD log-derived density-porosity data, the empirical value are a =1.17,

m=1.76, the blue line is Sh calculated from core-derived porosity with a= 1.64,

m=1.75. (b) The red line stands for the Sh calculated from LWD log-derived

density-porosity data, the empirical values are a =1.17, m=1.76, yellow line is

Sh calculated from LWD log-derived density-porosity data with m=0.43,

a=3.48. In all the calculations, the value n is chosen as 1.95.

102

(a)

5 10 15 20 25 30 35

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Interstitial water salinity

U1326

(b)

0.15 0.2 0.25 0.3 0.35

0

50

100

150

200

250

Dep

th b

elow

sea

floor

(m)

Interstitial water resistivity(ohm-m)

U1326

Fig 3.5 (a) Water salinity (in ppt) from recovered sediment at Hole U1326C and U1326D (b) estimated in situ water resistivity calculated from

the equation of state of sea water (Fofonoff, 1985).

103

(a)0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4

6

8

10U1326 Pickett Plot

LWD density porosity

FF(R

t/Rw

)

m=0.26, a=4.65m=1.76, a=1.57

(b) 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4

6

8

10U1326 Pickett Plot

core-derived Porosity

FF(R

t/Rw

)

m=0.23, a=4.76m=1.76, a=2.3

Fig. 3.6 (a) “Pickett Plot” in Site U1326, the blue scattered points are the density-porosity values plotted against the FF data. A power line is then

fit through the scattered points, the empirical parameters for the equation of the power line (as showed at bottom of the chart) are a=4.65,

m=0.27. The blue line stands for the empirical relationship calculated from LWD density-porosity data, with m fixing to 1.76, parameter a from

empirical fitting equals to 1.57. (b) Pickett Plot shows the results of core-derived porosity plotting against the FF data in Site U1326. A power

line is then fit through the scattered point, the empirical parameters from the equation of the power line (as showed at bottom of the chart) is

a=4.67, m=0.23. The blue line is the empirical relationship between core-derived porosity and FF from core-derived porosity with m=1.76, a=2.3.

104

(a)

0 0.2 0.4 0.6 0.8

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Gas Hydrate Saturation

U1326

Sh (density-porosity, m=1.76, a=1.57)Sh (core-derived porosity, m=1.76,a=2.3)

(b)

0 0.2 0.4 0.6 0.8

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Gas Hydrate Saturation

U1326

Sh (density-porosity, m=0.27,a=4.65)Sh (density-porosity, m=1.76, a=1.57)

Fig. 3.7 Gas hydrate saturation from density-porosity and core-derived porosity calculation; the yellow line stands for gas hydrate saturation

using the empirical value 4.65 and 0.27 for a and m respectively, the red line (in both (a) and (b)) stands for the gas hydrate saturation from

density-porosity using m=1.76, a =1.57. The blue line is the result from core-derived porosity, using m=1.76, a=2.3.

105

(a)

0 10 20 30 40

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Interstitial water salinity

U1327

(b)

0.2 0.25 0.3 0.35 0.4

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Interstitial water resistivity(ohm-m)

U1327

Fig 3.8 (a) U1327 water salinity (in ppt) derived from recovered core. Gas hydrate dissociation upon recovery is the cause of the low salinity

outliers (fresh water). (b) Estimated in situ water resistivity calculated from the equation of state of sea water (Fofonoff, 1985).

106

(a) 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4

6

8

10U1327 Pickett Plot

LWD log-derived Porosity

FF(R

t/Rw

)

m=0.26, a=3.74m=1.76, a=1.27

(b) 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4

6

8

10U1327 Pickett Plot

core-derived Porosity

FF(R

t/Rw

)

m=0.21, a=3.74m=1.76, a=1.27

Fig. 3.9 (a) Pickett Plot in Site U1327, the blue scattered points are the density-porosity values plotted against the FF data. A power-trendline is

then fit through the scattered points, and the empirical parameters from the equation of the best-fit trendline are a=3.74, m=0.26. The blue line

stands for the empirical relationship between LWD derived porosity and FF with m=1.76 and a=1.27. (b) Pickett Plot showing the results of

core-derived density porosity against the FF data at Site U1327. A power-trendline is then fit through the scattered points, the empirical

parameters from the equation of the best-fit trendline are a=3.74, m=0.21.

107

(a)

0 0.2 0.4 0.6 0.8

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Gas Hydrate Saturation

U1327

Sh (density porosity, m=1.76, a=1.27)Sh (density porosity, m=0.55, a=2.99)

(b)

0 0.2 0.4 0.6 0.8

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Gas Hydrate Saturation

U1327

Sh (core porosity, m=1.76, a=1.98)Sh (density porosity, m=1.76, a=1.27)

Fig.3.10 (a) Gas hydrate saturation from density-porosity and core-derived porosity calculation; the yellow line stands for gas hydrate saturation

calculated from core-density using the empirical value 1.98 and 1.76 for a and m respectively, the red line (in both (a) and (b)) stands for the gas

hydrate saturation from density-porosity using m=1.76, a =1.27. The blue line is the result from density- porosity, using m=0.55, a=2.99.

108

(a)

22 24 26 28 30 32 34

0

50

100

150

200

250

Dep

th b

elow

sea

floor

(m)

Interstitial water salinity

U1329

(b)

0.2 0.25 0.3 0.35 0.4 0.45

0

50

100

150

200

250

Dep

th b

elow

sea

floor

(m)

Interstitial water resistivity(ohm-m)

U1329

Fig. 3.11 (a) Water salinity (in ppt) from recovered sediment at Hole U1329C and U1329D, (b) Estimated in situ water resistivity calculated

from the equation of state of sea water (Fofonoff, 1985).

