95
Acoustic, elastic and physical properties of overconsolidated sands and reservoir fluids - Experimental measurements, modelling, and implications for reservoir characterization, time-lapse seismic monitoring, and geomechanics Sirikarn Narongsirikul Dissertation for the degree of Philosophiae Doctor (Ph.D.) Department of Geosciences Faculty of Mathematics and Natural Sciences University of Oslo, Norway Submitted: February 2020

Acoustic, elastic and physical properties of

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Acoustic, elastic and physical properties of

Acoustic, elastic and physical properties of

overconsolidated sands and reservoir fluids

- Experimental measurements, modelling, and implications for reservoir

characterization, time-lapse seismic monitoring, and geomechanics

Sirikarn Narongsirikul

Dissertation for the degree of Philosophiae Doctor (Ph.D.)

Department of Geosciences

Faculty of Mathematics and Natural Sciences

University of Oslo, Norway

Submitted: February 2020

Page 2: Acoustic, elastic and physical properties of

© Sirikarn Narongsirikul, 2020 Series of dissertations submitted to the Faculty of Mathematics and Natural Sciences, University of Oslo No. 2268 ISSN 1501-7710 All rights reserved. No part of this publication may be reproduced or transmitted, in any form or by any means, without permission. Cover: Hanne Baadsgaard Utigard. Print production: Reprosentralen, University of Oslo.

Page 3: Acoustic, elastic and physical properties of

i

Abstract

Hydrocarbons in easily accessible areas have been explored, with much of it having

been tapped out to supply the world’s energy demand. The growing consumption has

driven exploration to continue in harsher and more challenging areas, in parallel with

finding alternative renewable resources. In doing so, many oil and gas exploration

companies have shown an interest in applying for exploration licenses in areas towards

the northern hemisphere. These areas have been known to be affected, both in the

present day, and in its geological history by deglaciation. As the ice melted, the buried

sediment was uplifted due to an isostatic rebound. The methods and knowhow

commonly applied for exploring and developing hydrocarbons in more common areas

without such the effect, may not be sufficiently used to apply in the uplifted areas. This

study realized a potential to fill in the knowledge gaps with laboratory experiments,

required for better understanding stress release affecting acoustic and petrophysical

data. These data are important input parameters used during basin modelling, and

seismic and rock physics modelling studies for resource evaluation. This thesis details

the impact of stress reduction on acoustic and rock physical properties of sands, as

sandstone is commonly found as a reservoir rock. Laboratory experiments were

designed with a stress condition simulating complex burial history, which mimic

episodes of uplift and erosion. Measurements of acoustic, petrophysical, and

geomechanical properties were employed, followed by an investigation of a rock

physics model, suitable for modelling these types of sediments under these stress

conditions. The impact of preconsolidation stress prior to the sediments being uplifted

on time-lapse seismic responses due to fluid saturation changes was also explored. The

research outcome contributes to the knowledge required to de-risk hydrocarbon

prospect evaluation and field development strategies. On a separate but related subject,

this thesis also outlines the results from laboratory measurements and from theoretical

mixing average model calculations performed on pure brine, pure crude oil, and

mixtures of both. These are the main reservoirs fluids commonly found to fill in porous

reservoir rocks, in addition to gas. Modelling various fluid replacement and substitution

scenarios is a common process required both in hydrocarbon exploration and

development phases. This analysis is to validate whether the observations from seismic

responses or well logs show hydrocarbon potentials. During the modelling process,

especially when the pore fluids are filled with more than one fluid type, the estimation

of density and bulk modulus properties of the fluids is required. Therefore, it is

important to verify whether fluid property estimation from the model calculations and

from direct measurements give a good fit. This research outcome can contribute to

better parameters constraints during fluid substitution exercises. It can also aid in

reducing risks during hydrocarbons in-place and remaining oil reserve evaluations.

Page 4: Acoustic, elastic and physical properties of

ii

Page 5: Acoustic, elastic and physical properties of

iii

Preface

This dissertation entitled “Acoustic, elastic, and physical properties of

overconsolidated sands and reservoir fluids – Experimental measurements,

modelling, and implications for reservoir characterization, time-lapse seismic

monitoring, and geomechanics”, has been submitted to the Department of

Geosciences at the University of Oslo, in agreement with requirements for the degree

of Philosophiae Doctor (Ph.D.). This research is part of the Barents Sea Rock

Properties (BarRock) project. Funding for the project was provided by the Research

Council of Norway under the Petromaks program.

The experimental compaction study herein was based on natural sand samples, which

were laboratory prepared, in a condition simulating normal subsurface burial and uplift.

Well log data from the Barents Sea were also included to compare experimental results

with measured data. The fluid experiments were based on NaCl solution and a crude

oil sample from the Draugen Field, North Sea.

The thesis consists of two parts: Part I) an overview, and Part II) the embedded papers.

The overview in Part I is comprised of four chapters. Chapter 1 details the research

opportunities, and the main objectives of the study. Chapter 2 describes the scientific

background of the work. Chapter 3 contains the summaries of the four papers enclosed

in Part II of the thesis. Chapter 4 represents the concluding remarks. Four individual

papers, together with five expanded abstracts and two local conference abstracts are

included in Part II. Three papers have been published in the scientific journal

Geophysical Prospecting, whilst the fourth paper is in preparation for submission to the

peer reviewed journal Marine and Petroleum Geology.

A primary objective of this research was to perform experimental measurements and

modelling of unconsolidated sands, to gain a better understanding of the effects of

stress reduction on rock properties. A secondary objective, which constitutes a minor

portion of the research, was to compare experimental measurements of fluid properties

on crude oil and brine mixtures with theoretical mixing average models. The results of

the primary objective are captured in Papers 1, 2, and 3, whereas the results from the

secondary research objective can be found in Paper 4. Expanded abstracts 1, 2, and 3

were derived from the above-mentioned primary objective papers. The same dataset

was also used to develop expanded abstracts 4 and 5, to investigate the effects of

depletion on reservoir compressibility and compaction, and stress reduction induced

anisotropic effects, respectively. These five expanded abstracts were presented at

international conferences and workshops in the United Kingdom (x 2), the United States of

Page 6: Acoustic, elastic and physical properties of

iv

America, The Netherlands, and Switzerland. The two additional short abstracts that were

presented at local conferences in Norway are also included.

Sirikarn Narongsirikul, Stavanger, February 2020

Page 7: Acoustic, elastic and physical properties of

v

Acknowledgements

I am grateful to my main supervisor, Associate Professor Nazmul Hague Mondol, for

his supervision and dedication. His positive attitude and strong belief gave me a

constant drive, especially on days when motivation ran low. I am also indebted to

Professor Jens Jahren, my secondary supervisor, for all the constructive feedback he

has provided.

This experimental study will not be possible without support from the many staff at

Norwegian Geotechnical Institute (NGI). I would especially like to express my

gratitude towards Toralv Berre, for his time and dedication to teach me how to use

laboratory equipment. His patience is only exceeded by his passion, shown by his

guidance and assistance on many countless weekends and nights. I also thank Magnus

Soldal, Gundmund Havstad, Lars Grande, Reidar Øtter and Manzar Fawad for

discussions and assistance during my time at the laboratory.

I would like to thank my colleagues and friends at the Department of Geosciences

(UiO); Oluwakemi Ogebule, Mohammad Koochak Zadeh, Aatisha Mahajan, and

Mohsen Kalani for making my time at the university the most enjoyable and

memorable experience with their warm friendship.

I am thankful to my ConocoPhillips supervisor, Per Gunnar Folstad, for his continued

support and understanding, allowing me the time to complete the Ph.D. I appreciate

the opportunities ConocoPhillips management provided, especially Jonathan Scorer,

Olaf Knoth, Tim Austin, and Ole Eeg, for onboarding me with ConocoPhillips. I

would also like to thank my colleagues and friends in the Geophysical Implementation

team. Thank you, Marie Skadberg, Leila Bencherif-Sørensen, Lisa Mai, Janne Helen

Kristensen, Victoria Flatås, Alena Chernyshova, Samad Shokouhi, Reidar Midtun, and

Anna Oleksiak (AkerBP) for their friendship and perseverance, especially with never

giving up asking about my Ph.D. progress.

This journey has given me the opportunity to meet with many rock physics

professionals in both academia and industry. I would like to thank many people in the

Rock physics community whom I have crossed path with during conferences and

workshops, and for all those fruitful discussions that have arisen. Special thanks to

Dave Dewhurst for your friendship, support and help to improve the first draft of my

papers.

Page 8: Acoustic, elastic and physical properties of

vi

Finally, I would like to immensely express the appreciation for my family. This journey

would have been difficult without having you all by my side. Mum and Dad, Bung,

Bruce, A, Yada, and Kik, you have always sent your love and support from continents

on the opposite side of the world. Special thanks to Richard Bruce Ainsworth to help

me on the final effort of the dissertation. Robbie, you are beyond any words I could

find to express my thankfulness for. You have been there from day one. Thank you

for caring, supporting, and encouraging me to have this day. Thank you for always

being there for me.

Page 9: Acoustic, elastic and physical properties of

vii

Contents

Abstract ..................................................................................................................................... i

Preface ..................................................................................................................................... iii

Acknowledgements ................................................................................................................. v

List of Scientific Contributions ............................................................................................ ix

Part I Overview ....................................................................................................................... 1

Chapter 1 - INTRODUCTION ...................................................................................... 3

1.1 Research Opportunities .......................................................................................... 4

Chapter 2 - SCIENTIFIC BACKGROUND ................................................................ 7

2.1 Mechanical Compaction of Sands ......................................................................... 7

2.2 Uplift and Erosion ................................................................................................... 9

2.3 Effective Stresses ................................................................................................... 11

2.4 Normal Consolidation versus Overconsolidation ............................................. 13

2.5 Acoustic and Rock Physical Properties of Sandstones and Effects of

Mineralogy and Micro-textures .................................................................................. 14

2.6 Rock Physics of Mechanically Compacted Sands - Friable Sand Model ........ 15

2.7 Reservoir Fluids ..................................................................................................... 17

2.8 Gassmann Fluid Substitution ............................................................................... 19

2.9 Fluid Mixing Averages .......................................................................................... 20

Chapter 3 - SUMMARY OF THE PAPERS ............................................................... 23

3.1 Paper 1 ..................................................................................................................... 23

3.2 Paper 2 ..................................................................................................................... 24

3.3 Paper 3 ..................................................................................................................... 26

3.4 Paper 4 ..................................................................................................................... 27

Chapter 4 - CONCLUDING REMARKS ................................................................... 29

4.1 Future Research Outlook...................................................................................... 31

Bibliography .......................................................................................................................... 33

Part II Enclosed Papers ...................................................................................................... 39

Page 10: Acoustic, elastic and physical properties of

viii

Page 11: Acoustic, elastic and physical properties of

ix

List of Scientific Contributions

Part II of this thesis includes the following four journal papers, five conference

proceedings, and two conference abstracts:

Journal Papers

Paper 1: Narongsirikul, S., N. H. Mondol, and J. Jahren, 2019, Acoustic and

petrophysical properties of mechanically compacted overconsolidated sands: Part 1 –

Experimental results, Geophysical Prospecting, Vol 67, No 4, May 2019, pp. 804 – 824, DOI:

10.1111/1365-2478.12744

Paper 2: Narongsirikul, S., N. H. Mondol, and J. Jahren, 2019, Acoustic and

petrophysical properties of mechanically compacted overconsolidated sands: Part 2 –

Rock physics modelling and applications, Geophysical Prospecting, Vol 67, No 1, January

2019, pp. 114 – 127, DOI: 10.1111/1365-2478.12692

Paper 3: Narongsirikul, S., N. H. Mondol, and J. Jahren, 2020, Effects of stress

reduction on geomechanical and acoustic relationship of overconsolidated sands,

Geophysical Prospecting, Vol 68, No 3, March 2020, pp. 968 – 981, DOI: 10.1111/1365-

2478.12902

Paper 4: Narongsirikul, S., N. H. Mondol, and J. Jahren, Acoustic measurements of

brine – oil mixtures: Experiments versus modelling of the mixing law. In preparation for

submission (2020)

Expanded Abstracts/Conference Proceedings

Expanded Abstract 1: Narongsirikul, S., N. H. Mondol, and J. Jahren, 2013,

Density/porosity versus velocity of overconsolidated sands derived from experimental

compaction. Conference Proceedings, 75th EAGE Conference & Exhibition, London, UK, June

2013, We 06 10, DOI: 10.3997/2214-4609.20130755

Page 12: Acoustic, elastic and physical properties of

x

Expanded Abstract 2: Narongsirikul, S., N. H. Mondol, and J. Jahren, 2013,

Acoustic, electric, and petrophysical properties of mechanically compacted sands of

varying mineralogy– simulating the effects of uplift on rock properties. Conference

Proceedings, Second International Workshop on Rock Physics (2IWRP), Southampton, UK, August

2013, DOI: 10.3997/2214-4609-pdb.381.Narongsirikul_et_al

Expanded Abstract 3: Narongsirikul, S., N. H. Mondol, and J. Jahren, 2013,

Possible application of friable sand model for shallow mechanically compacted

overconsolidated sands. Conference Proceedings, SEG Technical Program Expanded Abstracts

2013, Houston, USA, pp. 2701-2706, DOI: 10.1190/segam2013-1303.1

Expanded Abstract 4: Narongsirikul, S., N. H. Mondol, and J. Jahren, 2013,

Depletion-induced reservoir compaction in shallow overconsolidated reservoir.

Conference Proceedings, EAGE – International Workshop on Geomechanics and Energy,

Lausanne, Switzerland, November 2013, We 01 05, DOI: 10.3997/2214-4609.20131970

Expanded Abstract 5: Narongsirikul, S., N. H. Mondol, and J. Jahren, 2014,

Velocity anisotropy of unconsolidated sands and its relation to induced stress response.

Conference Proceedings, 76th EAGE Conference and Exhibition 2014, Amsterdam, The

Netherlands, June 2014, Tu P11 16, Volume 2014, p.1 – 5, DOI: 10.3997/2214-

4609.20140937

Conference Abstracts

Conference Paper 1: Narongsirikul, S., N. H. Mondol, and J. Jahren, 2013,

Experimental insight into uplift effects on seismic velocities and petrophysical

properties of sandstones: Implication for the Barents Sea area. NGF Winter Conference,

Oslo, Norway

Conference Paper 2: Narongsirikul, S., N. H. Mondol, and J. Jahren, 2013, Rock

Physics aspects of uplifted sediments - Experimental compaction study, NFiP Seminar,

Stavanger, Norway

Page 13: Acoustic, elastic and physical properties of

1

Part I

Overview

Page 14: Acoustic, elastic and physical properties of

2

Page 15: Acoustic, elastic and physical properties of

3

Chapter 1

INTRODUCTION

Transformation from seismic or sonic wave propagation data to rock properties is an

important technique for resource evaluation in hydrocarbon exploration and

development. To invert seismic or well log data to reservoir properties, rock physics

modelling is required to link the two domains, as it provides model equations that relate

seismic to reservoir parameters. Together with traditional and advanced seismic

techniques in acquisition & processing to improve seismic imaging, rock physics

modelling allows subsurface interpretation to move from being qualitative to being

more quantitative. Such quantitative interpretation techniques are important processes

to extract reservoir properties (e.g. porosity, saturation, pressure) from seismic data.

These reservoir properties can readily assist in evaluating resource potential during

hydrocarbon exploration and appraisal. When the hydrocarbons are proven economic

and further developed, additional reservoir properties including geomechanical and

elastic parameters (e.g. compaction, pressure depletion, strain, elastic modulus) are also

commonly obtained from seismic data. This reservoir information can be used to

manage and reduce uncertainties of the remaining reserves.

The rock physics modelling process can range from a simple estimation of rock

properties from porosity-stress relationships, to establishing sediment compaction

trends (e.g. for an application of basin modelling), in a general two or three-

dimensional (2D and 3D) regime, to a more sophisticated fourth dimension (4D),

which involves additional reservoir parameters. The latter often have an application in

time-lapse (4D) seismic monitoring, which involves a computation of changes in

pressure and fluid saturation in the rock through time. This modelling process

commonly uses the Gassmann fluid substitution workflow (Gassmann, 1951; Smith et

al., 2013). This advanced step in the seismic modelling extends the need to estimate,

not only the rock properties, but also pore fluid parameters. A good estimation of such

properties requires the understanding of the relationship between the acoustic

parameters, stress and temperature.

Page 16: Acoustic, elastic and physical properties of

4

1.1 Research Opportunities

The section below details two areas of research (A and B) that are explored in this PhD

thesis via four research themes (1 to 4).

A) Rock property estimation often involves obtaining acoustic and petrophysical

parameters, which vary with depth or stress. In any given sedimentary basin, sediments

commonly experience progressive burial, where the increasing vertical load leads to

compaction. Such a process induces changes in rock properties, such as density and

velocity due to several mechanical dynamics, for example sliding, realignment, crushing

and breakage of grains, which resulting in a loss of porosity. Various experiments have

been designed to capture normal compaction processes, which allow acoustic,

mechanical, and petrophysical measurements to be acquired (Yin, 1992; Tao et al., 1995;

Dvorkin and Nur, 1996; Chuhan et al., 2003; Pettersen, 2007; Mondol et al., 2010;

Fawad et al., 2011; Grande et al., 2011; Bhuiyan et al., 2013). Therefore, laboratory

studies have been maturing in this area. In many sedimentary basins, the sediments

have not only experienced simple progressive burial, but have also undergone more

complex burial histories, like those found in uplifted basins. The process where the

sediments are deeply buried and later uplifted, can result in changes to the stress,

pressure and temperature regimes developed during initial burial. This can often mean

that the rocks appear “overconsolidated” for the depth at which they are encountered.

Several study approaches were developed in the past few years with the aim of

understanding the effects of complex burial histories or stress paths on rock properties,

for example using rock physics modelling (e.g. Grude et al., 2013; Draege et al., 2014;

Avseth and Lehocki, 2016). However, the understanding of the impact of stress

unloading on rocks is still immature. More experimental studies are required to better

estimate and understand rock properties (e.g. geophysical, petrophysical and

geomechanical properties) of sediments undergoing stress unloading during uplift.

Experiments designed to replicate complex burial and uplift histories can lead to

improvements in the interpretation of subsurface data and the characterization of

unloaded rocks. The outcome of these studies is relevant for basin modelling, reservoir

characterization, seismic inversion and time-lapse seismic monitoring applications.

B) Fluid property estimation through seismic forward modelling by Gassmann fluid

replacement (Gassmann, 1951) is a common method to understand different fluid

replacement scenarios in the subsurface. Gassmann’s equation comprises of the

saturated rock bulk modulus term that relates the bulk modulus of the porous rock

frame, the bulk modulus of the mineral matrix, and the bulk modulus of the pore-

filling fluids to its porosity. The equation requires that each of the properties are

independently estimated including the pore-filling fluids. If a fluid is fully substituted

by another, the modelling is straight forward as the averaging of mixed fluids is not

Page 17: Acoustic, elastic and physical properties of

5

required. However, a single fluid phase following the replacement process does not

commonly exist in the subsurface. For example, porous reservoir rocks that are filled

with oil, will never be fully, and homogeneously replaced by brine during water

flooding (Bahadori, 2018). Traces of oil will be left behind, which explains why oil

recovery never achieves 100%. This phenomenon means that the fill of the pore space

in a fluid saturated rock is often a mixture of oil, gas and brine. Estimating the

properties of mixed fluids requires knowledge of the property of the fluid end members

(e.g. bulk modulus and density). Average fluid mixing models (Reuss, 1929 or Voigt,

1910) can then be used to estimate the combined fluid constituents. Most of the fluid

properties are commonly acquired from laboratory measurements. These experiments

have been reported to directly measure through the fluids without saturating core

samples with the fluids (e.g. Han and Batzle, 2000). Many previous fluid experiments,

to measure acoustic velocity and density, were performed for miscible mixtures (e.g.

Han and Batzle, 2000; Han and Sun, 2013; Dashti and Riazi, 2014). A miscible mixture

is when fluids form a homogeneous mix when added together (e.g. decane mixed with

hexane, or CO2 mixed with oil). Experiments to directly measure immiscible fluid

mixtures (e.g. oil mixed with brine) are limited. An immiscible mixture is when fluids

do not form a homogeneous mixture when mixed with another, and the mixed fluids

remain in their own phases. Unlike miscible fluids, most acoustic wave measurements

for the immiscible mixtures were performed by a porous core sample being saturated

with one fluid and flooded or replaced by another (Alemu et al., 2013). The measured

mixed fluid properties (bulk modulus, density) are then later extracted from the

measured fluid saturated core using the Gassmann’s equation. Each fluid’s properties

are then derived using the averaging models through various saturation assumptions

and iterations that provide a good fit. A lack of laboratory measurements of acoustic

waves propagating directly through these immiscible fluids can be because seismic

waves propagate through saturated rocks instead of being through a fluid alone in the

subsurface. Therefore, experimentally measuring rock samples saturated with fluids is

a more natural way of replicating the true subsurface conditions. However, the

estimation of mixed fluid properties by the mixing average laws required in Gassmann

must be done separately prior to combining with the rock property term. It is therefore

important to validate whether fluid property estimation for immiscible fluids using the

existing mixing models gives a good match with direct measurements from

experiments on pure immiscible fluids without rock samples. The Gassmann fluid

substitution independent mixed fluid property term can then be further tested and

verified.

The two areas highlighted above provide the following four research opportunities that

have been addressed in this PhD thesis:

Page 18: Acoustic, elastic and physical properties of

6

1) Experimental measurements of acoustic and petrophysical properties

simulating uplifted sediments: This research study is devoted to understanding

the influence of stress reduction on acoustic velocities, and rock physical properties

based on laboratory experiments. The experiments were designed to simulate a

complex burial history with periods of uplift. The experimental measurements

acquiring compressional (P) and shear (S) wave velocities and corresponding

petrophysical (porosity and density) properties of unconsolidated natural sands

were reported. The sand samples vary in mineral compositions and textures. The

research outcome resulted in Paper 1.

2) Rock physics modelling of acoustic and petrophysical properties of stress

unloaded sediments: This research topic aims at modelling experimental

measurements of overconsolidated sand data from Paper 1. Existing rock physics

models and templates were utilised to evaluate the model validity for the

overconsolidated data. The study focuses on extending the validity of a friable sand

model, which was established for normally compacted sediments, to describe

behaviours of overconsolidated sediments. The effect of overconsolidation due to

stress unloading on fluid detection sensitivity for time-lapse seismic applications,

was also assessed. The research outcome resulted in Paper 2.

3) (Geo)mechanical properties of overconsolidated sands and its relation to

acoustic properties: As a continuation from the first two research themes, this

topic is focused on investigating the relationship between geomechanical (e.g.

stress, strain, Young’s modulus) and acoustic parameters from the same sand

experiments performed in Paper 1. Correlation between static and dynamic elastic

properties of stress unloaded sands was investigated, to improve the knowledge

applying to seismic-geomechanics applications for reservoirs undergoing stress

reduction. The research outcome resulted in Paper 3.

4) Experimental measurements of acoustic properties versus mixing average

law models for brine – crude oil mixtures: This research topic has the key

objectives of acquiring density and velocity measurements of two pure fluids and

three mixed fluids, together with their computed bulk modulus. The measurements

were performed directly through the fluids without saturating in a core sample. The

measurement results were also compared with theoretical mixing average models,

which are commonly used during Gassmann fluid substitution. The experiments

were performed under a range of pressures and temperatures. Pure brine and crude

oil from the Draugen field, North Sea were tested. Three brine - crude oil samples

were prepared by mixing different volume fractions. The research outcome resulted

in Paper 4.

Page 19: Acoustic, elastic and physical properties of

7

Chapter 2

SCIENTIFIC BACKGROUND

This thesis section aims to provide contextual information and scientific background

relevant to the current study. Previous studies, processes and model equations are

briefly described. The section also gives a broad context on how the current study can

contribute to the knowledge gaps in this research area. The main part of the scientific

background focuses on unconsolidated sand compaction and its petrophysical and

acoustic parameters. Definitions of some terms and properties related to mechanical

compaction and uplift process are described. The minor part of this section comprises

a review of reservoir fluid properties and existing models used to estimate mixed fluid

properties.

2.1 Mechanical Compaction of Sands

Newly deposited sands in sedimentary basins can be transformed into hard rocks

through mechanical and diagenetic processes when compacted during burial to deeper

depths. Such transformation increases rock strength as the sediments are subjected to

more stress, and higher temperatures through time (Bjørkum et al., 1998; Bjørlykke and

Jahren, 2010). Compaction starts when the vertical weight of the overlying sediment

increases. This process results in porosity reduction and increased density. Compaction

can be divided into mechanical and chemical domains depending on burial

temperatures (Bjørlykke and Egeberg, 1993; Bjørlykke and Jahren, 2010). In basins

with an average normal geothermal gradient (35°C/km), the transition from

mechanical to chemical compaction takes place at around 2000–2500 m corresponding

to 80–100 °C (Bjørlykke and Egeberg, 1993), (Figure 1). Beyond these depths,

dissolution of quartz takes place at stylolites and precipitation of quartz cement,

sourced from the dissolution at the stylolites, thereby form as overgrowths on the

detrital quartz grains found in between the stylolites increasing the rock strength

(Oelkers et al., 1992, 1993; Bjørkum, 1996; Oelkers et al., 1996; Walderhaug, 1996). The

chemical compaction process leads to a stiffer rock that becomes mechanically pseudo-

over-consolidated and therefore insensitive to the effective stress (Bjørlykke and Høeg,

1997; Storvoll et al., 2005).

Page 20: Acoustic, elastic and physical properties of

8

Figure 1. Schematic drawing to demonstrate compaction in sands (modified from Bjørlykke and Jahren,

2010). At approximately 80-100 C, mechanical compaction is transitioned into the chemical compaction domain.

The mechanical compaction process in sands causes grain realignment and an increase

of grain-to-grain contacts. Significant grain breakages and fracturing can occur when

subjected to high stress. Experimental mechanical compaction for unconsolidated

sands has the potential to reduce porosity of coarse-grained sands from 40-43% to

approximately 26% (Chuhan et al., 2002; Fawad et al., 2010). Finer sand grains are

compacted less because small grain sands have a higher number of grain-to-grain

contacts. Such phenomenon leads to less stress distribution per grain contact, resulting

in less grain fracturing and smaller porosity loss (Figure 2). The knowledge of sand

compaction and collection of laboratory measurements allows compaction trends to

be developed (Chuhan et al., 2002; Fawad et al., 2010). Such knowledge permits us to

determine the sediment’s acoustic (velocity) and petrophysical properties (porosity,

density, and permeability) as a function of stress and temperature. The current uses of

compaction trends are typically for basin and seismic modelling. Since the compaction

curve is different for each lithology, the trend can be used to lithofacies modelling. Any

deviation from normal mechanical compaction trends can be an indication of

cementation (Avseth et al., 2005), over pressure and the presence of gas (Prasad, 2002),

or uplift and erosion (Alchalabi and Rosenkranz, 2002). The latter is a focus of this

thesis.

Page 21: Acoustic, elastic and physical properties of

9

Figure 2. Experimental compaction of loose sand grains (after Chuhan et al., 2002). Coarse-grained sand is more compressible than fine-grained sand.

2.2 Uplift and Erosion

Earth surface uplift due, for example, to tectonic or isostatic deglaciation rebound leads

to enhanced elevations, which in turn induce erosion. In petroleum provinces, uplift

and erosion can have important consequences for petroleum systems. It can affect the

reservoir quality, seal integrity, source rock maturity/migration and reservoir pressure

(Henriksen et al., 2011). Examples of prolific petroleum provinces that have been

uplifted are listed in Table 1.

Riis and Jensen (1992) used the term “net uplift” for estimation of reduced burial

depth. The reduction of a burial depth or overburden removal results in a decrease in

stress on pre-compacted rocks. The resulting effect of stress unloading often shows

on well logs, where anomalously high density and velocity are observed at any given

depth. Figure 3 demonstrates anomalous density and velocity behaviours of well logs

from uplifted compared to non-uplifted areas in the Barents Sea.

Porosity (density)/depth and velocity/depth trends have been used to quantify the

amounts of uplift and erosion (e.g. Hansen, 1996; Densley et al., 2000). As compaction

Page 22: Acoustic, elastic and physical properties of

10

is mostly inelastic and causes permanent damage to sediment grains, only a small elastic

part of the deformation is reversible during unloading. This means that porosities and

velocities resemble the maximum burial depth (maximum effective stress) the

sedimentary rock has experienced. The rocks are said to be “overcompacted” for their

current depth. Higher velocities or lower porosities than expected for a given burial

depth can therefore be used to predict the difference between present day and

maximum burial depth.

Table 1 List of some important uplifted Petroleum provinces (modified from

Henrikson et al., 2011)

Figure 3. Narongsirikul et

al. (2013) shows well log

data from the Barents Sea

comparing P-wave

velocity and bulk density

data between an uplifted

(red curves) and non-

uplifted wells (black

curves).

Basin Country Timing uplift Nature of uplift

San Juan USA Late Eocene–Recent Epeirogenic–isostatic

Permian USA Cretaceous–Recent Epeirogenic–isostatic

Maracaibo Venezuela Early Miocene- Late Eocene Orogenic

Zagros Foreland Iran Miocene–Recent Orogenic

Jungar China Miocene–Recent Orogenic

Western Canada Canada Oligocene-Recent Post Orogenic,

Epeirogenic–isostatic

Timan Pechora Russia Miocene/Pliocene Orogenic–isostatic

Barents Sea Norway Palaeogene and Neogene Orogenic–isostatic

Page 23: Acoustic, elastic and physical properties of

11

2.3 Effective Stresses

The term effective stress, σ’, is defined as σ’ = σ – αp; where σ is the total stress, and

subscript v and h are the direction of applied vertical load and the applied cell pressure

controlling horizontal deformation, respectively; p is the pore pressure; and α is an

effective stress coefficient. Following Terzaghi (1943) the effective stress coefficient

value is assumed to be 1 for unconsolidated sediments.

Effective vertical stress (σ'v) is the difference between vertical total stress (σv) and pore

fluid pressure (p), with unit in Pascals and can be expressed by (Bjørlykke and Jahren,

2010):

σ'v = σv − p (1)

The vertical total stress or the lithostatic stress (σv) can be calculated by integrating the

density (ρb) over the depth of the sediment column (h)

σv=ρbgh (2)

where ρb is the average sediment bulk density in g/cm3 of the overlying sediment layers,

h is the sediment thickness in meters and g is the acceleration of gravity in m/s2.

Pore fluid pressure (p) is the pressure in the pore fluid. Like a total stress, pore fluid

pressure, at any given depth, can be calculated by integrating the weight of density of

the water from sea level to a depth of interest. Pore fluid pressure can be defined by:

p=ρwgh (3)

where ρw is the average fluid density of the fluid that filled the sediment column in

g/cm3, h is the sediment thickness (meters) and g is the acceleration of gravity (m/s2).

The transmission of stress through the grain framework helps carrying the weight of

the overburden stress (Figure 4). Therefore, the effective stress is sometimes called the

intergranular stresses (Bjørlykke and Jahren, 2010). This is the effective stress that

governs the mechanical compaction of sediments.

Page 24: Acoustic, elastic and physical properties of

12

Figure 4. Schematic representation of the effect of stress from overburden which is carried by mineral grain framework and the pore pressure. The effective stress is defined as the difference between overburden and pore pressure.

The effective stress in the horizontal direction (σ’h) is defined as total horizontal stress

(σh) minus the pore pressure (p). The horizontal stress is customarily expressed as a

proportion of the vertical stress.

σ’h = K0 σ’v (4)

where K0 is a coefficient of earth pressure at rest in terms of effective stresses.

K0 can be used for an estimation of the effective horizontal stress as well as an

indication of stress anisotropy. K0, also usually defines a stress path for the sediment

loaded under a uniaxial strain condition in soil mechanics. K0 is the ratio between

horizontal effective stress σ’h and vertical effective stress σ’v, that is (Mayne and

Kulhawy, 1982):

K0 =σ’h

σ’v (5)

For example, for K0 = 1, the stress is hydrostatic implying that the vertical and

horizontal effective stresses are equal. For K0 < 1, the stress is uniaxial, and the vertical

effective stress is greater than the horizontal effective stress. For loose sediments

loaded under normal consolidation, K0 equals approximately 0.5.