109

(a) 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4

6

8

10U1329 Pickett Plot

LWD log-derived Porosity

FF(R

t/Rw

)

m=0.31, a=2.94m=1.76, a=1.18

(b) 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

2

4

6

8

10U1329 Pickett Plot

core-derived Porosity

FF(R

t/Rw

)

m=0.26, a=3.02m=1.76, a=1.43

Fig. 3.12 (a) Pickett Plot in Site U1329, the blue scattered points are the density-porosity values against the FF data. A power-trendline is then

fit through the scattered points, the empirical parameters from the equation of the best-fit trendline are a=2.94, m=0.31. The blue line stands for

the empirical relationship between LWD density porosity and FF with m=1.76 and a=1.18. (b) Pickett Plot showing the results of core-derived

density porosity against the FF data. A power-trendline is fit through the scattered points, and the empirical parameters from the equation of the

best-fit trendline are a=3.02, m=0.26. The blue line is the power relationship with core-derived porosity and FF with m=1.76 and a=1.43.

110

(a)

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

120

140

Dep

th b

elow

sea

floor

(m)

Gas Hydrate Saturation

U1329

Sh (density-porosity, m=1.76, a=1.43)Sh (density-porosity, m=1.76, a=1.18)

(b)

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

120

140

Dep

th b

elow

sea

floor

(m)

Gas Hydrate Saturation

U1329

Sh (density-porosity, m=0.31, a=2.94)Sh (density-porosity, m=1.76, a=1.15)

Fig. 3.13 Gas hydrate saturation from density-porosity and core-derived porosity calculation of Site U1329; the blue line stands for gas hydrate

saturation using the empirical value 2.94 and 0.31 for a and m respectively, the red line (in both (a) and (b)) stands for the gas hydrate

saturation from density-porosity using m=1.76, a =1.18. The yellow line is the result from core-derived porosity, using m=1.76, a=1.43.

111

(a)

1200 1400 1600 1800 2000

0

50

100

150

200

250

300

350

400

Velocity (m/s)

Dep

thbe

low

seaf

loor

(m)

U1325

LWD SonicVis ionWL VelocityMCS Velocity

Base GHSZ

(b) 1400 1600 1800 2000 2200 2400

0

50

100

150

200

250

300

350

400

Dep

thbe

low

seaf

loor

(m)

Velocity (m/s)

U1326

LWD SonicVisionWL VelocityMCS Velocity

Base GHSZ

112

(c) 1300 1400 1500 1600 1700 1800 1900 2000

0

50

100

150

200

250

300

350

400

450

Dep

thbe

low

seaf

loor

(m)

Velocity (m/s)

U1327

LWD SonicVisionWL VelocityMCS Velocity

Base GHSZ

(d) 1200 1400 1600 1800 2000 2200

0

50

100

150

200

250

300

350

400

Dep

thbe

low

seaf

loor

(m)

Velocity (m/s)

U1329

LWD SonicVisionWL VelocityMCS Velocity

Base GHSZ

113

Fig. 3.14 The velocity data from the four transect drilling sites. The red line stands in all plots for the LWD velocity, which comes from the A-

Hole, the blue line is the WL velocity and the black points are the MCS velocity values. The purple line in Site U1327 stands for the

unconstrained interval velocity calculated from the VSP data. The depth of the Base of the Gas hydrate stability zone (equivalent to BSR) is

marked by the black horizontal line in each graph.

114

0 0.05 0.1 0.15 0.2 0.25 0.3

0

50

100

150

200

250

300

350

Dep

th b

elow

sea

floor

(m)

Gas hydrate saturation (fraction of pore space)

U1325

Fig.3.15 Gas hydrate saturations at Site U1325. The red line shows the results

calculated from LWD velocity data acquired from Hole U1325A.The blue line

is for the results calculated from WL velocity collected from Hole U1325C.

The black stars are the results from the MCS data calculated from the main

transect line MCS line 89-08. The location of the estimated base of GHSZ has

been marked by a thick black line.

115

0 0.1 0.2 0.3 0.4

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Gas hydrate saturation (fraction of pore space)

U1326

Fig.3.16 Gas hydrate saturations calculated for Site U1326. The red line is the

result from LWD velocity data acquired from Hole U1326A. The blue line is

the result calculated from WL velocity collected from Hole U1326D. The

black stars are the result from the MCS data calculated from the main transect

line MCS line 89-08. The location of the estimated base of GHSZ is marked

by a black line.

116

0 0.1 0.2 0.3 0.4

0

50

100

150

200

250

300

Dep

th b

elow

sea

floor

(m)

Gas hydrate saturation (fraction of pore space)

U1327

Fig.3.17 Gas hydrate saturations at Site U1327. The red line stands for the

results calculated from LWD velocity data acquired from Hole U1327A. The

blue line stands for the results calculated from WL velocity collected from

Hole U1327E. The black stars are the results from the MCS data calculated

from the main transect line MCS line 89-08. The location of the estimated

base of GHSZ has been marked in this graph.

117

0 0.1 0.2 0.3 0.4 0.5 0.6

0

20

40

60

80

100

120

140

160

180

U1329

Gas Hydrate Saturation (fraction of pore space)

Dep

th B

elow

Sea

floor

(m)

Fig. 3.18 Gas hydrate saturations for Site U1329. The red line stands for the

results calculated from LWD velocity data acquired from Hole U1329A. The

blue line stands for the results calculated from WL velocity collected from

Hole U1329D. The black stars are the results from the MCS data calculated

from the main transect line MCS line 89-08. The location of the estimated

base of GHSZ has been marked in this graph.

118

Fig. 4.1 Multibeam bathymetry map with main seismic sections used for Site

U1325 along the transect across the accretionary prism offshore Vancouver

Island (courtesy of D. Kelley, J. Delaney, and D. Glickson, University of

Washington, and C. Barnes, C. Katnick, NEPTUNE Canada, University of

Victoria; funded by the University of Washington and the W.M. Keck

Foundation). MCS = multichannel seismic

119

Fig. 4.2 Seismic line 89-11 with NW-SE direction, located ~1.5km NW away from Site U1325. The location of Site U1325 and Site U1326 can

be projected onto this profile, which has 5.3km offset from each other. The “bowl-like” sedimentation structure within the V-shape basin is

clearly depicted in this seismic profile. Different sediment stratigraphy can be distinguished due to different intensity of seismic reflections. The

BSR is seen clearly, except where it falls into the V-shaped slope-basin.

120

Fig. 4.3 Subbottom profiler data from line ODP7, the red line indicates the location of line 89-11. The transparent layer of Holocene sediment

ontop of higher reflective material can be clearly seen in this image. The layer is composed of the sediment with extremely high porosity and low

density, indicating the pelagic sedimentation on the seafloor without any influx of new turbidite sediment and compaction effects (Novosel,

2005). The vertical time scale is a reduced traveltime as a deep-water delay of 1.0 seconds was used in the recording of the 3.5 kHz profiler data.

121

Fig. 4.4 Section of MCS Line 89-08 over Site U1325. The yellow line indicates the boundary between lithostratigraphic Subunits IA and IB,

which can be traced for ~1.5km around Site U1325. The BSRs is indicated by the green line. There are several small-scale faults cutting through

the buried ridge and some of them almost extend to the BSRs.

122

Fig. 4.5 Seismic profile of MCS line ODP3, the green dashed line shows the discontinuous BSR. The buried uplifted ridge can be located in the

centre just below the projected location of Site U1325.

123

Fig. 4.6 Seismic Profile of MCS line ODP2, which is located 2.5km SE of Site U1325. The rough location of the uplifted ridge in centre is

marked by the seismic high reflectivity shown at the projected location of Site U1325. The complicated appearances of the sediments at the NE

side of the basin indicate progressive deformation inside the basin. BSRs are difficult to identify in this seismic section.

124

Fig. 4.7 Seismic Profile of MCS line ODP7. The yellow line indicates the unconformity; the deeply cutting faults within the buried ridge are

indicated by dashed black lines. The green arrow points out the location where a BSR is speculated to appear.

125

Fig. 4.8 Section of SCS line CAS02B_line05_04 over Site U1325. The green line indicates the unconformity boundary between lithostratigraphic

Subunits IA and IB, which can be traced for ~1.2 km around Site U1325. The proposed location of BSRs is indicated by the yellow line, which

was projected from observations along the other seismic profiles. The black dashed line indicates the general pattern of the sediment, which is

semi-similar to the seafloor. There are several small-scale fractures distributed inside the highly deformed sediments. Note the various side-

echoes, which have been marked by black dashed lines on the two sides of the basin.

126

(a) (b) Fig. 4.9 (a) extracted wavelet from MCS line 89-08 with a sample interval is 0.002s; the length in time is 0.1s. The wavelet was generated by

choosing traces within a 500m distance from Site U1325A, and a starting time of 2.9 s TWT to 3.5 s TWT. A total of 94 traces were selected to

generate the wavelet. (b) Computed spectrum of the extracted wavelet. The computed spectrum within the frequency domain of 0 - 32Hz shows

little noise, but from 32-64 Hz, which is the main frequency band of the seismic data, the noise increases.

127

128

Fig. 4.10 Synthetic seismogram using the wavelet shown in Figure 4.10 (from MCS line 89-08). The first column is the integrated log time–

depth conversion; the second column is the computed T-D chart velocity (m/s). The following columns are the LWD velocity data, the LWD

density values from Hole U1325A, computed acoustic impedance (AI), and reflection coefficients (RC). The extracted wavelet is shown used in

the convolution to produce the synthetic traces (both polarities are shown). The resistivity is used as the reference log here. The seismic trace

extracted from line 89-08 at U1325 is shown giving a correlation coefficient of 0.42. The location of the BSR is marked by the yellow dashed

line. The synthetic seismogram in the zone indicated by the two green dashed lines has relatively low reliability because of the abnormally low

density values from the LWD log, which may be caused by degraded borehole conditions.

129

130

Fig. 4.11 The seismic profile of MCS line 89-08 with generated synthetic seismogram, logging data, core pore-fluid chlorinity data and

lithologic units. The first column shows the pore-fluid chlorinity data. The second column shows the resistivity data (the red line stands for

U1325A LWD data and the green line stands for U1325C WL data). The gray bar is showing the lithologic units. The column with red wiggle

line inside is the synthetic generated from the MCS 89-08 seismic data convolved with LWD velocity and density data. The column following is

the velocity data (the red line stands for the WL logging data, the green line is the MCS velocity). The last column includes the gas hydrate

saturation calculated from LWD resistivity data.

131

Fig. 4.12 The seismic profile of MCS line ODP7 with generated synthetic seismogram, logging data, core-derived pore-fluid chlorinity data and

lithologic units. Pattern of data shown is identical to those shown in Fig. 4.12.

132

Fig. 4.13 Bathymetry map and seismic lines used at Site U1326; (courtesy of D. Kelley, J. Delaney, and D. Glickson, University of Washington,

and C. Barnes, C. Katnick, NEPTUNE Canada, University of Victoria; funded by the University of Washington and the W.M. Keck Foundation).

133

Fig. 4.14 Seismic profile of SCS line CAS03B_inline01, the red vertical line marks the location of CAS03B Xline10 which almost goes through

the Site U1326. The BSR is marked in the graph and the location of the slump is pointed out by the arrow.

134

Fig.4.15 Profile of SCS line CAS03B-Xline01. The BSR is indicated by the yellow dashed line. The vertical red line indicates the location of the

intersecting MCS line CAS3MCS99 (shown in Fig. 4.18).

135

Fig. 4.16 Seismic profile of SCS line CAS03_Xline03. The BSR has been marked in the graph by the black dashed line which extends from SP

140-200. The red line is also the location of the transect line CAS3MCS99. The location of the U1326A is marked at the top.

136

Fig. 4.17 Seismic profile MCS line 89-08, the red line marks the location of seismic line CAS3MCS99, which is roughly the location of Site

U1326. The BSR is indicated with a yellow dashed line. The green dashed line inside the frontal basin is the boundary between the well-bedded

and slightly uplifted sediment layers.