Effective horizontal stresses are important for drilling applications. The magnitude of

horizontal stresses is critical for a mud weight design, whist horizontal stress direction

is important for well trajectory optimization to improve wellbore stability. Mud weight

used to counteract subsurface pore pressure should not exceed the minimum

horizontal stress to prevent drilling fluid losses into the formation through reopening

of existing fractures or invasion in permeable formations.

Page 25: Acoustic, elastic and physical properties of

13

The accuracy in the estimation of the horizontal stresses is normally improved using

calibrated data from Leak Off Tests (LOT) and borehole breakouts (Zoback et al.,

1985). However, these data are not available in a new frontier exploration, as such data

can only be obtained when drilling a well. Therefore, being able to estimate the

horizontal stresses through the knowledge of K0 relationship with effective vertical

stress, can reduce drilling risks in a new frontier basin.

2.4 Normal Consolidation versus Overconsolidation

Normal consolidation refers to a continuous stress increase during sediment

compaction. Normal consolidation, normal compaction, normal loading, and virgin

compaction are terms used interchangeably. Sediments presently found at lower stress

than previous experienced levels due to uplift, erosional unloading or excess pore

pressure, are termed “overconsolidated” (e.g. Pender, 1978; Bjørlykke and Jahren,

2010). Alternative terms describing overconsolidation are unloaded, reloaded, or

uplifted. When the sediments have been precompacted before uplift, the term

“preconsolidation stress” is used for the past maximum effective stress level acting on

the sediments prior to stress reduction.

When describing the degree of overconsolidation, the overconsolidation ratio (OCR)

is used. OCR is defined by the difference between past maximum effective vertical

stress σ’v max, and the present effective vertical stress σ’v, (Casagrande, 1936) that is:

OCR =σv max

σv ′ (6)

When the OCR is equal to 1, sediments are normally consolidated, i.e. when the past

maximum effective stress and the present effective stress are equal. When OCR is a lot

greater than one, the sediments are previously buried at significantly deeper depth than

at present day depth.

Mayne and Kulhawy (1982) determined the relationship between K0 and

overconsolidation ratio (OCR) on the effect of stress history by compiling data from

hundreds of different soils. The relation is expressed as:

K0oc = (1- sin θ') OCR sin θ’ (7)

The analysis of unloading stress linking K0 and OCR are built on Jaky’s simplified

equation (Jaky, 1948). For normally consolidated materials, OCR=1, and equation (7)

reduces to K0oc= (l-sin θ') = K0nc, where K0nc is K0 during normal compaction. The θ

Page 26: Acoustic, elastic and physical properties of

14

is frictional angle. For example, if θ’ = 30°, K0nc = 0.5. If we relate K0oc with effective

horizontal and vertical stresses in equation 4 and 5, the effective horizontal stress in

uplift basins can be determined.

2.5 Acoustic and Rock Physical Properties of Sandstones and Effects of

Mineralogy and Micro-textures

Acoustic, moduli and petrophysical properties of sands and sandstones are controlled

by grain contacts, mineralogy, and microtextures such as sorting and grain size. These

controlling parameters affect initial porosity. How much the porosity is further reduced

depends on the degree of applied stress and temperature following burial and

compaction. At a certain stress and temperature (as reviewed in section 2.1), the sands

begin to increase in stiffness due to quartz precipitation and cementation.

Experimental compaction studies of unconsolidated sands or sandstones have

previously been documented to show the effect of stress dependent velocities and

porosity (Yin, 1992; Tao et al., 1995; Dvorkin and Nur, 1996; Chuhan et al., 2003;

Pettersen, 2007; Mondol et al., 2010; Fawad et al., 2011; Grande et al., 2011; Bhuiyan et

al., 2013). The measurements were performed for samples with different mineral

compositions and textural variations. Varying mineralogy and textural differences

(such as sorting, grain size, and roundness) also has the effect on both acoustic and

petrophysical properties (Chuhan et al., 2002; Fawad et al., 2011). Sorting is one of the

most common microtextural properties that influences the mentioned parameters

(Dvorkin and Nur, 1996; Fawad et al., 2011). Fawad et al. (2011) demonstrates that

sand samples with poor sorting shows low initial porosity, while sands with high grain

angularity has high initial porosity. Compaction and velocities increased with

decreasing degree of sorting. However, at the same porosity under low stresses, the

velocities of the composite mixture with varying grain size distribution were slightly

lower than in the well-sorted sands. This indicates that the presence of loose smaller

grains in-between the framework grains contributes to porosity reduction but does not

form grain-to-grain contacts contributing to velocity increase under low stresses.

Fawad et al.’s (2011) sand compaction tests also show that at a given stress, well-sorted,

coarse-grained sands are more compressible and have higher velocities both for

compressional and shear waves than fine grained sands when the mineralogy is similar.

Such phenomenon can be attributed to grain fracturing, where coarser grains lead to

high compressibility and large grain-to-grain contact areas result in high velocities. On

the other hand, small grain-to-grain contact areas allow higher deformation at grain

contacts, more crushing and increased porosity loss resulting in high velocities. The

mineralogy also influences the velocities. Quartz-poor sands show higher velocities

Page 27: Acoustic, elastic and physical properties of

15

compared to that of quartz rich sands. This could be the result of increasing grain

contacts of ductile minerals in the quartz-poor sands increasing the effective bulk and

shear stiffness (Fawad et al., 2011). Since these sand compaction studies were

performed by applying stress loading to the samples, compaction experiments that

include the opposite stress application by unloading, can improve the understanding

of how the variations in sand compositions and textures in such a stress condition can

impact acoustic and rock physical parameters.

2.6 Rock Physics of Mechanically Compacted Sands - Friable Sand

Model

Rock physics has become an important tool in reservoir geophysics and quantitative

seismic interpretation. To successfully apply rock physics modelling in such an

integrated study, appropriate rock physics models suitable to describe certain rock

types need to be selected. Many empirical site-specific models have been used for

specific settings. However, such models may not be applicable elsewhere due to

differences in geologic settings. At the same time, more advanced physics-based

models can be too uncertain because of poor constraints on the input parameters

without well or laboratory data to calibrate these parameters. A hybrid modelling

approach, which combines those two model types, has been proposed and applied to

siliciclastic unconsolidated to moderately consolidated sediments (Mavko et al., 2009;

Avseth et al., 2010).

In sandstones, a physical-contact theory Hertz-Mindlin (HM, Mindlin, 1949) can be

used to calculate elastic moduli and velocities as a function of porosity and pressure.

This theory can describe the porosity-pressure dependence in any unconsolidated

sediment, when the porosity reduction is only due to mechanical compaction. The

HM model combined with theoretical elastic bounds such as the Hashin-Shtrikman

(HS) bounds mimics the elastic signatures of porosity reduction associated with

depositional sorting and diagenesis, including mechanical and chemical compaction.

For unconsolidated sands, such a theoretical combination of models was developed

into a hybrid model called “a friable sand model”.

The friable (unconsolidated) sand model was first introduced by Dvorkin and Nur

(1996). The model expands the possibility of predicting seismic velocity and porosity

of unconsolidated sands if rock microtextures (e.g. sorting) are known (Figure 5).

Without such geometric details the best possible way to predict seismic velocities are

by using upper, and lower bounds, and the geometric average (Voigt, 1910; Reuss,

1929; Hills, 1952; Hashin and Shtrikman, 1962, 1963). Therefore, the model allows

Page 28: Acoustic, elastic and physical properties of

16

better characterization of unconsolidated reservoir sands as it helps to discriminate

between poorly-consolidated and well-cemented rocks.

The friable sand model suggested by Dvorkin and Nur (1996) was based on a dataset

of high porosity, normally consolidated sands from the North Sea. If the sands buried

and uplifted within a mechanical compaction domain where the sands remained

poorly-consolidated after uplift, the model based on compaction, may be used for

uplifted rocks. This PhD study aims at exploring and expanding the validity of the

model for overconsolidated sands.

The friable sand model uses Hertz-Mindlin (HM, Mindlin, 1949) contact theory to

calculate the dry rock moduli at depositional porosity. To connect this porosity end

point with the solid mineral phase, the modified Hashin-Shtrikman lower bound (HS-

, Hashin and Shtrikman, 1963) is employed to interpolate between the two data points.

Figure 5. Schematic representation of the friable sand model. Elastic moduli or velocities increase as grain sorting deteriorates.

The bulk and shear moduli from the HM theory are given by:

𝐾HM = [ 𝑛2(1−𝜙𝑐)2 µ2

18𝜋2(1−𝜈)2𝑃 ]

1

3 (8)

µHM = 5−4𝜈

5(2−𝜈)[

3𝑛2(1−𝜙𝑐)2 µ2

2𝜋2(1−𝜈)2𝑃 ]

1

3 (9)

, and for the modified HS- lower bound are given by:

Page 29: Acoustic, elastic and physical properties of

17

𝐾dry = [

𝜙

𝜙𝑐

𝐾HM+4

3µHM

+1−

𝜙

𝜙𝑐

𝐾+4

3µHM

]

−1

− 4

3µHM (10)

µdry = [

𝜙

𝜙𝑐

µHM+ɀ +

1−𝜙

𝜙𝑐

µ+ɀ ]

−1

− ɀ (11)

ɀ = µHM

6(

9𝐾HM+8µHM

𝐾HM+2µHM) (12)

where KHM and µHM are the dry rock bulk and shear moduli, respectively, at critical

porosity (depositional porosity), ϕc. µ and ν are the shear modulus of the solid mineral

and Poisson’s ratio of the mineral. P is the effective pressure. Based on Terzaghi’s

principle the effective pressure is the difference between total applied pressure and

pore pressure. The pore pressure is assumed to be hydrostatic. n is the coordination

number (the average number of contacts per grain). Kdry and µdry are the dry

unconsolidated sand bulk and shear moduli. K is the bulk modulus of mineral. ϕ is the

porosity.

2.7 Reservoir Fluids

Reservoir fluids fall into three broad categories; (i) aqueous solutions with dissolved

salts (e.g. brine), (ii) liquid hydrocarbons (e.g. crude oil), and (iii) gases (hydrocarbon

and non-hydrocarbon e.g. H2S). Their fluid compositions depend upon their source,

depositional history, and present thermodynamic (temperature and stress) conditions.

Such variations can influence their physical and acoustic properties. Knowledge of

thermodynamic and physical properties of hydrocarbons, including their complex

mixtures, is critical in petroleum production and processing. Accurately predicting

reservoir fluid properties required in seismic interpretation, reservoir monitoring and

direct hydrocarbon indicators (DHI) may be the difference between a successful and

an unsuccessful hydrocarbon discovery and development.

Fluid properties can be calculated through PVT relations and generalized correlations

based on the principles of corresponding states (Equation of State, McCain, 1990;

Danesh, 1998). However, the PVT method involves an isothermal process and is not

appropriate for seismic modelling studies. Batzle and Wang (1992) made an effort in

applying engineering properties to develop the geophysical properties of hydrocarbon

fluids. Their equations are found to be more suitable to estimate fluid properties for

geophysical applications, as the measurements through wave propagation involve an

adiabatic process instead of isothermal.

Page 30: Acoustic, elastic and physical properties of

18

The primary seismic properties of pore fluids; density, bulk modulus, velocity, and

viscosity, vary substantially but systematically under the pressure and temperature

conditions typically seen in sedimentary basins. Brines and hydrocarbon gases and oils

are the most abundant pore fluids.

Based on Batzle and Wang (1992), gas and oil density and modulus as well as oil

viscosity, increase with molecular weight and pressure, and decrease with temperature.

Gas viscosity has a similar behaviour, except at higher temperatures and lower

pressures, where the viscosity will increase slightly with increasing temperature. Oil

density and the bulk modulus are shown in Figure 6 and 7.

Figure 6. Oil densities as a function of temperature, pressure, and composition (Batzle and Wang, 1992)

Figure 7. The bulk modulus of oil as a function of temperature, pressure, and composition (Batzle and Wang, 1992)

Page 31: Acoustic, elastic and physical properties of

19

Brine modulus, density, and viscosities increase with increasing salt content and

pressure. Brine shows anomalous temperature behaviour as the bulk modulus increase

with temperature and reaches a maximum at a temperature between 40 to 80°C (Figure

8). Beyond these temperature points, the brine’s modulus establishes a reversing trend.

It decreases with increasing temperature (Batzle and Wang, 1992).

Figure 8. Calculated brine modulus as a function of pressure, temperature, and salinity (Batzle and Wang, 1992)

2.8 Gassmann Fluid Substitution

Gassmann theory (Gassmann, 1951) is the most used theoretical approach for fluid

substitutions. Gassmann’s equation assumes that the rock is macroscopically

homogenous and isotropic, and that all pores are interconnected. Gassmann’s

assumption is valid only at low frequency to allow pore pressures sufficient time to

equilibrate over the pore space. Gassmann’s model relates the saturated bulk modulus

of the rock to its porosity, the bulk modulus of the porous rock frame, the bulk

modulus of the mineral matrix, and the bulk modulus of the pore-filling fluids using

the following equation:

𝐾𝑠𝑎𝑡

(𝐾𝑜−𝐾𝑠𝑎𝑡)=

𝐾𝑑𝑟𝑦

(𝐾𝑜−𝐾𝑑𝑟𝑦)+

𝐾𝑓𝑙

𝜙 (𝐾𝑜−𝐾𝑓𝑙) (13)

where Ksat is the saturated bulk modulus in the undrained condition, Ko is the bulk

modulus of the mineral matrix, Kfl is the bulk modulus of the pore fluid, Kdry is the

bulk modulus of the dry porous rock frame, and ø is porosity.

Page 32: Acoustic, elastic and physical properties of

20

The equation has two main terms which shows that the application of the Gassmann’s

model is a two-part process. The first term is the part determining the bulk modulus

of the dry porous rock frame, whilst the second term relates to the bulk modulus of

the saturated fluid, which is required to be estimated independently. When there is

more than one fluid mixture, the fluid mixing average law is employed to calculate the

mixed fluid bulk modulus, Kfl.

To convert the elastic moduli to seismic wave velocity, Vp, the following equation is

employed:

Vp = √(K+

4

3μ)

ρ (14)

where K is the saturated bulk modulus (Ksat), μ is the effective shear modulus of rock

with fluid (μsat = μdry), and ρ is bulk density.

2.9 Fluid Mixing Averages

Voigt and Reuss Averages

Theoretical prediction of the effective elastic moduli of a mixture of grains or pores

requires that detailed descriptions of rock microtextures e.g. geometrical details and

volume fraction of the mixtures are known. However, in most cases, such information

is not available. The best possible way to achieve the estimation of the elastic moduli

is to predict upper and lower bounds using Voigt and Reuss averages (Figure 9).

Figure 9. Effect of changing the volume fraction of constituent materials. The bulk modulus will move along the vertical dotted line between the two bounds (after Mavko et al., 2009).

Page 33: Acoustic, elastic and physical properties of

21

At any given volume fraction of considered constituents, the effective elastic modulus

will fall between the upper and lower bounds. The Voigt (Arithmetic) and Reuss

averages are interpreted as the ratio of average stress and average strain within the

composite mixture. The stress and strain distribution in the rock are generally unknown

and are expected to be nonuniform. The upper bound (Voigt, 1910) assumes the strain

is uniform or called iso-strain (Figure 10, left). The lower bound (Reuss, 1929) assumes

the stress is uniform or called iso-stress (Figure 10, right). In a geological context, the

Voigt bound describes an extreme condition that all the sediment grains experience

the same strain when the sediment is compacted and deformed under stress. Whilst,

Reuss describes another extreme (iso-stress) under which all deposited sediments

experience identical stress when forces are applied to the rocks. Since the Reuss average

describes an iso-stress situation, it also applies accurately to suspensions and fluid

mixtures and therefore is an appropriate model used for comparison with fluid

experiments herein.

Figure 10. Schematic representation of Voigt or iso-strain (Left) and Reuss or iso-strain (Right) which are theoretically served as the upper bound and lower bounds, respectively.

The Voigt average of the effective elastic modulus, Mv, of any given number (n) of the

mixed constituents is:

𝑀𝑉 = ∑ 𝑓𝑖𝑀𝑖𝑛𝑖=1 (15)

The Reuss average of the effective elastic modulus, MR, of any given number (n) of the

mixed constituents is:

1

𝑀𝑅 = ∑

𝑓𝑖

𝑀𝑖

𝑛𝑖=1 (16)

where fi is the volume fraction of the ith constituent, and Mi is the elastic modulus of

the ith constituent.

Page 34: Acoustic, elastic and physical properties of

22

Wood’s Equation

In fluid suspension or fluid mixtures, compressional wave velocity is given by Wood’s

equation (Wood, 1955), given that heterogeneities are small when compare with a

wavelength.

𝑉𝑃 = √𝐾𝑅

𝜌 (17)

where VP is P-wave velocity, KR is the Reuss average (iso-stress) of the constituents

1

𝐾𝑅 = ∑

𝑓𝑖

𝐾𝑖

𝑛𝑖=1 (18)

and ρ is the average density which is estimated by Voigt averaging method

𝜌 = ∑ 𝑓𝑖𝜌𝑖𝑛𝑖=1 (19)

fi. Ki, and ρi are volume fraction, bulk moduli and density of the constituent,

respectively.

Page 35: Acoustic, elastic and physical properties of

23

Chapter 3

SUMMARY OF THE PAPERS

This thesis section provides a summary of the four papers. Each paper summarizes

the study objectives and the main findings.

3.1 Paper 1: Acoustic and petrophysical properties of mechanically

compacted overconsolidated sands: Part 1 – Experimental results

Objectives:

This study reports an experimental investigation of seven brine-saturated

unconsolidated sands with varying mineralogical compositions and textures. The study

applies complex stress paths under a zero horizontal strain condition including three

stages of loading, partial unloading and reloading to simulate sediments that underwent

several uplift-erosion episodes. The samples were compacted in the uniaxial strain

configuration from 0.5 up to 30 MPa effective stresses. The measurements reported

here include porosity and ultrasonic P- and S-wave wave velocities. Relationship

between porosity and velocities were assessed. Quantification of changes in the

measured parameters after stress reduction was also included to study the effect of

overconsolidation. The results were compared with previously published sand

compaction datasets and with well logs from the Barents Sea Shelf.

Main findings:

• The experimental results show that stress unloading impacts velocities both P-

and S-waves, porosity, and density of sands differently compared to the sands

subjected normal loading. Lower porosity and higher velocities are found at a

given applied stress in pre-consolidated compared to normally consolidated

sands. This can be explained by deformation of the sediments due to pre-

compaction is being permanent and irrecoverable. Porosity and P- and S- wave

velocities deviate from the normal compaction trends during unloading and the

degree of deviation increases with increasing preconsolidation stress. However,

varying preconsolidation stress magnitudes (maximum applied stress) do not

significantly affect the change in the porosity - velocity relation during

unloading.

Page 36: Acoustic, elastic and physical properties of

24

• The estimation of the amount of porosity loss and the change of P- and S-wave

velocities due to the stress changes during unloading caused the total porosity,

and the two velocities to change a maximum of 5%, 25%, and 50%, respectively.

At any given stress change, these relations can constitute a porosity loss (Δϕ)

and a change in the P-wave velocity (ΔVP) relationship to an approximation of

Δϕ ~ 5ΔVP. This means one unit of porosity loss causes five times P-wave

velocity increase. For the velocities, the 25% and 50% changes of the P- (ΔVP)

and S- (ΔVS) waves as a result of stress reduction gives ΔVP ~ 2ΔVS. This means

the change in S-wave velocity is approximately two times larger than the change

in P-wave at any given stress reduction.

• Mineral composition and sorting are found to influence porosity and velocities

mostly during the normal consolidation, but less significantly during unloading

and reloading. Published experimental data of sand samples with varying

mineral compositions and textures, but subjected to a similar stress condition,

resemble the same unloading trend on the velocity – porosity space as the study

herein.

• The research findings and the dataset can be used in velocity modelling, basin

analysis, time-lapse seismic monitoring, and potential seismic-geomechanics

applications for basins subjected to stress unloading or uplift. The experimental

results are only valid for unconsolidated sands that have been compacted and

unloaded/reloaded within the mechanical compaction domain.

3.2 Paper 2: Acoustic and petrophysical properties of mechanically

compacted overconsolidated sands: Part 2 – Rock physics modelling and

applications

Objectives:

The objective of this paper is to utilize existing rock physics models to evaluate the

validity of the model for the measured data from Paper 1. A friable sand model was

used to describe relationships between velocities and porosity of the laboratory data.

Evaluation of the validity of the friable sand model, established for normally

compacted sediments to describe micro-textural changes, in particular sorting, suitable

to be used for overconsolidated sediments was investigated. Overconsolidated rock

properties in the P-wave and S-wave velocity ratio (VP/VS) and acoustic impedance

(AI) domain were also included in the study. Observations of geological trends utilizing

Rock Physics Templates (RPTs) were discussed. For the purpose of time-lapse seismic

Page 37: Acoustic, elastic and physical properties of

25

applications, fluid saturation sensitivity was also further analysed in order to address

whether 4D fluid detectability can be affected if the rocks experienced pre-

consolidation. Barents Sea well log data were included for comparison with the

experimental measurements.

Main findings:

• The results show that a friable sand model established from normally

consolidated sediments can also be used to predict velocity and porosity of

overconsolidated sediments. At a given stress, the sands which were

subjected to stress reduction, plotted on the same model line as the sands

that experienced normal consolidation. The overconsolidated sand data

plotted towards the lower porosity and higher velocity ends as a result of

the porosity loss due to the effect of preconsolidation. This also means that

the data from overconsolidated sand samples plotting along the friable sand

model lines, not only describes the deterioration of grain sorting as

traditionally explained by other studies, but also outlines the differences in

the preconsolidation stresses associated with a maximum burial depth or

amount of uplift. These findings extend the validity of a friable sand model

and have a potential to be used for estimating preconsolidation stress of

overconsolidated sediments.

• Different behaviours are found for velocity and porosity (density)

relationship and VP/VS versus AI space between overconsolidated sands

and normally consolidated sands. The overconsolidated sands exhibit

steeper velocity-porosity and VP/VS-AI gradients compared to the normally

consolidated sand data. The VP/VS-AI trends established on Rock Physics

Templates (RPTs) also add geological burial history information, such as a

degree of stress reduction, on top of the already existing relations previously

reported.

• The effect of preconsolidation on a 4D fluid detectability for reservoir

monitoring applications was investigated. The result showed that fluid

sensitivity of the overconsolidated sands decreases with increasing degree

of preconsolidation stress prior to the sands being unloaded. The sensitivity

of the VP/VS and AI responding to the fluid replacement is less in

overconsolidated sediments compared to normally compacted sediments.

This implies that there could be a limitation on the change of fluid saturation

detectability if time-lapse seismic monitoring is employed in a field where

the reservoir rocks are affected by uplift. These findings allow better rock

Page 38: Acoustic, elastic and physical properties of

26

physics diagnostics and fluid sensitivity understanding for uplifted

sediments.

3.3 Paper 3: Effects of stress reduction on geomechanical and acoustic

relationships of overconsolidated sands

Objectives:

Experimental investigation to acquire mechanical properties (e.g. stress, strain, and

overconsolidation ratio) measured of the same sand samples as used in Papers 1 and 2

are reported. The mechanical data reported together with acoustic velocity and density

are used to compute static and dynamic elastic moduli. Relationships between rock

acoustic and mechanical properties and the different behaviours between normal

compaction and overconsolidation are investigated. The objective is to compare such

properties between normal and overconsolidated stress conditions in order to

understand the effect of stress reduction. The knowledge obtained can improve the

understanding of seismic – geomechanics inter-relations of overconsolidated

sediments.

Main findings:

• Mechanical parameters including stress, strain, effective horizontal and vertical

stress coefficient or stress path under a uniaxial strain condition (K0), and

overconsolidation stress ratio (OCR) show significantly different relationships

when comparing normal compaction to overconsolidation (unloaded and

reloaded). K0 value of approximately 0.5 is found for most of the normally

consolidated sands, but slightly varies during unloading depending on mineral

compositions and textural differences. The relationship between K0 and OCR

from previously published studies can be further simplified from the K0 and

OCR relation established from the experimental data study herein. Given that

K0 has a direct relationship with an effective horizontal and vertical stress, the

effective horizonal stress can be derived if effective vertical stress is known. An

ability to estimate a horizonal stress is important especially in exploration of a

new frontier basin. This is because more accurate data like Leak Off Tests

(LOT), which is usually tested and acquired during drilling is not available in an

early hydrocarbon exploration phase. The finding is especially helpful to the

estimation of horizontal stress for designing drilling parameters (e.g. mud

weight) in a well being drilled in an uplifted basin.

Page 39: Acoustic, elastic and physical properties of

27

• The investigation results on the elastic moduli of the sand samples (e.g.

constrained modulus, Young’s modulus, bulk and shear modulus) show that the

static moduli, which were computed from stress and strain data, of the

overconsolidated sands are much higher than for normally consolidated sands

as the deformation is smaller (small strain). The correlation between dynamic

(computed from velocities and density data) and static (computed from stress

and strain data) elastic moduli is stronger for the sands subjected to the stress

during overconsolidation stage than for a normal consolidation stage. Such

correlation can be used to transform dynamic elastic moduli which are more

frequently obtained from well log measurements to static elastic moduli which

are rarely obtained from core measurements. The static moduli are the

parameters required for geomechanical model input for many geomechanics

application, for example collapse analysis for wellbore stability understanding.

3.4 Paper 4: Acoustic properties of brine – crude oil mixtures:

Experimental measurements versus mixing average law model

Objectives:

Experimental measurements to acquire density and velocity of two pure fluids and

three mixed fluids were performed under applied pressure up to 35 MPa at three

temperatures 23˚C, 75˚C and 120˚C. Pure brine with NaCl concentration of 35000

ppm and crude oil from the Draugen field, North Sea were tested. Three brine - crude

oil samples were prepared by mixing different volume fractions. The mixtures

constitute of 70% brine - 30% crude, 50% brine - 50% crude, and 30% brine - 70%

crude. The study aims to report measurement results and compare the measured mixed

fluid properties including velocity, density, and calculated bulk modulus with

theoretical mixing average models that are commonly used during the Gassmann fluid

substitution workflow.

Main findings:

• The results show that measured velocity, density, and its bulk modulus product

decrease with increasing temperature and oil volume fraction. The above-

mentioned fluid properties increase with pressure for most samples. However,

pure brine behaves abnormally as the velocity and bulk modulus increase with

increasing temperatures up to 70 C and then decrease as temperature rises to

120 ˚C. This pure brine behaviour is similar to that reported in other published

studies.

Page 40: Acoustic, elastic and physical properties of

28

• Experimental measurements of fluid properties under elevated pressure and

temperature show that most mixed fluids have similar behaviours compared to

the commonly used models, with a few exceptions. Arithmetic averaging model

(Voigt average), matched measured density reasonably well for most samples,

except 100% crude oil at 120˚C, where the model overpredicted the density

especially at a high stress level. Bulk moduli estimated from Wood’s (or Reuss)

averaging methods give a reasonably good match to the values calculated from

the measured density and velocity data for most of the mixed fluids, except the

sample with 70% brine and 30% crude oil. At 75 ˚C and 120 ˚C, Voigt’s mixing

average matches the bulk modulus of this high brine content sample (70-30:

brine-oil mixture) better. However, at 23˚C both Wood’s and Reuss models fail

to capture bulk modulus for this sample. The large discrepancy between the

model and measured moduli is possibly due to anomalous brine temperature

behaviour in mixed samples with a high proportion of brine. The study shows

that careful fluid substitution modelling is required for mixed fluids with high

brine concentrations when estimating bulk modulus of the mix fluid which is

performed as part of the fluid substitution workflow.

Page 41: Acoustic, elastic and physical properties of

29

Chapter 4

CONCLUDING REMARKS

The main scientific contributions of this PhD dissertation were to improve the

understandings of; i) stress reduction affecting rock physical and acoustic properties

and ii) fluid property measurements versus estimation using models. The first

contribution constitutes a major part of the thesis, which combines experimental

measurement, rock physics, and geomechanics studies.

This experimental compaction study of natural sand samples document significant

measurement differences in velocity and density (porosity) between normal

consolidated (stress loading) and overconsolidated (stress unloading) samples. The

degree of the differences increases with preconsolidation stress prior to unloading.

Velocity and porosity loss remained higher after stress unloading when the sands were

subjected to higher preconsolidation stress. This implies that at a given present day

depth, the velocity and porosity of sediments, that undergo different burial depths

before stress release, varies after uplift. By choosing the sand samples with varying

mineralogy and textures, the study can detail the effect of mineral composition and

sorting on stress reduction. This is applicable as in the subsurface various sands with

different mineral and microtextural compositions are found. The change in velocity

and porosity trends, found for the different samples with different mineral constituents

and sorting, is very similar even during the application of stress reduction, while the

trends for the normal compaction varies. This behaviour implies that mineralogy and

sediment textures do not strongly influence acoustic response and porosity of the

sediments in uplifted basins. In one way, this can make a model of stress dependence

acoustic velocities and porosity of stress unloaded sediments less complicated, as the

similar velocity – porosity trend can be established no matter how the sediments’

mineralogy and microtextural (e.g. sorting) properties are constituted. In another way,

the same velocity – porosity trend used for a lithology prediction of overconsolidated

sands subjected to uplift becomes more complex as the work suggests it is difficult to

discriminate the sand types (e.g. sands with varying mineralogies have similar

behaviours). This complexity needs to be accounted for in an uncertainty analysis

during basin modelling studies.

In rock physics studies, a new or modified rock physics model is often proposed when

the measurements of rock or fluid properties are obtained from a laboratory

experiment. In this study however, a new or modified rock physics model equation is

Page 42: Acoustic, elastic and physical properties of

30

not established. An existing rock physics model was used instead with the aim to

investigate and expand its model validity to the sand types that were subjected to stress

release similar to the studied samples herein. A friable or unconsolidated sand model

was utilized. The rock physics modelling result using the friable sand model applied

to the measured sand data shows that the model is not only valid for normally

consolidated but also overconsolidated sands. At any given stress, the overconsolidated

sand data plots along the friable sand model lines, not only describing deteriorating

sorting but also outlining the differences in preconsolidation stresses the sands were

subjected to before unloading. The result expands the validity of the friable sand

model for use in overconsolidated sediments. The model can also provide information

about the degree of preconsolidation stress the sediments were subjected to. These

findings can also pose a level of ambiguity in discriminating sediment sorting properties

from well to poorly sorted using the velocity-porosity relationship. This is because the

sand data plots on the model toward the lower porosity end may be a poorly sorted

normally compacted sand, but can also be a well sorted sand that was previously pre-

compacted at a deeper than present day depth. These important findings mean that

knowing whether a studied basin experienced uplift or no-uplift before a rock physics

modelling exercise is important. The results of the study can help to improve velocity

modelling, basin analysis and seismic modelling studies of shallow reservoir sands

subjected to stress reduction at burial depths where mechanical compaction is the

dominant process.

If a hydrocarbon discovery in an uplifted basin is proven economic and the field is

further developed, it is important to include oil or gas recovery strategies as part of a

field development study. The fluid saturation sensitivity results in this study help to

understand the impact of uplift on the detectability of fluid saturation change (e.g.

water replacing oil). Results show that when fluid saturation changes, reservoir sands

that experienced stress release have reduced acoustic responses and porosity changes.

The higher the pre-compaction depth prior to uplift the lesser fluid sensitivity to the

seismic response which could impact the detectability of fluid saturation changes on

time-lapse seismic. This can mean that another alternative reservoir monitoring

method (e.g. electromagnetic or resistivity time-lapse) needs to be considered as part

of an oil recovery plan during field development studies.

On a separate, but related subject from the above highlighted aspects of stress

reduction affecting rock properties, the final part of the study focuses on laboratory

measurements of fluid properties. The measurements of velocity and density were

performed for both pure and mixed fluids under elevated pressures and temperatures.