137

Fig. 4.18 Section of MCS line CAS3MCS99 crossing the ridge in a southeast–northwest direction. A series of faults is clearly visible, which are

indicated by black dash lines, resulting in a seafloor displacement of up to 25 m. A possible slope slump (additional to the main slump already

identified earlier) is indicated by the yellow line on the seafloor, and the black arrow points out the location of the BSR. The locations of Hole

U1326A and the originally proposed site CAS-03B have been marked on the top.

138

(a) (b)

Fig. 4.19 (a) extracted wavelet from MCS line CAS03MCS99 with a sample interval of 0.002s; the length in time is 0.1s. The wavelet is

generated from traces within a 300m distance from Site U1326A, with a time range of 2.5 s TWT to 2.95 s TWT. A total of 247 traces were

selected to generate the wavelet. (b) Computed spectrum of the extracted wavelet. The computed spectrum within the frequency domain of 0 -

192Hz shows little noise, except within the frequency band of 32-64 Hz.

139

140

Fig. 4.20 Synthetic seismogram using the wavelet shown in Figure 4.20 (from MCS line CAS3MCS99). The first column is the integrated log

time –depth conversion; the second column is the computed T-D chart velocity (m/s). The following columns are the LWD velocity data from

Hole U1326A, computed acoustic impedance (AI), and reflection coefficients (RC). The extracted wavelet is shown used in the convolution to

produce the synthetic traces (both polarities are shown left and right of the real seismic trace). The gas hydrate enrichment zone and the depth of

the BSR have been marked out by the yellow dashed lines. The seismic trace was extracted from MCS line CAS3MCS99 at Site U1326 are

shown giving an overall correlation coefficient of 0.52 to the synthetics.

141

142

Fig. 4.21 Detailed multibeam bathymetry map around Site U1327 and Cold Vent Site U1328 (also known as Bullseye Vent). The location of the

main MCS 3-D grid is indicated by a yellow box; the 1999 vent field SCS grid and the 2000 SCS grid around Bullseye Vent are indicated in

black boxes. The orange dots show the location of Site U1325, Site U1327 and Site U1328. The orange line shows the main transect line MCS

89-08. The survey focuses within the two relatively topography high, which goes a NW-SE direction on the base map. The green line has pointed

out the westward dipping thrust fault.

143

Fig. 4.22 Detailed multibeam bathymetry map around Site U1327 and Cold Vent Site U1328 (also known as Bullseye Vent). The location of the

main MCS 3-D grid is indicated by a yellow box; the 1999 vent field SCS grid and the 2000 SCS grid around Bullseye Vent are indicated in

black boxes. The orange dots show the location of Site U1325, Site U1327 and Site U1328. The survey focuses within the two topographic highs,

which are in a NW-SE direction. The green line has pointed out the westward dipping thrust fault. (courtesy of D. Kelley, J. Delaney, and D.

Glickson, University of Washington, and C. Barnes, C. Katnick, NEPTUNE Canada, University of Victoria; funded by the University of

Washington and the W.M. Keck Foundation).

144

Fig.4.23 Seismic profile of MCS line 8908 around site U1327, the location of Site U1327A is projected onto this line. The steeply westward

dipping thrust fault is indicated by the black dashed line. A strong and continuous BSR can be seen from the mid-slope to the thrust fault outcrop.

The green arrows point out the ridges of accreted sediments.

145

Fig. 4.24 Seismic profile of MCS99_Inline38 of the 3D grid. The BSR has been marked by a yellow line. The location of Site U1327A is

indicated on top of the image. The NW boundary of the sediment-filled basin has been marked by a green dashed line, and small scale faults

located within the basin near CDP 200 are shown by black dashed lines.

146

Fig. 4.25 Seismic line 2008007PGC DTSEXT060, Huntec envelope data. The orange dashed line marks the boundary between the first and the

second layers, the green line indicates the boundary between the second and the third layer. A deep-water delay of 1.465s was used to record the

Huntec data. The red vertical line is the intersection point with MCS99_Inline38.

147

Fig.4.26 Seismic line PGC007-2008-line60 showing the airgun data. The red vertical line indicates the location of MCS Inline 38. The BSR

depth is marked in the figure by arrows (vertical time is reduced by a deep-water delay).

148

Fig. 4.27 Seismic line 2008007PGCDTSEX058 with Huntec envelope data. The red vertical line is MCS Inline 38.

149

Fig. 4.28 Seismic line PGC007-2008line 58 profile showing airgun data; The BSR is highlighted by the arrows.

150

(a) (b)

Fig. 4.29 (a) extracted wavelet from MCS Inline38 with a sample interval is 2ms; the total length is 0.1 s. The wavelet was generated from traces

within 200 m distance from Site U1327A, with a time window of 1.8 s TWT to 2.15 s TWT. A total of 40 traces were selected to generate the

wavelet. (b) Computed spectrum of the extracted wavelet. The computed spectrum within the frequency domain of 0 – 192 Hz shows relatively

high noise levels, especially at ~100 Hz, which is the main frequency band of the seismic data.

151

152

Fig. 4.30 Synthetic seismogram generated using the wavelet shown in Figure 4.29 (from MCS Inline38). The first column is the integrated log

time –depth conversion; the second column is the computed T-D chart velocity (m/s). The following columns are the LWD velocity data, the

LWD density values from Hole U1327A, computed acoustic impedance (AI), and reflection coefficients (RC). The extracted wavelet is used to

produce the synthetic traces (both polarities are shown). The seismic trace closest to the borehole was extracted from MCS Inline38 at U1327

and is repeated five times for easier visual comparison. A correlation coefficient of 0.179 is achieved without adjustment of the synthetic traces.

153

154

Fig. 4.31 Seismic MCS99_Inline38 profile with logging and core derived information as well as the synthetic seismogram superimposed. The

first column is the pore-fluid chlorinity profile. The second column shows the resistivity log-data (the red line is for U1327A LWD data and the

green line stands for U1327E WL data). The gray bar is showing the lithologic units. At the exact borehole location the synthetic trace is

superimposed. The next column shows the velocity data (the red line stands for the WL logging data, the green line is the VSP interval velocity).

The last column shows the gas hydrate saturation calculated from LWD resistivity data. BSR has been traced by the red line inside the graph.