Some discrepancies between the measured properties and the model calculation using

fluid mixing average laws are found for one mixed fluid sample, but other mixtures

show a good match between the measured and the modelled properties. Arithmetic

Page 43: Acoustic, elastic and physical properties of

31

averaging model (Voigt average), matched measured density reasonably well. Reuss or

Wood’s averaging method gives a reasonably good match to most mixed fluids, except

for the sample with high brine volume (70:30 brine-crude oil). At low temperature,

Voigt’s model does not capture the measured data of this mixture. The effect of brine

abnormal temperature behaviour for mixed samples with high proportions of brine

possibly explains the discrepancies. The study shows that modelling error due to an

incorrect mixed fluid property estimation using the averaging models can occur at this

step of estimating fluid property during fluid substitution modelling. Careful fluid

substitution exercises are required, and a new mixing average model may be needed.

Since most studied fluid samples show a good fit between the measurements and the

model calculation except for this one, a possible cause for the discrepancy, can be an

experimental error. This aspect of the research cannot be explored further unless a

repeated experiment is performed. This is beyond the scope of this study and hence a

repeat experiment represents a future research opportunity.

Overall, the present work benefits the industries in the direction they are heading,

which is moving the exploration of hydrocarbons toward the northern hemisphere.

Area such as the SW Barents Sea, have had isostatic deglaciation rebound in its burial

history. Therefore, sediments in the basin were affected by uplift and erosion. Many

resource evaluation methods, for example basin modelling, lithology mapping, and

rock physics modelling require the inter-relationship knowledge between

acoustic/seismic response and petrophysical properties. Understanding how complex

burial histories affect these properties can make the difference between a successful

and an unsuccessful exploration outcome.

4.1 Future Research Outlook

The experimentation and modelling studies performed on both sand compaction and

fluid property measurements have encountered numerous ideas and challenges.

Despite of being interest, some ideas and arisen questions emerged during this study

were not pursued further, and they were not addressed due to current work scope and

time allocation. However, these present potential future research opportunities.

Scientific communities as well as the petroleum industries will benefit, if the following

research areas are considered.

A reduction in the effective stress has various potential causes, with the following two

being the main contributions; 1) uplift and erosion where vertical weight of overburden

was removed and 2) an increase in pore pressure. In basins that have been affected by

uplift, the sediments may have also been experienced over-pressure, which could occur

pre- or post-uplift. The ability to distinguish between these two effects, either through

Page 44: Acoustic, elastic and physical properties of

32

laboratory experiments, rock physics modelling, or well log observations, can help us

prepare for drilling in uplifted basins, which may encounter risks due to overpressure.

In addition, the current experimental studies presented herein, were only performed

on natural sand samples. Similar experiments employed for clays or mudstones can

complement the understanding of the stress release impacting the acoustic and physical

properties of these fine-grained lithologies. If the reservoir sands were uplifted, it is

likely that the entire overburden, comprised mainly clays and mudstone that overlying

reservoir rocks, were also uplifted.

In this research, fluid saturation sensitivity for time-lapse seismic applications were

investigated. The study, however, did not explore the absolute detectable limit where

time-lapse seismic monitoring will not work for uplifted reservoir rocks. To highlight

how much maximum burial depth (or preconsolidation stress) prior to uplift will limit

time-lapse seismic response, more rock physics modelling, combining different fluid

replacement scenarios with rock properties affected by various degrees of pre-

consolidation stress is required. With these results, alternative reservoir monitoring

methods can be justified whether they are required to improve oil or gas recovery in

uplifted reservoirs.

Lastly, the fluid property measurements for brine and oil mixtures for 70-30 brine-oil

components, give distinctive results which do not fit with the mixing average models,

especially when compared to other mixtures. Repeated experiments on brine and oil

mixtures, focusing on high brine proportions, can help explain the discrepancies found

in this sample. It can also determine whether a new mixing average model is required

to estimate fluid bulk modulus for fluid substitution exercises.

Page 45: Acoustic, elastic and physical properties of

33

Bibliography

Alemu, B. L., E. Aker, M. Soldal, Ø. Johnsen, and P. Aagaard, 2012, Effect of sub‐core

scale heterogeneities on acoustic and electrical properties of a reservoir rock: a

CO2 flooding experiment of brine saturated sandstone in a computed

tomography scanner, Geophysical Prospecting, 61, no. 1, pp. 235-250

Al-Chalabi, M. and P. L. Rosenkranz, 2002, Velocity-depth and time-depth

relationships for a decompacted uplifted unit, Geophysical Prospecting, 50, no.

6, 661–664

Andersen, F. C. and T. A. Johansen, 2010, Test of rock physics models for prediction

of seismic velocities in shallow unconsolidated sands: A well log data case,

Geophysical Prospecting, 58, no. 6, pp. 1083-1098

Avseth, P., T. Mukerji, and G. Mavko, 2005, Quantitative seismic interpretation:

applying rock physics tools to reduce interpretation risk: Cambridge University

Press

Avseth P., T. Mukerji, G. Mavko, and J. Dvorkin, 2010, Rock physics diagnostics of

depositional texture, diagenetic alterations, and reservoir heterogeneity in high-

porosity siliciclastic sediments and rocks – A review of selected models and

suggested workflows, Geophysics, 75, pp. 7531–7547

Avseth P. and I. Lehocki, 2016, Combining burial history and rockphysics modeling to

constrain AVO analysis during exploration, The Leading Edge, 35, 528–534

Bahadori A., 2018, Fundamentals of Enhanced Oil and Gas Recovery from

Conventional and Unconventional Reservoirs

Batzle, M. L. and Z.-J. Wang, 1992, Seismic properties of pore fluids, Geophysics, 57,

1396–1408.

Bhuiyan, M. H., R. M. Holt, I. Larsen, and J. Stenebråten, 2013, Static and dynamic

behaviour of compacted sand and clay: Comparison between measurements in

Triaxial and Oedometric test systems, Geophysical Prospecting, 61, no. 2, pp.

329-340

Bjørkum, P.A., Oelkers, E.H., Nadeau, P.H., Walderhaug, O., Murphy, W.M., 1998.

Porosity prediction in quartzose sandstones as a function of time, temperature,

depth, stylolite frequency, and hydrocarbon stauration. AAPG Bulletin, 82, pp.

637-648

Page 46: Acoustic, elastic and physical properties of

34

Bjørkum, P.A., 1996. How important is pressure in causing dissolution of quartz in

sandstones? Journal of Sedimentary Research, 66, no. 1, pp. 147-154

Bjørlykke, K. and P. K. Egeberg, 1993, Quartz cementation in sedimentary basin,

AAPG Bulletin, 77, pp. 1536-1548

Bjørlykke, K. and K. Høeg, 1997, Effects of burial diagenesis on stresses, compaction

and fluid flow in sedimentary basins, Marine and Petroleum Geology, 14, no. 3,

pp. 267-276

Bjørlykke, K., 2010, Petroleum Geoscience. From Sedimentary Environments to Rock

Physics, Springer, 508 p

Bjørlykke, K. and J. Jahren, 2010, Sandstones and sandstone reservoirs. In Petroleum

Geoscience, Springer, Berlin, Heidelberg, pp. 113-140

Bulat, J. and S. J. Stoker, 1987, Uplift determination from interval velocity studies, UK,

southern North Sea. In: Brooks, J., Glennie, K.W. Eds., Petroleum Geology of

North West Europe, Graham and Trotman, London, pp. 293–305

Casagrande, A., 1936, The determination of the pre-consolidation load and its practical

significance", Proceedings of the international conference on soil mechanics

and foundation engineering 3, Harvard University Cambridge. pp. 60–64

Chuhan, F. A., A. Kjeldstad, K. Bjørlykke, and K. Høeg, 2003, Experimental

compression of loose sands: relevance to porosity reduction during burial in

sedimentary basins, Canadian Geotechnical Journal, 40, no. 5, pp. 995–1011

Danesh, A., 1998, PVT and phase behaviour of petroleum reservoir fluids: Elsevier.

Dashti, H. H. and M. R. Riazi, 2014, Acoustic velocities in petroleum fluids:

Measurement and prediction, Journal of Petroleum Science and Engineering,

124, pp. 94-104

Densley, M.R., Hillis, R.R., and J.E.P. Redfearn, 2000, Quantification of uplift in the

Carnarvon Basin based on interval velocities, Australian Journal of Earth

Sciences, 47, no. 1, pp. 111-122

Dræge A., K. Duffaut, T Wiik, and K. Hokstad, 2014, Linking rock physics and basin

history — Filling gaps between wells in frontier basins, The Leading Edge, 33,

240–246

Dutta, T., G. Mavko, and T. Mukerji, 2010, Improved granular medium model for

unconsolidated sands using coordination number, porosity, and pressure

relations, Geophysics, 76, no. 2, E91–E99

Dvorkin, J., and A. Nur, 1996, Elasticity of high-porosity sandstones: Theory for two

North Sea data sets, Geophysics, 61, pp. 1363–1370

Page 47: Acoustic, elastic and physical properties of

35

Fawad, M., N. H. Mondol, J. Jahren, and K. Bjørlykke, 2011, Mechanical compaction

and ultrasonic velocity of sands with different texture and mineralogical

composition: Geophysical Prospecting, 59, no. 4, p. 697- 720

Gassmann, F., 1951, Uber die Elastizitat Poroser Medien: Vier. Der Natur.

Gesellschaft in Zurich ,96, pp. 1–23

Grande, L., N. H. Mondol, and T. K. Berre, 2011, Horizontal stress development in

fine-grained sediments and mudstones during compaction and uplift, EAGE

Extended Abstract, EarthDoc, P372, 73rd EAGE Conference & Exhibition

incorporating SPE EUROPEC, 23-26 May 2011, Vienna, Austria

Green, P. F., 1989, Thermal and tectonic history of the East Midlands shelf onshore

UK. and surrounding regions assessed by apatite fission track analysis, Journal

Geological Society of London, 146, pp. 755–774

Grude S., M. Landrø and B. Osdal, 2013, Time-lapse pressure saturation discrimination

for CO2 storage at the Snøhvit field, International Journal of Greenhouse Gas

Controls, 19, 369–378

Han, D. and M. L. Batzle, 2000, Velocity, density and modulus of hydrocarbon fluids—

Data measurement: 70th Annual International Meeting, Society of Exploration

Geophysicists, Expanded Abstracts, paper RPB 7.2

Han, D. H. and M. Sun, 2013, Velocity and density of water with dissolved CH4 and

CO2, SEG Expanded abstract, pp. 2846-2850

Hansen, S., 1996a. Quantification of net uplift and erosion on the Norwegian Shelf

south of 66 degrees N from sonic transit times of shale. Norsk Geologisk

Tidsskrift, 76, no. 4, pp. 245-252

Hashin Z. and S. Shtrikman, 1962, A variational approach to the theory of the elastic

behaviour of polycrystals, Journal of the Mechanics and Physics of Solids, 10,

343–352

Hashin Z. and S. Shtrikman, 1963, A variational approach to the elastic behaviour of

multiphase materials, Journal of the Mechanics and Physics of Solids, 11, pp.

127–140

Hatchell, P. and S. Bourne, 2005, Rock under strain: Strain-induced time-lapse time

shifts are observed for depleting reservoirs, The Leading Edge, 24, pp. 1222-

1225

Henriksen E., H. Bjørnseth, T. Hals, T. Heide, T. Kiryukhina, O. Kløvjan, G. Larssen,

A. Ryseth, K. Rønning, and K. Sollid, 2011, Uplift and erosion of the greater

Barents Sea: impact on prospectivity and petroleum systems, Geological Society

of London Memoirs 35, pp. 271–281

Page 48: Acoustic, elastic and physical properties of

36

Hill R., 1952, The elastic behaviour of a crystalline aggregate, Proceedings of the

Physical Society, A65, 351

Jaky, J., 1948, Pressure in silos, Proceedings of the 2nd International Conference on

SMGE, 1, 103–109

Japsen, P., 1998, Regional velocity-depth anomalies, North Sea Chalk: a record of

overpressure and Neogene uplift and erosion, American Association of

Petroleum Geology Bulletin, 82, pp. 2031–2074

Jensen, L.N. and B. J. Schmidt, 1992, Late Tertiary uplift and erosion in the Skagerrak

area; magnitude and consequences, Norsk Geologisk Tidsskrift, 72, pp. 275–

279

McCain, W. D., 1990, The Properties of petroleum fluids: Penn Well Books.

Mavko, G., T. Mukerji, and J. Dvorkin, 2009, The rock physics handbook, Cambridge

University Press.

Mayne, P. W. and F.H. Kulhawy, 1983, K-OCR relationships in soil, Journal of the

Geotechnical Engineering Division, 20, no. 1

Mindlin R.D., 1949, Compliance of elastic bodies in contact. Journal of Applied

Mechanics 16, 259–268

Mondol, N. H., P. Avseth, M. Fawad, and T. Smith, 2010, Vs prediction in

unconsolidated sands, Physical and Geological controls on shear wave velocity,

72nd EAGE Meeting, Expanded Abstract, p351

Narongsirikul, S., N. H. Mondol, and J. Jahren, 2013, Possible application of friable

sand model for shallow mechanically compacted overconsolidated sands,

Conference Proceedings, SEG Technical Program Expanded Abstracts 2013,

Houston, USA, pp. 2701-2706

Oelkers, E.H., P.A. Bjørkum, and W.M. Murphy, 1992, The mechanism of porosity

reduction, stylolite development and quartz cementation in North Sea

sandstones. In: Kharaka, Y.K., Maest, A.S. (Eds.), Water-rock interaction.

Rotterdam, Balkema, 2, pp. 1183-1186

Oelkers, E.H., P.A. Bjørkum, and W.M. Murphy, 1993, Calculation of the rate and

distribution of chemically driven quartz cementation in North Sea sandstones.

In: Cuney, M., Cathelineau, M. (Eds.), Proceedings of the Fourth International

Symposium on Hydrothermal Reactions. Institute Lorain des Geosciences,

Nancy, pp. 169-172

Page 49: Acoustic, elastic and physical properties of

37

Oelkers, E.H., P.A. Bjørkum, and W.M. Murphy, 1996, A Petrographic and

computational investigation of quartz cementation and porosity reduction in

North Sea sandstones, American Journal of Science, 296, pp. 420-452

Pender, M. J., 1978, A model for the behavior of overconsolidated soil, Geotechnique,

28, 1, pp. 1-25

Pettersen, Ø., 2007, Sandstone compaction, grain packing and critical state theory,

Petroleum Geoscience, 13, pp. 63-67

Prasad, M., 2002, Acoustic measurements in sands at low effective pressure:

Overpressure detection in sands, Geophysics, 67, 2, pp. 405 – 412

Reuss, A., 1929, Berechnung der Fliessgrenzen von Mischkristallen. Z. Angew, Math.

Mech, 9, pp. 49-58

Riis, F., and L.N. Jensen, 1992, Introduction; measuring uplift and erosion; proposal

for a terminology Norsk Geologisk Tidsskrift, 72, pp. 223–228

Smith, T. M., C. H. Sondergeld, and C. S. Rai, 2003, Gassmann fluid substitutions: A

tutorial. Geophysics, 68, pp. 430–440

Storvoll, V., Bjørlykke, K., Mondol, N.H., 2005, Velocity-depth trends in Mesozoic

and Cenozoic sediments from the Norwegian shelf, AAPG Bulletin, 89, no. 3,

pp. 359-381

Terzaghi, K., 1943, Theoretical soil mechanics, John Wiley & Sons, Inc.

Tao, G., M. S. King, and M. Nabi-Bidhendi, 1995, Ultrasolic wave propagation in dry

and brine-saturated sandstones as a function of effective stress: laboratory

measurements and modelling, Geophysical Prospecting, 42, pp. 299-328

Voigt, W., 1910, Lehrbuch der Kristallphysik. Teubner, Leipzig

Walderhaug, O., 1996, Kinetic modeling of quartz cementation and porosity loss in

deeply buried sandstone reservoirs. AAPG Bulletin, 80, no. 5, pp. 731-745

Wood, A. B., 1955, Textbook of Sound, G. Bell and Sons td., London, p 360

Yin, H., 1992, Acoustic velocity and attenuation of rocks: Isotropy, intrinsic anisotropy

and stress induced anisotropy. Ph. D. Thesis, Stanford University, CA, United

States

Zoback, M.D., D. Moos, L. Mastin and R. N. Anderson, 1985, Well bore breakouts

and in situ stress, Journal of Geophysical Research, 90, no. B7, pp. 5523–5530

Page 50: Acoustic, elastic and physical properties of

38

Page 51: Acoustic, elastic and physical properties of

39

Part II

Enclosed Papers

Page 52: Acoustic, elastic and physical properties of

40

Page 53: Acoustic, elastic and physical properties of

Paper 3

Effects of stress reduction on geomechanical and acoustic relationship of overconsolidated sands

By

Sirikarn Narongsirikul, Nazmul Haque Mondol, and Jens Jahren

Geophysical Prospecting (2020)

Vol 68, No 3, March 2020 pp. 968–981

DOT: 10.1111/1365-2478.12902

3

Publication

Page 54: Acoustic, elastic and physical properties of
Page 55: Acoustic, elastic and physical properties of

Geophysical Prospecting, 2020, 68, 968–981 doi: 10.1111/1365-2478.12902

Effects of stress reduction on geomechanical and acoustic relationshipof overconsolidated sands

Sirikarn Narongsirikul1,3∗, Nazmul Haque Mondol1,2 and Jens Jahren1

1Department of Geosciences, University of Oslo, Problemveien 7, 0315, Oslo, Norway, 2Norwegian Geotechnical Institute (NGI), Oslo,Norway, and 3ConocoPhillips Norway, Ekofiskvegen 35, 4056, Tananger, Norway

Received April 2019, revision accepted October 2019

ABSTRACTRelationship between different geomechanical and acoustic properties measured fromseven laboratory-tested unconsolidated natural sands with different mineralogicalcompositions and textures were presented. The samples were compacted in the uni-axial strain configuration from 0.5 to 30 MPa effective stress. Each sand sample wassubjected to three loading–unloading cycles to study the influence of stress reduc-tion. Geomechanical, elastic and acoustic parameters are different between normalcompaction and overconsolidation (unloaded and reloaded). Stress path (K0) datadiffer between normal consolidated and overconsolidated sediments. The K0 value ofapproximately 0.5 is founded for most of the normal consolidated sands, but variesduring unloading depending on mineral compositions and textural differences. TheK0 and overconsolidation ratio relation can be further simplified and can be influ-enced by the material compositions. K0 can be used to estimate horizontal stress fordrilling applications. The relationship between acoustic velocity and geomechanicalis also found to be different between loading and unloading conditions. The staticmoduli of the overconsolidated sands are much higher than normal consolidatedsands as the deformation is small (small strain). The correlation between dynamicand static elastic moduli is stronger for an overconsolidation stage than for a normalconsolidation stage. The results of this study can contribute to geomechanical andacoustic dataset which can be applied for many seismic-geomechanics applications inshallow sands where mechanical compaction is the dominant mechanism.

Key words: Overconsolidation, Stress path, Static, Dynamic elastic.

1 INTRODUCTI ON

Effective stress reduction due to sediment uplift or excess porefluid pressure affects acoustic, geomechanical and petrophysi-cal properties (Holt 1990; Bowers and Katsube 2002; Zimmeret al. 2007; Narongsirikul, Mondol and Jahren 2019a,b). In-crease of vertical weight on the sediments results in poros-ity loss, as well as increase of velocity and density in theunderlain sediments. In a contrary, unloading the stress act-ing on rocks results in velocity and porosity reversal. How-ever, the reversal of velocity and density is not fully re-

∗E-mail: [email protected]

gained back the values as the previous stress level due topermanent deformation (Bower 1995; Bowers and Katsube2002).

The different effects the stress condition has on the sed-iment properties, due to different burial histories, need to beconsidered in reservoir characterization and seismic modellingstudies. Moreover, in the seismic-related geomechanics field,understanding the correlation between geomechanical param-eters and acoustic relations is critical, as one data type (e.g.seismic velocities) can be used to extract another (e.g. mechan-ical stress and strain). A link between acoustic/seismic andgeomechanical parameters can be expressed through various

968 C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers

This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproductionin any medium, provided the original work is properly cited.

Page 56: Acoustic, elastic and physical properties of

Effects of stress reduction of overconsolidated sands 969

relationships. The relation between dynamic and static elasticmoduli is one of the most common correlations in geome-chanics. This relation is especially required when the dynamicmoduli calculated from acoustic logs are needed to convert tostatic elastic moduli, as the static moduli are rarely obtaineddue to limited core measurements. The static elastic proper-ties such as Young’s and Shear Modulus, and Poison’s ratio,are important for geomechanical modelling. Uniaxial or con-strained modulus is important for reservoir compaction study.Another common relationship is between velocity and strain.Previous studies have demonstrated that such a correlationis effective in evaluating seismic-geomechanics coupling viaR-factor, a ratio of the change between velocity and in thevertical strain (e.g. Hatchell and Bourne 2005; Holt, Nes andFjær 2005; Røste, Stovas and Landro 2005; Rickett et al.

2007; Herwanger and Horne 2009). This correlation meansseismic responses of four-dimensional time-shift or time-straincan give information about geomechanical changes due toproduction, that is deformation and dilation. Another geome-chanical parameters which are important for geomechanicsapplications are horizonal stresses. These parameters are oftenaccurately obtained through leak off test or borehole breakoutdata. However, without drilling, the horizontal stresses needto be estimated by other means. Previous studies show that thestress path (K0) can give an estimation of effective horizontalstress if effective vertical stress is known (Holt 1999; Grande,Mondol and Berre 2011; Holt et al. 2016). K0 can also be es-timated by overconsolidation ratio (OCR, Jaky 1948; Mayneand Kulhawy 1983).

The seismic and rock mechanical relations of nor-mally compacted sediments have been investigated extensivelythrough laboratory experiments on compacted sands studies(Yin 1992; Tao, King and Nabi-Bidhendi 1995; Dvorkin andNur 1996; Chuhan et al. 2003; Pettersen 2007; Mondol et al.

2010; Fawad et al. 2011; Bhuiyan et al. 2013). Such exper-iments lead to a good collection of data that can be utilizedfor geomechanics and seismic integration for normal consol-idation (NC) stress conditions. However, a few experimen-tal approaches have been developed with the aim of under-standing the effects of stress unloading on rock properties foroverconsolidated (OC) sediments (Holt 1994; Nygard et al.

2004; Bhuiyan, Kolstø and Holt 2011; Grande et al. 2011;Dræge et al. 2014; Avseth and Lehocki 2016). More studiescan mature the understanding of the subject and increase thecollection of database for seismic and geomechanical appli-cations such as reservoir compaction and time-lapse seismicstudies.

This study reports an experimental investigation of sevenunconsolidated sands with varying mineralogical composi-tions and textures, aiming at mechanical parameters ob-tained from the same laboratory measurements performedby Narongsirikul et al. (2019a). The study applies cyclicalstress paths under zero horizontal strain conditions (K0) in-cluding three stages of loading, partial unloading and reload-ing to investigate different behaviours between NC and OC.Mechanical properties (stress, strain, OCR and K0) mea-sured from the experiments are reported. These data are usedto estimate static elastic moduli. The acoustic velocity datastudied in Narongsirikul et al. (2019a,b) are used to esti-mate dynamic elastic properties. The relationships betweenrock acoustic and mechanical properties, and the differentbehaviours between normal compaction and OC are inves-tigated. The study aims at correlating different geomechan-ical parameters together with acoustic and computed elasticproperties, which are common for geomechanics and seismicapplications.

2 NORMAL CONSOLIDATION VERSUSOVERCONSOLIDATION

Normal consolidation (NC) refers to a continuous stressincrease during sediments compaction. NC, normal com-paction, normal loading and virgin compaction are the termsused interchangeably. In contrast, sediments that are atpresent found at lower stress than previously experienced dueto a reduction of the overburden (erosional unloading) or dueto excess pore pressure are termed overconsolidated (OC; e.g.Pender 1978; Bjørlykke 2010). The term ‘pre-consolidationstress’ was used for the past maximum effective stress priorto stress reduction if the sediments have been preloaded. Theoverconsolidation ratio (OCR) is defined by the difference be-tween past maximum effective vertical stress σ ′

v max and thepresent effective vertical stress σ ′

v, (Casagrande 1936) that is

OCR = σ ′v max

σ ′v

. (1)

When the OCR is equal to 1, sediments are normallyconsolidated, that is when the past maximum effective stressand the present effective stress are equal.

K0 can be used for an estimation of the effective hori-zontal stress as well as an indication of stress anisotropy. K0

usually defined a stress path for the sediment loaded undera uniaxial strain condition in soil mechanics. K0 is the ratio

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 57: Acoustic, elastic and physical properties of

970 S. Narongsirikul, N.H. Mondol and J. Jahren

between horizontal effective stress σ ′h and vertical effective

stress σ ′v, that is

K0 = σ ′h

σ ′v

. (2)

For example, for K0 = 1, the stress is hydrostatic implyingthat the vertical and horizontal effective stresses are equal.For K0 < 1, the stress is uniaxial and the vertical effectivestress is greater than the horizontal effective stress. For loosesand sediments loaded under normal consolidation, K0 equalsapproximately 0.5 (Grande et al. 2011).

Finally, the term effective stress, σ′, is defined as σ

′ = σ −αp, where σ is the total stress, and subscript v and h denotethe direction of vertical and horizontal stresses, respectively, p

is the pore pressure and α is an effective stress coefficient. Fol-lowing Terzaghi (1943), the effective stress coefficient valueis assumed to be 1 for unconsolidated sediments.

3 S TATIC A ND DY N A MI C ELA ST I CM O D U L I

Elastic moduli are elastic parameters which can be derivedfrom laboratory measurement and well logs. The elastic mod-uli calculated from deformational experiments are the staticmoduli. Whilst, the elastic moduli calculated from elastic wavevelocities (P- and S- waves, Vp and Vs) and density (ρ) are thedynamic moduli (Mavko, Mukerji and Dvorkin 2009). Fiveelastic parameters considered in the present study are listedbelow. The definition and the equation are based on Mavkoet al. (2009).

Young’s Modulus, E, for static is defined as the ratio ofthe extensional stress to the extensional stain in a uniaxialstress state:

Estatic = σzz

εzz. (3)

For dynamic:

Edynamic = ρV2s

(3V2

p − 4 V2s

)

(V2

p − V2s

) . (4)

Poisson’s ratio, ν, for static is defined as minus the ratioof the lateral strain to the axial strain in a uniaxial stress state:

νstatic = σzz

εzz. (5)

For dynamic:

νdynamic =(V2

p − 2 V2s

)

2(V2

p − V2s

) . (6)

Constrained or P-wave modulus, M, for static is definedas the ratio of the axial stress to the axial strain in a uniaxialstrain state:

Mstatic = σzz

εzz. (7)

For dynamic:

Mdynamic = ρV2p . (8)

Bulk Modulus, K, for static is defined as the ratio of thehydrostatic or isotropic stress, σ 0, to the volumetric strain(εαα):

Kstatic = σ0

εαα

. (9)

For dynamic:

Kdynamic = ρ

(V2

p − 43

V2s

). (10)

Shear Modulus, μ, for static is defined as the ratio ofthe shear stress to the shear strain. It can be defined by thefunctions between Young’s Modulus (E) and Poisson’s Ratio(ν) as

μstatic = E2 (1 + ν)

. (11)

For dynamic:

μdynamic = ρV2s . (12)

4 D ATASET AND EXPERIMENTALP R O C E D U R E

The same data as reported in Narongsirikul et al. (2019a)are used herein. A total of seven experimental compactiontests were performed on seven brine-saturated natural sandaggregates with varying mineral compositions and textures(Tables 1 and 2).

An axi-symmetric triaxial cell located at the NorwegianGeotechnical Institute was used for measurements of rockproperties in this study. The samples were initially isotrop-ically compacted to an effective stress of 0.52 MPa in thetriaxial cell. The effective stress was then increased from 0.52to 30 MPa under a uniaxial strain condition (K0), which wasmaintained through adjustment of radial stress (σ

′h, onset,

Fig. 1). Pore pressure was kept constant at 1 MPa throughoutthe experiments. To assure complete saturation prior to com-paction, vacuum was applied before fluid saturation to ensureno air was left inside the samples. Carbon dioxide (CO2) wasthen applied and left within the samples for a few minutes

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 58: Acoustic, elastic and physical properties of

Effects of stress reduction of overconsolidated sands 971

Table 1 Mineral compositions from the XRD analysis (after Fawad et al. 2011)

Mineral constituents (weight percentage, %)

Sample name Quartz Feldspar1 Clay2 Other minerals3 Effective grain density (g/cc) Initial porosity (%)

Quartz arenite 1 (QA1) 95.27 4.73 – – 2.65 36Quartz arenite 2 (QA2) 97.88 1.77 – 0.35 2.65 41Sub arkose 1 (SA1) 91.27 8.73 – – 2.65 38Sub arkose 2 (SA2) 77.19 22.81 – – 2.64 34Arkosic arenite (AA) 54.84 34.18 10.97 – 2.65 42Volcanic arenite (VA) 4.84 52.51 – 42.65 2.81 40

1Includes K-feldspar, albite, and plagioclase.2Includes kaolinite and illite.3Mostly present in volcanic arenite including aragonite, calcite, ankerite, amphibole and augite.

before brine solution (35 g NaCl per litre of water) was in-jected into the samples. Finally, the back pressure applicationwas introduced to ensure fully liquid saturation. Details ofthe back pressure application can be found in Berre (2011).At 15, 25 and 30 MPa, effective stresses partial unloadingwas carried out by decreasing the vertical load followed byreloading to the next stress level (Fig. 1).

A maximum strain amplitude of the order of 10−2 canbe achieved in this experiment setup. The strains were con-tinuously recorded during each loading and unloading cyclethrough the changes in the sample height and diameter usingvertical and horizontal deformation sensors (Fig. 1). Wavevelocities both P- and S- and density data were taken fromNarongsirikul et al. (2019a) to estimate the dynamic elasticproperties of the sands. The detailed setup of the triaxial cellis described in detail in Berre (2011) and Narongsirikul et al.

(2019a).

5 EX P E R I M E N T A L R E S U L T S

5.1 Stress, strain, K0 and overconsolidation ratio

Figure 2 shows relations between (a) stress–strain (b) stress–K0 (c) overconsolidated ratio (OCR)–strain and (d) OCR–K0

for all sand samples at all pressure steps. Solid lines in allplots denote normal consolidation and dashed lines signifyoverconsolidation from unloading/reloading. OCR and K0 arecalculated using equations (1) and (2), respectively. The gen-eral observation for all the plots shows that all samples in anoverconsolidated (OC) condition exhibit distinctive behaviourcompared to the samples during normal compaction. The dif-ference in magnitude of the parameters between the samplesis attributed to the variations in the mineral compositions andtextures of each sample.

Figure 2(a) shows the stress–strain relationship plottedfor all samples. The stress versus strain relation is an important

coupled attribute for a reservoir during compaction, which iscommonly known to result from production/injection-relatedstress changes. The slopes calculated from stress–strain rela-tion yield the uniaxial compaction coefficient which is oneof the key parameters for the estimation of the magnitude ofreservoir deformation (Geertsma 1973). From the plot, strainincreases with an increase in effective stress in the range 50–180 millistrain at the maximum effective stress at 30 MPa. Ingeneral, the samples with quartz rich content of >90% quartzcontent [quartz arenite 1 (QA1), quartz arenite 2 (QA2) andsub arkose 1 (SA1)] establish flatter slopes than the sampleswith medium to low quartz content [sub arkose 2 (SA2), arko-sis arenite (AA) and feldspathic greywacke (FG), except vol-canic arenite (VA)] during normal loading. The VA sample hasthe lowest quartz content. However, the sample exhibits in-termediate stress–strain gradient between high and low quartzsamples, instead of steep slope like what observed for otherlow quartz members. Other minerals (e.g. aragonite and cal-cite) of VA composite may contribute to this behaviour (SeeTable 2). At any given stress level, decreasing quartz con-tent (increase of ductile and clay minerals) results in higherstrain or compressibility. This shows that rock compositionaffects the compaction behaviour. However, this behaviourcontrasts with that of the samples under unloading/reloadingconditions, where different constituents do not significantlygovern the degree of deformation, since all samples exhibitsimilar unloading strain–stress gradients (Fig. 2a). This may bedue to the effect of permanent deformation the sands experi-enced from pre-stressing. The strain observed in all samples forOC curves shows elastic hysteresis. For all unloading cycles,the slopes of stress and strain curve (dashed lines) are of thesame degree regardless of the change in the pre-consolidationstresses.