155

Figure 4.32 Image showing the heterogeneity of the high-concentration zone

of Site U1327. Shown are (from left to right) LWD RAB image and extracted

ring-resistivity log, Infra-red image of the recovered core of Hole U1327C,

electrical WL log and Infra-red image of the recovered core of Hole U1327D,

and WL log of Hole U1327E. The distance between all boreholes is 60m. The

high-concentration zone (seen as bright zone in the RAB image and dark-blue

colors in the IR image) thins to the east and is entirely lost in Hole U1327E.

156

(a) (b)

Fig. 4.33 (a) extracted wavelet from MCS line8908 with a sample interval is 2ms; the total length is 0.1s. The wavelet was generated from traces

within 500m distance from Site U1327A, with a time window of 1.7s TWT to 2.2 s TWT. (b) Computed spectrum of the extracted wavelet. The

computed spectrum within the frequency domain of 0 - 64Hz shows relatively very low noise levels, except at frequency ~40Hz where the noise

ratio reaches ~10%.

157

158

Fig. 4.34 Synthetic seismogram using the wavelet shown in Figure 4.39 (from MCS line 8908). The first column is the integrated log time –

depth conversion; the second column is the computed T-D chart velocity (m/s) which is identical as for line MCS99_inline38. The following

columns are the velocity data from LWD velocity measurement, the LWD density values from Hole U1327A, computed acoustic impedance

(AI), and reflection coefficients (RC). The extracted wavelet is used to produce the synthetic traces (both polarities are shown). The seismic trace

closest to the borehole was extracted from MCS line8908 at U1327 and is repeated five times for easier visual comparison. A correlation

coefficient of 0.106 is achieved without further adjustment of the synthetic traces. The gas hydrate enrichment region and the depth of BSR have

been marked out by the yellow dashed lines. The green circle shows the part of the LWD data where some data points in the velocity have been

removed due to poor data quality. Note that the corresponding density data also show some degradation (low values). The synthetic seismogram

thus yields a false reflection event. However, the location of the poor data quality coincides with the depth range of the BSR (borehole condition

may be degraded at the transition from some hydrate-bearing sediment to free gas bearing sediment).

159

160

Fig. 4.35 Seismic MCS line 8908 profile with logging and core derived information as well as the synthetic seismogram superimposed. The first

column is the pore-fluid chlorinity profile. The second column is showing the resistivity log-data (the red line is for U1327A LWD data and the

green line stands for U1327E WL data). The gray bar is showing the lithologic units. At the exact borehole location the synthetic trace is

superimposed. The next column is the showing velocity data (the red line stands for the WL logging data, the green line is the VSP interval

velocity). The last column shows the gas hydrate saturation calculated from LWD resistivity data. BSR has been traced by the red line inside the

graph.

161

162

Fig. 4.36 Seismic profile of SCS line PGC007-2008line60 (airgun data) with logging and core derived information as well as the synthetic

seismogram generated from MCSinline38. The first column is the pore-fluid chlorinity profile. The second column is showing the resistivity log-

data (the red line is for U1327A LWD data and the green line stands for U1327E WL data). The gray bar shows the lithologic units. At the exact

borehole location the synthetic trace is superimposed. The next column shows the velocity data (the red line stands for the WL logging data, the

green line is the VSP interval velocity). The last column shows the gas hydrate saturation calculated from LWD resistivity data.

163

Fig. 4.37 Multibeam bathymetry map shows the northeast end of the accretionary prism on Northern Cascadia, offshore Vancouver Island.

(courtesy of D. Kelley, J. Delaney, and D. Glickson, University of Washington, and C. Barnes, C. Katnick, NEPTUNE Canada, University of

Victoria; funded by the University of Washington and the W.M. Keck Foundation). Site U1329 is indicated by the orange circle, transect MCS

line 89-08 which goes through the drilling site is shown as the red line.

164

Fig. 4.38 Seismic line CAS05C_line03 added with lithology bar. The BSR can be clearly traced for ~2.5km and extends from CDP 120 to 280.

The location of the MCS Line 89-08 is indicated for reference.

165

Fig. 4.39 MCS line 89-08 across Site U1329 with lithology units superimposed. The BSR can be traced identified in the SW portion of the line,

but it is progressively lost towards the NE into shallower water. The unconformity at Site U1329 (boundary between lithologic units II and III) is

close to the BSR at Site U1329 and thus a more complex interference pattern can be seen in the low-frequency data than in the higher-frequency

SCS data shown in Figure 4.38. The locations of the SCS lines are indicated for reference.

166

Fig. 4.40 Seismic profile of COAMS MCS99 data of ODP1. The BSR can be tracked both at the SW edge of the basin and around Site U1329.

The unconformity boundaries between different sedimentary strata have been marked out by the yellow dashed line. Strong seismic reflectivity is

seen at the NE edge of the basin which may be an indicator of free gas.

167

(a) (b)

Fig. 4.41 (a) extracted wavelet from MCS line 89-08 with a sample interval of 0.002 s; the length in time is 0.1 s. The wavelet generated by

choosing traces within 500m distance from Site U1329A, with a starting time of 1.25 s TWT to 2.0 s TWT. A total of 41 traces were selected to

generate the wavelet. (b) Computed spectrum of the extracted wavelet. The two frequency peaks reach respectively at ~24 Hz and 36 Hz. The

signal to noise ration is almost equal to 1.

168

169

Fig. 4.42 Synthetic seismogram using the wavelet shown in Figure 4.41 from MCS line 89-08. The first column shows the location of data points

(in time) used as log time–depth function; the second column is the computed T-D chart velocity (m/s). The following columns are the LWD

velocity log, the LWD density values from Hole U1329A, computed acoustic impedance (AI), and reflection coefficients (RC). The extracted

wavelet is used in the convolution to produce the synthetic traces (both polarities are shown.) The seismic traces extracted from line 89-08 at

U1329 yield a correlation coefficient with the synthetic seismogram of 0.3.

170

171

Fig. 4.43 The seismic profile of line 89-08 with generated synthetic seismogram, logging data, chlorinity data and lithologic units. The first

column stands for chlorinity data. The layers with dispersed fresher data intermingle inside the whole sediment. The second column stands for

the resistivity data, the red line stand for U1329A LWD data. The column with yellow wiggle line is the synthetic generated from MCS seismic,

LWD velocity and density data. The column following is the velocity, the red line stand for the WL logging data, the green line is the MCS

velocity. The last column includes the gas hydrate saturation we calculated from LWD resistivity data.

172

(a) (b)

Fig. 4.44 (a) extracted wavelet from CAS05B_line03 with a sample interval of 0.002s; the length in time is 0.1 s. The wavelet generated by

choosing traces within 200 m distance from Site U1329A, with a starting time of 1.2 s TWT to 1.7 s TWT. (b) Computed spectrum of the

extracted wavelet. The noise ratio reaches 0.3 at 120 Hz. However, when signal is strongest at ~70 Hz, the noise ratio is low.

173

174

Fig. 4.45 Synthetic seismogram using the wavelet shown in Fig. 4.43(a) from SCS line CAS05B_03. The first column shows the location of data

points (in time) used as log time–depth function; the second column is the computed T-D chart velocity (m/s). The following columns are the

LWD velocity log, the LWD density values from Hole U1329A, computed acoustic impedance (AI), and reflection coefficients (RC). The

extracted wavelet is shown used in the convolution to produce the synthetic traces (both polarities are shown.) The seismic traces extracted from

line CAS05B_line03 at U1329 yield a correlation coefficient with the synthetic seismogram of 0.7.

175

176

Fig. 4.46 The seismic profile of line CAS05B_line03 with generated synthetic seismogram, logging data, chlorinity data. The first column

stands for chlorinity data. The second column is stands for the resistivity data fromU1329A LWD data. The column with red wiggle line

inside is the synthetic generated from SCS seismic, LWD velocity and density data. The column following is the velocity, the red line stand

for the WL logging data, the green line is the MCS velocity generated from line8908. The last column includes the gas hydrate saturation we

calculated from LWD resistivity data.

177

(a) (b)

Fig. 4.47 (a) extracted wavelet from line ODP1MCS99 with a sample interval of 0.002 s; the length in time is 0.1 s. The wavelet is generated by

choosing traces within 200 m distance from Site U1329A, with a starting time of 1.2 s TWT to 1.65 s TWT. (b) Computed spectrum of the

extracted wavelet. The noise ratio reaches 0.3 at ~70 Hz.

178

179

Fig. 4.48 Synthetic seismogram generated from ODP1MCS99. The first column shows the location of data points (in time) used as log time–

depth function; the second column is the computed T-D chart velocity (m/s). The following columns are the LWD velocity log, the LWD density

values from Hole U1329A, computed acoustic impedance (AI), and reflection coefficients (RC). The extracted wavelet is shown used in the

convolution to produce the synthetic traces (both polarities are shown). The seismic traces extracted from line ODP1MCS99 at U1329 yield a

correlation coefficient with the synthetic seismogram of ~0.7.

180

181

Fig. 4.49 The seismic profile of line ODP1MCS99 with generated synthetic seismogram, logging data, chlorinity data and lithologic units. The

first column is the chlorinity data. The second column is the LWD resistivity data from Hole U1329A. The synthetic trace generated is shown in

yellow. The following column is velocity data (the red line is the WL logging data, the green line is the MCS velocity generated from MCS line

89-08). The last column includes the gas hydrate saturation calculated from LWD resistivity data.

182

(a)

0 0.1 0.2 0.3 0.4 0.5 0.6

0

50

100

150

200

250

300

350

400

U1325

Gas Hydrate Saturation (fraction of pore space)

Dep

th B

elow

Sea

floor

(m)

LWD v elocityLWD resistiv ity m=0.43, a=3.48

LWD resistiv ity m=1.76,a=1.17

(b)

0 0.1 0.2 0.3 0.4 0.5 0.6

0

50

100

150

200

250

300

U1326

Gas Hydrate Saturation (fraction of pore space)

Dep

th B

elow

Sea

floor

(m)

LWD velocityLWD resistivity m=0.27, a=4.65LWD resistivity m=1.76,a=1.57

(c)

0 0.1 0.2 0.3 0.4 0.5 0.6

0

50

100

150

200

250

300

U1327

Gas Hydrate Saturation (fraction of pore space)

Dep

th B

elow

Sea

floor

(m)

LWD velocityLWD resistivity m=0.26, a=3.74LWD resistivity m=1.76,a=1.27

(d)

0 0.1 0.2 0.3 0.4 0.5 0.6

0

20

40

60

80

100

120

140

160

180

U1329

Gas Hydrate Saturation (fraction of pore space)

Dep

th B

elow

Sea

floor

(m)

LWD velocityLWD resistivity m=0.31, a=2.94LWD resistivity m=1.76,a=1.18

Figure 5.1 Comparison of gas hydrate saturation estimates from logging data

(resistivity and velocity) at all four transect sites of IODP Expedition 311. The

depth of the BSR is indicated as horizontal dashed line.