Figure 2(b) shows the stress–K0 relation for all load-ing, unloading and reloading cycles. As reviewed in the

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 59: Acoustic, elastic and physical properties of

972 S. Narongsirikul, N.H. Mondol and J. Jahren

Tab

le2

Tex

tura

lvar

iati

ons

ofth

esa

ndsa

mpl

es(a

fter

Faw

adet

al.2

011)

,san

dgr

ain

size

dist

ribu

tion

(wei

ght

perc

enta

ge,%

),so

rtin

gcl

assi

fica

tion

and

grai

nsh

apes

(fra

ctio

n)

Ver

yco

arse

Coa

rse

Med

.Fi

neV

ery

fine

Silt

Mea

ngr

ain

size

(µm

)So

rtin

gde

gree

(phi

)So

rtin

gC

lass

ific

atio

nR

ound

ness

(fra

ctio

n)Sp

heri

city

(fra

ctio

n)Sa

mpl

e>

1000µ

m50

0–10

00µ

m25

0–50

m12

5–25

m63

–125µ

m<

63µ

m

QA

10.

524.

7739

.86

40.2

34.

909.

7221

3.2

1.00

Poor

lyso

rted

0.57

0.48

QA

20.

021.

0978

.69

18.4

80.

840.

8830

3.5

0.51

Mod

erat

ely

wel

lsor

ted

0.54

0.42

SA1

4.66

32.3

261

.80

1.19

–0.

0346

6.5

0.60

Mod

erat

ely

wel

lsor

ted

0.58

0.48

SA2

0.37

22.3

840

.57

20.0

70.

3016

.31

257.

01.

31Po

orly

sort

ed0.

450.

35A

A0.

024.

0958

.49

28.7

61.

047.

6025

0.0

0.90

Mod

erat

ely

sort

ed0.

470.

38

VA

–0.

0618

.13

74.2

10.

217.

3918

0.5

0.69

Mod

erat

ely

wel

lsor

ted

0.53

0.43

introductory section, K0 can be used to estimate the effectivehorizontal stress as well as an indicator of stress anisotropy.Here, we find that as the samples underwent normal com-paction under a uniaxial strain condition, K0 ranges between0.45 and 0.55 (solid lines) and remains fairly constant acrossthe stress range between 5 and 30 MPa. This observation iscomparable with a typical K0 value of 0.5 for unconsolidatedsands under normal consolidation in soil mechanics. For nor-mal consolidation, the difference in mineral compositions isresponsible for variations in the K0 development where thesamples with high quartz content (>90% weight percent), thatis QA2, QA1 and SA1 (black, magenta and blue solid lines)show high K0 compared to samples with lower quartz con-tents (i.e. SA2, AA, VA and FG). Increasing amount of quartz(decreasing amount of clays and ductile minerals) increases K0.

For OC when the stress was removed during the unloadingwith resultant decreasing effective stress, the development ofK0 increases approaching 1 (dashed lines). This means that thevertical and horizontal stress becomes less anisotropic fromstress reduction.

Figure 2(c) illustrates the OCR–strain relation for all sam-ples. The definition of the OCR is referred to in equation (1).During normal consolidation (NC), the OCR is equal to 1(solid lines) as the maximum effective stress is the same aspresent effective stress. Therefore, the NC condition plots atOCR = 1 for all samples. When the samples were de-stressedand re-stressed, depicted by dashed lines, increasing the OCRdoes not change the strain significantly due to hysteresis effect(permanent damage of pore structure during preconsolida-tion). This behaviour is observed for all unloading cycles.

Figure 2(d) demonstrates the relation between OCR andK0. A clear separation can be observed between the normalcompaction trend (solid lines plotted at OCR = 1) and OClines (dashed lines). The unloading/reloading trends show lin-early proportional relationship between OCR and K0 whereK0 increases with an increase in the OCR. At the points wherethe OCR is equal to 1 and K0 is equal approximately to 0.5,the samples are at the maximum effective stress prior to thestress reduction. As the stress was further reduced from themaximum stress point, the OCR and K0 begin to increase ina linear fashion. As previously stated, that the K0 can be anindicator of stress anisotropy, this means the reduction in theeffective stress causes the stress anisotropy to become weakeras K0 approaches 1 (isotropic).

In the same plot (Fig. 2d), we also find that the samplesthat comprised greater amounts of quartz exhibit higher K0

at any given OCR. For example, at OCR 2.5, a group ofsamples that comprised high quartz (QA2, QA1, SA1) has

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 60: Acoustic, elastic and physical properties of

Effects of stress reduction of overconsolidated sands 973

0 5 10 15 20 250

5

10

15

20

25

30

35

Loading path (step number)

(MP

a)ssert

Sevitceff

E

Cycle 1

Cycle 3

Cycle 2

K0-ConsolidationIsotropic

σ'v K0 condition

σ'v > σ’h εh = 0σ'h

Normal load = Normal consolidationUnload/Reload = Overconsolidation

Figure 1 Loading protocol for experimental mechanical compaction applied to the sand samples in the present study (from Narongsirikul et al.2019a). The procedure employs loading, unloading and reloading cycles. Unloading was applied at 15, 25 and 30 MPa effective stresses beforefurther reloading. The uniaxial strain loading (K0-consolidation) condition is demonstrated in the upper left corner of the figure where zerohorizontal strain, εh = 0, is controlled by adjustment of effective horizontal stress, σ ′

h, while the effective vertical stress is increased, σ ′v .

K0 of approximately 0.85–0.9 while the group of sampleswith lower quartz (SA2, AA, FG, VA) has K0 of 0.65–0.75.The SA2 sample has lower K0 than what is expected from asample with a moderate amount of quartz content. This maybe explained by textural parameter differences, possibly thesorting in this sample may have a stronger control than themineralogy alone.

5.2 Relationship between acoustic and mechanicalproperties

Measurements of the acoustic and mechanical properties ofthe samples expressed as relationships between strain ver-sus velocities, K0 versus velocities and overconsolidation ratio(OCR) versus velocities plots are shown in Fig. 3. The P-wave velocities are on the left column (Fig. 3a,c,e) and theS-wave velocities are on the right column (Fig. 3b,d,f). In gen-eral, all three mechanical properties exhibit similar relation-ships with P-wave and S-wave velocities. The difference is themagnitude where the shear wave propagation phenomenon isoverall slower than the compressional wave. The normal con-solidation (NC) trends (solid lines) are noticeably different

from the overconsolidation (OC) trends (OC, dashed lines).In addition, mineral compositions still play a significant rolein exhibiting the relation varying between samples to samplesin all plots.

In the strain–velocities plot (Fig. 3a,b), the inter-relationbetween these attributes are non-linear at low strain and laterdevelop a linear trend at higher strain as the deformationcontinues. These behaviours are observed for all samples dur-ing normal compaction (solid lines). As with the strain–stressplot (Fig. 2), a significant deviation from normal compactionoccurs for OC (dashed lines). The strain–velocity coefficientfor OC is low compared to normal compaction, indicated bythe low gradient on all unloading cycle lines. This means asmall change in strain can be expected from a large changeof velocities in overconsolidated sediments. In both OC andNC conditions, we find that the magnitude of strain changescorrelate linearly with the changes in velocity. The link be-tween these attributes can be useful approximation of theR-factor in seismic geomechanics application where R-factoris defined the ratio between travel time changes due to veloc-ity changes and the path length changes due to physical strainchanges (Hatchell and Bourne 2005; Røste et al. 2005). This

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 61: Acoustic, elastic and physical properties of

974 S. Narongsirikul, N.H. Mondol and J. Jahren

Figure 2 Geomechanical parameters for all samples and all loading/unloading/reloading cycles expressed in term of (a) stress and strain, (b)stress and K0, (c) Overconsolidation ratio (OCR) and strain and (d) OCR and K0. The normal consolidation trend (NC, solid lines, OCR =1) separates clearly from overconsolidation trend (OC, dashed lines). The data span in all plots is attributed to the mineralogical and texturaldifferences comprised in each sample.

R-factor allows the transformation from seismic travel-timeto displacement.

In the K0 and OCR versus velocities plot (Fig. 3c–f), anincrease of either K0 or OCR means the effective stress isdecreased. As a consequence, the velocities for both P- andS-waves decrease from the maximum pre-consolidation stress(dashed lines). The K0 and OCR versus velocities show thesame behaviour both during NC where the values are constantand during OC where the value shows an inverted polynomialtrend.

5.3 Dynamic and static moduli

Dynamic elastic moduli obtained by wave propagation mea-surements or log-derived data are a well-known applicationfor seismic modelling studies. For example, fluid substitutionrequired effective bulk and shear stiffness information. Staticelastic moduli are of the same importance as this informationis often used in many in situ stress applications, for examplewellbore stability, compaction and hydraulic fracturing. In lin-ear elasticity of isotropic homogenous materials, static elasticmoduli or stiffness is generalized by Hooke’s law, σ = –C/ε,

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 62: Acoustic, elastic and physical properties of

Effects of stress reduction of overconsolidated sands 975

Figure 3 Geomechanical versus acoustic relations for all samples demonstrating links between velocities, both P- and S-waves, and, from the topdown, strain (a, b), K0 (c, d) and OCR (e, f), respectively. The normal consolidation (NC) stage is plotted as solid lines and overconsolidation(OC) stage is plotted as dashed lines. The NC and OC stages clearly have different trends.

where the modulus, C, is a ratio between stress, σ , and strain,ε. The static moduli are not often acquired due to expensivecoring. Therefore, the conversion from commonly obtaineddynamic moduli to static moduli is important. Figures 4–6show the static and dynamic elastic parameters derived fromexperimental measurements in the present study. The equationand the description of the elastic moduli for each parameter,both static and dynamic, were explained in the Section 3. Theparameters presented herein comprised constrained modulus,bulk modulus, shear modulus, Young’s modulus and Poisson’sRatio.

In Fig. 4, measurement of static (open symbol) and dy-namic (filled symbol) elastic parameters between the two endmembers, the quartz arenite 2 (QA2, red) and the felds-pathic greywacke (FG, blue), were plotted. VA has the low-est quartz content. However, VA’s other composite mineralsare uniquely different from other samples. Therefore, FG wasused to represent the end member instead. Both the normalconsolidated stages and unloading/reloading stages for all cy-cles are included in the plots and depicted as a circle symboland a square symbol, respectively. All the static and dynamicproperties were plotted as functions of effective stress. For

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 63: Acoustic, elastic and physical properties of

976 S. Narongsirikul, N.H. Mondol and J. Jahren

Figure 4 Elastic moduli parameters, both static (open symbols) and dynamic (filled symbols), versus effective stress for two representativesamples, QA2 (red) and FG (blue). The plots are for all pressure steps both normal loading (circles) and unloading/reloading (square).

all the elastic parameters, dynamic moduli (fill symbol) over-all are higher than static moduli (filled symbol) as expected.The quartz arenite 2 sample that represents the high quartzend member has lower elastic moduli than the low quartzfeldspathic greywacke. However, the magnitude differencebetween the two end members is less pronounced for staticelastic moduli.

All plots in Figs 5 and 6 show a relationship betweenstatic and dynamic elastic moduli for all samples at all pres-sure steps. Figure 5 colour-codes the data with effective stress

which is grouped into low, medium and high stress levels.Figure 6 colour-codes the data with quartz content separatinginto three groups of 0–40%, 40–80% and >80% quartz per-centage. The red solid line in all plots is one-to-one correspon-dence (1:1 relationship) between the moduli. The black solidlines are the differences where the correspondences are one-to-n, where n = 0.25, 0.5, 2, 4 and 8. The higher number of n

shows the greater difference between the static and dynamicmoduli. All derived data plot on the left side of the one-to-onecorrespondence indicate that the dynamic moduli are higher

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 64: Acoustic, elastic and physical properties of

Effects of stress reduction of overconsolidated sands 977

Figure 5 Relations between static and dynamic properties for all samples coloured by applied stress for all samples. Circles are normalconsolidated (NC) sands, and squares are overconsolidated (OC) sands. The correspondences between 1:0.25 and 1:8 were projected on allstatic and dynamic elastic moduli relations. In all plots, for NC (circles) dynamic moduli are higher than static moduli and increase with stressas the data plotted beyond 1:8 correlation. The overconsolidated samples (squares) have static and dynamic elastic relationship closer to 1:1correspondence.

than static moduli, except for Poisson’s ratio where the dataare plotted close to 1:1 relationship. Note that Poisson’s ratiois the mechanical parameter that represents strain–strain re-lationship, instead of stress–strain relationship like the otherfour moduli. Therefore, the dynamic-static elastic relation canbe different from the other elastic moduli.

The deformation during normal compaction is far greaterthan the deformation during unloading/reloading at any givenstress. Therefore, the derived static elastic moduli for normalcompaction data (circles) were expected to be low (due to

higher strain). This agrees with the observations for all elasticmoduli in Figs 5 and 6 (circles) that the normal compactiondata are plotted as very low values of static moduli, indicatingthat the sediments loaded under normal compaction are morecompliant. The relation between static and dynamic elasticproperties is very different, indicating by the observed data,is beyond the one-to-eight correspondence. In contrast, thestatic moduli of the overconsolidated (OC) sands are muchhigher as the deformation is small (small strain). Therefore,the static moduli are high and the rock is stiff due to the effect

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 65: Acoustic, elastic and physical properties of

978 S. Narongsirikul, N.H. Mondol and J. Jahren

Figure 6 Relations between static and dynamic properties for all samples coloured by quartz content for all samples. Circles are normalconsolidated (NC) sands, and squares are overconsolidated sands. The correspondences between 1:0.25 and 1:8 were projected on all staticand dynamic elastic moduli relations. In all plots, for normal consolidation (circles) dynamic moduli are higher than static moduli as thedata plotted beyond 1:8 correlation. Overconsolidated (OC) samples with high quartz show dynamic–static relation closer to 1:1 relationshipcompared to groups of samples with lower or medium quartz content. The OC (squares) have static and dynamic elastic relationship closer to1:1 correspondence.

of pre-consolidation. Thus, the static moduli of the unload-ing/reloading sands correlate more closely with the dynamicelastic moduli. The static and dynamic relation of the con-strained and bulk modulus exhibit a correlation between 1:3and 1:8 correspondences. The shear modulus and the Young’smodulus are closer to a one-to-one which observed to be be-tween 1:1.5 and 1:8. Comparing between the influence ofstress and sand compositions, the variations in all moduli are

affected by the mineral compositions as can be seen fromFig. 6. For example, the OC samples grouped as high quartzcontent is closer to a one-to-one relationship compared to theother two groups. This observation suggested that static mod-ulus estimated from static–dynamic relations for OC sandscan vary from 30% to 87% (1:1.5 to 1:8 correspondences)depending on mineral compositions (Fig. 6). On the otherhand, the stress magnitude affects mainly the stiffness of the

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 66: Acoustic, elastic and physical properties of

Effects of stress reduction of overconsolidated sands 979

sands; for example, the static–dynamic relationship betweenthe Young’s moduli observed at a 1:2 correspondence still ploton the 1:2 line for all levels of stress magnitude (Fig. 5).

6 D I S C U S S I O N

6.1 Horizontal stress development and stress path

Knowing the magnitude of the minimum horizontal stress isimportant for drilling operations in petroleum explorationand development which is imperative for safe drilling. Mudweight used to counteract subsurface pore pressure shouldnot exceed the minimum horizontal stress to prevent drillingfluid losses into the formation through reopening of existingfractures or invasion into permeable formations. Therefore,predicting horizontal stress, especially in the area where thereis no well control, through the relationship between the K0

and effective stress can be useful.Several plots of stress path (K0, i.e. Fig. 2b–d) show that

K0 under normal consolidated conditions (K0nc) ranges be-tween 0.45 and 0.58 while K0 under overconsolidated (OC)conditions (K0oc) starts from the K0nc points and further in-creases to 1 as a function of overconsolidation ratio (OCR).Mayne and Kulhawy (1983) determine relationship betweenK0 and OCR on the effect of stress history by compiling datafrom hundreds of different soils. The relation is expressed as

K0oc= (1 − sin θ ′) OCRsinθ. (13)

The analysis of unloading stress path that linked betweenK0 and OCR are built on Jaky’s simplified equation (Jaky1948). For normally consolidated materials, OCR = 1, andequation (13) reduces to K0oc = (l – sinθ ′) = K0nc. The θ isfrictional angle. For example, if θ ′ = 30°, K0nc = 0.5.