183

Tables

Hole Latitude Longitude Water depth (mbsl)

Cores (N)

Cored (m)

Recovered (m)

Recovery (%)

Drilled (m)

Penetration (m)

U1325A 48°38.691′N 126°58.991′W

2201.1 0 0 0.00

350.0 350.0

U1325B 48°38.694′N 126°58.999′W 2194.8 28 205.5 141.17 68.7 1.0 206.5

U1325C 48°38.701′N 126°59.007′W 2194.8 15 115.5 62.49 54.1 188.8 304.3

U1325D 48°38.701′N 126°59.007′W 2194.6 1 4.7 4.69 99.8 0.0 4.7

Site U1325 totals: 44 325.7 208.35 64.0 539.8 865.5

U1326A 48°37.635′N 127°3.029′W

1828.1 0 0 0.00

300.0 300.0

U1326B 48°37.638′N 127°3.023′W 1828.4 1 1.5 1.55 103.3 0.0 1.5

U1326C 48°37.638′N 127°3.023′W 1828.0 13 85.7 54.39 63.0 1.0 86.7

U1326D 48°37.628′N 127°3.043′W 1827.9 20 192.6 121.96 63.3 107.4 300.0

Site U1326 totals: 34 279.8 177.9 229.6 408.4 688.2

U1327A 48°41.887′N 126°51.921′W 1322.1 0 0 0.00

300.0 300.0

U1327B 48°41.889′N 126°51.914′W 1306.1 1 9.5 9.79 103.1 0.0 9.5

U1327C 48°41.889′N 126°51.914′W 1304.5 35 297 262.17 88.3 3.0 300.0

U1327D 48°41.895′N 126°51.906′W 1303.7 17 75.9 53.99 71.1 224.1 300.0

U1327E 48°41.900′N 126°51.896′W 1303.4 4 12.5 12.37 99.0 287.5 300.0

Site U1327 totals: 57 394.9 338.32 85.7 814.6 1209.5

U1328A 48°40.072′N 126°51.022′W

1268.1 0 0 0.00

300.0 300.0

U1328B 48°40.048′N 126°51.056′W 1267.8 10 54.5 39.65 72.8 2.0 56.5

U1328C 48°40.057′N 126°51.044′W 1267.7 27 242.5 195.64 80.7 57.5 300.0

U1328D 48°40.067′N 126°51.020′W 1266.7 3 15 10.03 66.9 0.0 15.0

U1328E 48°40.080′N 126°50.987′W 1264.7 13 49 13.33 27.2 186.0 235.0

Site U1328 totals: 53 361 258.65 71.6 545.5 906.5

U1329A 48°47.369′N 126°40.713′W 959.1 0 0 0.00

220.0 220.0

U1329B 48°47.375′N 126°40.720′W 953.0 1 9.5 10.02 105.5 0.0 9.5

U1329C 48°47.375′N 126°40.720′W 946.4 23 187.5 186.16 99.3 2.0 189.5

U1329D 48°47.362′N 126°40.716′W 946.4 1 9.5 1.03 10.8 201.0 210.5

U1329E 48°47.385′N 126°40.716′W 945.5 10 48 37.74 78.6 79.0 127.0

Site U1329 totals: 35 254.5 234.95 92.3 502.0 756.5

Table 2.1 Expedition 311 operations summary for location of boreholes, and penetration depths (from Riedel et al. 2005)

184

a

Reservoir lithology

n

High salinity Low salinity Sandstone (quartz)

2.1661 1.761 2.264

Limestone (Calcite)

1.834 1.299 1.674

Unconsolidated material

1.715 0.551 0.710

Pooled estimate 1.9386 1.035

Table 3.1 Empirical parameters for the modified Archie relation (Modified

from Pearson et al., 1983) (Collett, 2000).

185

Site Depth (mbsf)

U1325 57

U1326 71

U1327 71

U1329 90

Table 3.2 minimum depth of reliable Vp estimation from LWD data at IODP

Expedition 311 sites (Goldberg et al., 2008).

186

Reservoir

constituent

Vp

(km/s)

Vs

(km/s)

Bulk-

density

(g/ 3cm )

Bulk

moduli

(K,

Mpa)

Shear

moduli

(G, Mpa)

Reference

Methane hydrate

structure- Ι

3.35* 1.68** 0.9 6713*** 2540*** *Estimated from Whiffen et al. (1982),

**Estimated from Pandit and King (1982),

***Calculated fromVp-Vs

Pure water 1.5 0.00 1.03 2250 0 Schlumberger Educational Services (1989)

Ice 3.8 2.00 0.92 8400 3700 King (1984)

Quartz 5.98 4.04 2.65 38000 44000 Toksoz et al. (1976)

Clay 3.40 1.60 2.60 21200 6666 Tosaya and Nur (1982)

Marine sediment 4.37 2.60 2.75 27200 18300 Lee et a. (1996)

Table 3.3 Acoustic velocity, elastic constants, and bulk-densities for various reservoir constituents.

187

1325A 1325C 1326A 1326D 1327A 1327D 1327E 1329A 1329C 1329D

Pre-cruise seismic BSR [mbsf] * 230 234 223 125

(a) Water Depth (m) 2196 1828 1308 950

(a) Distance from Deformation

Front (km) 11 5 21 38

(a)Depth below seafloor [s] TWT 0.2883 0.2817 0.272 0.154

(a) LWD-Rt 240 * 260 * 230 * * n.d. * *

(a)CWL-Vp * d.n.r.b * 260 * * 228 * * 125

(a)CWL-Rt * d.n.r.b. * 260 * 235 235 * * 125

(a)Deepest Cl− anomaly * 238.9 * 264 * 222 * * 128 *

188

(a)C1/C2 * 243 * * * 225 * * 119 *

(a)i-C4/n-C4 or C3 * * * * * 225 * * * *

(a)VSP * * * * * 245 * * * *

(a)Best estimate of BGHSZ 240.5

264 230 124

(b)Thickness of the GHOZ (m)

168

217

119

0

(c)Thickness of potential GHOZ

(m)

238

261

221

113

Table 4.1. Summary of Site-information for the individual boreholes of all four transect sites including water depth, distance to deformation front,

and estimated BSR depth using various techniques. * Velocity used for BSR depth conversion from TWT: 1636 m/s maximum average velocity

from seafloor to BSR; 1619 m/s minimum average from seafloor to BSR, reported depth is average between min and max estimate; note that

velocity was determined from VSP at ODP Site 889, n.d. Not detected, d.n.r.b. Wire-line logging tools did not reach BSR depth. (a) from

Riedel et al. (2006) , (b) from Malinverno et al. (2008), (c) from Pohlman et al. (2008).

189

TVD (Seismic)(meter) Time (2-way travel time)(s) 2201.195 2.93987 2221.942 2.9678

2237 2.98666 2239.036 2.99312

2241 2.99912 2246.806 3.00762

2250 3.01244 2258.719 3.03072

2260 3.03344 2262 3.03527

2272.188 3.04068 2276 3.04261 2288 3.05979

2300.16 3.08261 2307 3.09502

2310.002 3.09835 2311 3.09943

2321.398 3.10275 2329 3.10513

2331.758 3.10619 2345.744 3.11172

2347 3.11221 2349 3.11348 2350 3.11423 2351 3.11528 2353 3.11724 2354 3.11804 2357 3.11994 2375 3.12994

2377.342 3.13184 2378 3.13237 2395 3.1456

2405.313 3.15455 2412 3.16023 2414 3.16198 2417 3.1646 2419 3.16627

2423.444 3.17021 2445 3.18914

2461.258 3.20544 2465 3.20916 2466 3.21014 2467 3.21109

2500.108 3.24677 2541.548 3.29102 2559.16 3.30971

Table 4.2 Time-Depth Chart in U1325.

190

TVD (Seismic) (Meters) Time (TWT) (s) 1829 2.50457

1830.1 2.50567 1838 2.51322 1844 2.51998

1848.1 2.52428 1852 2.52831 1853 2.5293 1854 2.53055 1859 2.53711

1860.1 2.53865 1864 2.544 1869 2.55429 1870 2.55529 1871 2.55607

1871.6 2.55654 1876 2.55992 1879 2.56197

1881.1 2.56341 1883 2.56472 1885 2.56627

1886.1 2.56831 1890 2.58302 1891 2.58413

1895.1 2.58543 1899 2.58669

1900.1 2.58888 1901.6 2.59139 1905.1 2.59809 1907.6 2.60273 1909 2.60554

1909.6 2.6067 1910 2.60734 1911 2.60832

1912.1 2.60973 1915.6 2.61418 1918 2.61759

1919.1 2.61915 1920 2.62046

1928.6 2.62822 1936.1 2.63526 1939 2.6381

1940.6 2.64049 1941 2.64106 1946 2.64872 1953 2.65704 1956 2.66024 1957 2.66159

1959.6 2.66495

191

1971 2.68011 1972 2.68171

1972.6 2.68264 1977 2.68928

1978.6 2.69298 1980 2.69623 1981 2.6976 1982 2.69844

1983.1 2.6994 1993 2.70793 2000 2.71869

2001.6 2.72128 2009 2.73299 2013 2.73652 2014 2.73775

2015.1 2.73893 2021 2.74528 2022 2.74636 2030 2.75455

2033.1 2.75794 2034 2.75899 2035 2.76015

2036.1 2.76259 2040.1 2.77497 2044 2.78194 2045 2.78329 2047 2.7845

2049.1 2.78567 2053.1 2.79396 2055.1 2.79828 2059 2.80315 2061 2.80444 2071 2.81045

2072.1 2.8114 2078.1 2.81734 2081.6 2.82077 2084 2.8231 2085 2.82415

2086.6 2.82583 2089.1 2.82809 2091 2.83016 2095 2.83482

2100.6 2.84154 2102 2.84321

2104.6 2.84621 2109.1 2.85144 2112.1 2.85483

2326.126 3.10788

Table 4.3 Time-Depth Chart for Site U1326.

192

TVD(seismic) (meters) Time(TWT)(s) 1322.226 1.76973 1325.6 1.77425 1342.6 1.79758 1354.6 1.81364 1373.1 1.83804 1386.1 1.85496 1396.1 1.86803 1399.1 1.87175 1417.6 1.89528 1432.6 1.91406 1435 1.91699 1441 1.93033

1444.6 1.9344 1447.6 1.93765 1451.1 1.94096 1460.1 1.94948 1464.6 1.95393 1494.1 1.98612 1499.1 1.9914 1506.6 1.99951 1514.1 2.00761 1524.6 2.01528 1533.6 2.02154 1538.1 2.02473 1543 2.02818 1545 2.02995

1546.1 2.03104 1549 2.03396

1554.1 2.03975 1566.6 2.05512 1579.6 2.07149 1588.1 2.08236 1590.6 2.0853 1594.1 2.0896 1598.1 2.09478 1605.1 2.1032 1608.6 2.10733 1613.6 2.11342 1618.1 2.11852 1624.1 2.12556 1626 2.12783 1626 2.12783 1626 2.12783

1625.6 2.12735

Table 4.4 Time Depth Chart for Site U1327.

193

TVT (seismic) (meters) TWT (s) 959.64 1.27408

963.145 1.27881 973.673 1.29302 987.167 1.31084 1001.74 1.32969

1015.774 1.34751 1034.125 1.37003

1050 1.38918 1052 1.39401

1054.636 1.39687 1072 1.41549 1082 1.42742

1099.435 1.44194 1115 1.45507 1116 1.45607

1123.724 1.46428 1128 1.46886 1131 1.47193 1132 1.47312 1133 1.47432 1135 1.47664 1216 1.57528

1216.45 1.57583

Table 4.5 Time-Depth Chart in U1329.

194

Appendix I: Programming of gas hydrate

saturation estimates from velocity data

*Main program:*

for i=1:length(b)

ok_num(i,1)=saturation(b(i),f(i),vb(i),W);

end

*Subprogram: “saturation”*

Function ok_num=saturation(b,f,Vb,W)

xx=strcat('1/',num2str(Vb),'=(',num2str(W),'*',num2s

tr(f),'*(1-

Sh))/((5.816254e+10)/((6.441562e+19*',num2str(b),'+1

.395296e+21*',num2str(b),'*',num2str(f),'-

1.124781e+21*Sh*',num2str(b),'*',num2str(f),')^0.5))

+(1-',num2str(W),'*',num2str(f),'*(1-

Sh))/((87837)/(20100+38458*',num2str(f),'-

32338*Sh*',num2str(f),'))');

ok_num_t=Solve(xx,'Sh');

ok_num=double(ok_num_t(1,1));

for i=2:length(ok_num_t)

if (abs(double(ok_num_t(i)))<abs(ok_num))

ok_num=double(ok_num_t(i));

else

end

end