In the present study, the K0 and OCR relationship ex-hibits two distinct trends separating between a group of highquartz samples and a group of low-medium quartz samples(Fig. 7). The best fit was performed in the form of the powerfunction as in equation (13) on the two groups of samples andarrived at the following correlations:

For high quartz:

K0oc= 0.53 OCR0.53. (14)

For low to medium quartz:

K0oc= 0.47 OCR0.47. (15)

If the material is lacking internal friction angle informa-tion (θ ′), above best fit functions can be used for a simplifiedestimation of K0 if OCR is known.

Figure 7 Polynomial best fit functions established for K0 and over-consolidation ratio (OCR). Thin solid lines aligned across the wholerange of K0 (0.4–0.95) at OCR = 1 and dashed lines are measureddata from all samples in this study. The data trends show a group ofhigh quartz (red thick line) and low quartz samples (black thick line).

The effect of stress release as the reduction of verticaleffective stress causes an increase in horizontal effective stressthrough the increase of K0 value (K0 = σ

′h/σ

′v). Therefore,

this may have implications for safe drilling in uplifted regionbecause the difference between horizontal stress minus porepressure (assuming hydrostatic pore pressure) may be large(wide safely drilling window) as the horizontal stress is highdue to the effect of overburden removal.

6.2 Dynamic and static elastic behaviour

The dynamic elastic moduli derived from wave propagationthrough ultrasonic measurement and static elastic moduli de-rived from deformation loading are different as previouslyreported (Figs 4–6). Several previous studies have explainedpotential causes of the difference (e.g. Jizba and Nur 1990;Fjær 1999; Wang 2000). One of which is the difference instrain amplitude between the two measurements (Martin andHaupt 1994). In the dynamic wave propagation experiment,the elastic strain amplitude is <10−6 while static strain is typ-ically >10−2 (Wang 2000). Fjær (2009) also supported thisassumption on his experiment on dry weak sandstone. Otherplausible causes giving rise to the difference between staticand dynamic moduli were summarized by Wang (2000) andMavko et al. (2009). The suggested cause includes the effect

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 67: Acoustic, elastic and physical properties of

980 S. Narongsirikul, N.H. Mondol and J. Jahren

of pore pressure and testing conditions in open and closedsystems. The dynamic elastic moduli derived from wave prop-agation usually denote an undrained or closed system and areinfluenced by the elastic properties of the pore fluid. This is incontrast to static moduli, where in the experiment, the porepressure is kept constant, thereby representing a drained oropen system. Bhuiyan et al. (2013) shows that by removingthe effect of pore fluids through a Biot–Gassmann approach,static and dynamic moduli become comparable for sandstone.

The static–dynamic elastic moduli relation of overconsol-idated (OC) sand condition in the present study is closer toa 1:1 relationship compared to the normal compacted sandcondition (Figs 4–6). This shows that the sands that previ-ously underwent OC or unloading deformed under smallerlinear elastic strains while normal compacted sands deformedunder significant larger nonlinear plastic strains. Therefore,the small elastic strain amplitude of static moduli of unload-ing is closer to the strain amplitude of dynamic moduli. As aresult, the dynamic–static elastic moduli during unloading arenearly comparable, especially for Young’s and shear modu-lus. This phenomenon supports strain amplitude as plausiblecause of the significant differences observed between static anddynamic properties in unconsolidated sands. This behaviourwas also observed by Fjær (2009) on sandstones and Soneand Zoback (2013) on gas shales. Static and dynamic elasticmoduli of unloading (or OC) can be appropriately comparedwith almost equal strain amplitudes and at identical physicalproperties of the solid rock skeleton. This implies that dur-ing the depletion and injection of a shallow unconsolidatedreservoir experienced uplift (unloading) from overburden re-moval, approximating static modulus from dynamic moduluswill be less difference in the magnitude as the relation is closerto one-to-one correspondence than the normally compactedreservoir.

7 C ONCLUSION

Experimental mechanical compaction of sands integratingloading, unloading and reloading cycles confirms that stressrelease affects the rock physical properties and influences theseismic and mechanical properties. It is shown here that ge-omechanical parameters such as strain and stress are dif-ference between normal compaction and overconsolidation(OC). Stress path (K0) data differ between normal consoli-dated (NC) and OC sediments. The K0 value of approximately0.5 is found for most of the NC sands, but varies during un-loading depending on mineral compositions and textural dif-ferences. The K0 equation can be further simplified and can

be influenced by the material compositions. Acoustic velocityand geomechanical relation is also found to be the differencesbetween two stress conditions. The static moduli of the OCsands are much higher than normal consolidated sands asthe deformation is small (small strain). Dynamic–static elasticmoduli relations more closely correlate for the OC stage com-pared to the normally consolidated stage. The experimental re-sults presented herein are only valid for unconsolidated sandsthat have been compacted and unloaded/reloaded within themechanical compaction domain. The results of the study haveseismic-geomechanics applications for shallow reservoir sandswhere mechanical compaction is the dominant process.

ACKNOWLEDGEMENTS

We would like to thank the Norwegian Research Council(NFR) for funding the BarRock (Barents Sea Rock Properties)project under the program PETROMAKS (Programme for theOptimal Management of Petroleum Resources). We are alsograteful to many NGI personnel, especially Toralv Berre, fortheir dedicated help with sample preparation, experimentalsetup and testing program.

ORCID

Sirikarn Narongsirikul

https://orcid.org/0000-0001-5242-4663

REFERENCES

Avseth P. and Lehocki I. 2016. Combining burial history and rock-physics modeling to constrain AVO analysis during exploration.The Leading Edge 35, 528–534.

Berre T. 2011. Triaxial testing of soft rocks. Geotechnical TestingJournal 34, 61–75.

Bhuiyan M.H., Holt R.M., Larsen I. and Stenebraten J. 2013. Staticand dynamic behaviour of compacted sand and clay: compari-son between measurements in triaxial and oedometric test systems.Geophysical Prospecting 61, 329–340.

Bhuiyan M.H., Kolstø M.I. and Holt R.M. 2011. Effects of stress andstrain on wave velocities in compacted sand-kaolinite and kaolinite-smectite. 73rd EAGE Conference & Exhibition Incorporating SPEEUROPEC 2011, Extended Abstracts.

Bjørlykke K. 2010. Petroleum Geoscience: From Sedimentary Envi-ronments to Rock Physics. Springer

Bowers G.L. 1995. Pore pressure estimation from velocity data: ac-counting for overpressure mechanisms besides undercompaction.Society of Petroleum Engineers 10, 89–95.

Bowers G.L. and Katsube T.J. 2002. The role of shale pore-structureon the sensitivity of wireline logs to overpressure. In: PressureRegimes in Sedimentary Basins and their Prediction, Vol. 76 (eds

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 68: Acoustic, elastic and physical properties of

Effects of stress reduction of overconsolidated sands 981

A.R. Huffman and G.L. Bowers), pp. 43–60. American Associationof Petroleum Geologists.

Casagrande A. 1936. The determination of the pre-consolidation loadand its practical significance. Proceedings of the International Con-ference on Soil Mechanics and Foundation Engineering 3. pp. 60–64. Harvard University Cambridge.

Chuhan F.A., Kjeldstad A., Bjørlykke K. and Høeg K. 2003. Experi-mental compression of loose sands: relevance to porosity reductionduring burial in sedimentary basins. Canadian Geotechnical Jour-nal 40, 995–1011.

Dræge A., Duffaut K., Wiik T. and Hokstad K. 2014. Linking rockphysics and basin history — Filling gaps between wells in frontierbasins. The Leading Edge 33, 240–246.

Dvorkin J. and Nur A. 1996. Elasticity of high-porosity sandstones:theory for two North Sea datasets. Geophysics 61, 1363–1370.

Fawad M., Mondol N.H., Jahren J. and Bjørlykke K. 2011. Me-chanical compaction and ultrasonic velocity of sands with differenttexture and mineralogical composition. Geophysical Prospecting59, 697–720.

Fjær E. 1999. Static and dynamic moduli of weak sandstones. The37th U.S. Symposium on Rock Mechanics, ARMA-99-0675, Vail,CO, 7–9 June. American Rock Mechanics Association.

Fjær E. 2009. Static and dynamic moduli of a weak sandstone. Geo-physics 74, 1942–2156.

Geertsma J. 1973. Land subsidence above compacting oil and gasreservoirs. Journal of Petroleum Technology 25, 734–744.

Grande L., Mondol N.H. and Berre T.K. 2011. Horizontal stressdevelopment in fine-grained sediments and mudstones during com-paction and uplift. 73rd EAGE Conference & Exhibition incor-porating SPE EUROPEC, Vienna, Austria, 23–26 May 2011, Ex-tended Abstracts.

Hatchell P. and Bourne S. 2005. Rock under strain: Strain-inducedtime-lapse time shifts are observed for depleting reservoirs. TheLeading Edge 24, 1222–1225.

Herwanger J.V. and Horne S.A. 2009. Linking reservoir geomechan-ics and time-lapse seismics: predicting anisotropic velocity changesand seismic attributes. Geophysics 74, 13–33.

Holt R.M. 1994. Effects of coring on petrophysical measurements.International Symposium of the Society of Core Analysts PaperSCA9407.

Holt R.M. 1999. Laboratory acoustic measurements for reservoircharacterization: consequences of core alteration. InternationalSymposium of the Society of Core Analysts Paper SCA9926.

Holt R.M., Bauer A., Bakk A. and Szewczyk D. 2016. Stress-pathdependence of ultrasonic and seismic velocities in shale, SEG Tech-nical Program Expanded Abstracts: 3159–3163.

Holt R.M., Nes O–M. and Fjær E. 2005. In-situ stress dependenceof wave velocities in reservoir and overburden rocks. The LeadingEdge 24, 1268–1274.

Jaky J. 1948. Pressure in silos. Proceedings of the 2nd InternationalConference on SMGE. pp. 103–109.

Jizba D. and Nur A. 1990. Static and dynamic moduli oftight gas sandstones and their relation to formation properties.

SPWLA 31st Annual Logging Symposium, Lafayette, LA, 24–27 June.

Martin R.J. and Haupt R.W. 1994. Static and Dynamic Elastic Mod-uli in Granite: The Effect of Strain Amplitude. American RockMechanics Association.

Mavko G., Mukerji T. and Dvorkin J. 2009. The Rock Physics Hand-book: Tools for Seismic Analysis in Porous Media. Cambridge Uni-versity Press.

Mayne P.W. and Kulhawy F.H. 1983. K-OCR relationships in soil.Journal of the Geotechnical Engineering Division 20, 851–869.

Mondol N.H., Avseth P., Fawad M. and Smith T. 2010. Vs pre-diction in unconsolidated sands: physical and Geological controlson shear wave velocity. 72nd EAGE meeting, Expanded Abstracts,351.

Narongsirikul S., Mondol N.H. and Jahren J. 2019a. Acoustic andpetrophysical properties of mechanically compacted overconsoli-dated sands: part 1 – Experimental results. Geophysical Prospecting67, 804–824.

Narongsirikul S., Mondol N.H. and Jahren J. 2019b. Acoustic andpetrophysical properties of mechanically compacted overconsoli-dated sands: part 2 – Rock physics modelling and applications.Geophysical Prospecting 67, 114–127.

Nygard R., Gutierrez M., Høeg K. and Bjørlykke K. 2004. Influenceof burial history on microstructure and compaction behavior ofKimmeridge clay. Petroleum Geoscience 10, 259–270.

Pender M.J. 1978. A model for the behaviour of overconsolidatedsoil. Geotechnique 28, 1–25.

Pettersen Ø. 2007. Sandstone compaction, grain packing and criticalstate theory. Petroleum Geoscience 13, 63–67.

Rickett J., Duranti L., Hudson T., Regel B. and Hodgson N. 2007. 4-Dtime strain and the seismic signature of geomechanical compactionat Genesis. The Leading Edge 26, 644–647.

Røste T., Stovas A. and Landro M. 2005. Estimation of layer thicknessand velocity changes using 4D prestack seismic data. 67th EAGEAnnual Conference and Exhibition, Extended Abstracts.

Sone H. and Zoback M.D. 2013. Mechanical properties of shale-gasreservoir rocks — Part 1: static and dynamic elastic properties andanisotropy. Geophysics 78, D381–D392.

Tao G., King M.S. and Nabi-Bidhendi M. 1995. Ultrasolic wave prop-agation in dry and brine-saturated sandstones as a function of ef-fective stress: laboratory measurements and modelling. GeopgysicalProspecting 42, 299–328.

Terzaghi K. 1943. Theoretical Soil Mechanics. John Wiley & Sons.Wang Z. 2000. Dynamic versus static elastic properties of reservoir

rocks. In: Seismic and Acoustic Velocities in Reservoir Rocks, Vol. 3(eds Z. Wang, and A. Nur), pp. 531–539. Society of ExplorationGeophysicists.

Yin H. 1992. Acoustic velocity and attenuation of rocks: Isotropy,intrinsic anisotropy and stress induced anisotropy. PhD thesis, Stan-ford University, Stanford, CA.

Zimmer M.A., Prasad M., Mavko G. and Nur A. 2007. Seismic ve-locities of unconsolidated sands: Part 1 – Pressure trends from 0.1to 20 MPa. Geophysics 72, 1–13.

C© 2019 The Authors. Geophysical Prospecting published by John Wiley & Sons Ltd on behalf of European Association ofGeoscientists & Engineers, Geophysical Prospecting, 68, 968–981

Page 69: Acoustic, elastic and physical properties of

Expanded Abstract 1

Density/porosity versus velocity of overconsolidated sands derived from experimental compaction

By

Sirikarn Narongsirikul, Nazmul Haque Mondol, and Jens Jahren

75th EAGE Conference & Exhibition (2013)

London, UK, June 2013, We0610

DOI: 10.3997/2214-4609.20130755

5

Publication

Page 70: Acoustic, elastic and physical properties of
Page 71: Acoustic, elastic and physical properties of

Expanded Abstract 4

Depletion-induced reservoir compaction in shallow overconsolidated reservoir

By

Sirikarn Narongsirikul, Nazmul Haque Mondol, and Jens Jahren

EAGE – International Workshop on Geomechanics and Energy (2013)

Lausanne, Switzerland, November 2013, We 01 05

DOI: 10.3997/2214-4609.20131970

8

Publication

Page 72: Acoustic, elastic and physical properties of
Page 73: Acoustic, elastic and physical properties of

International Workshop on Geomechanics and Energy – The Ground as Energy Source and Storage Lausanne, Switzerland, 26-28 November 2013

We 01 05Depletion-induced Reservoir Compaction inShallow Overconsolidated ReservoirS. Narongsirikul* (University of Oslo), N.H. Mondol (University of Oslo andNorwegian Geotechnical Inst) & J. Jahren (University of Oslo)

SUMMARYProduction induced reservoir depletion in unconsolidated reservoirs may result in reservoir compactionand surface subsidence. The potential for such impacts can be assessed by integrating geomechanicalmodels with reservoir properties obtained from laboratory measurements. The response of reservoircompaction to pressure depletion depends on the burial history. This study investigates reservoircompressibility and compaction due to depletion from reservoirs of shallow overconsolidated sands causedby uplift compared to normally compacted sands from experimental mechanical compaction. Static moduliobtained from the experimental mechanical compaction show that uniaxial compressibility of theoverconsolidated sands is less compared to the normally consolidated sands. The reservoir compaction dueto depletion in such a reservoir is therefore less compared to the normally compacted reservoir. Themineral compositions and textural variations do not influence the compressibility in the overconsolidatedcase. The preconsolidation stress associated to the maximum burial depth before uplift is an importantestimation to predict when the pore pressure drawdown will change the compaction behaviour from poro-elastic to plastic. The results can be applied to reservoir compaction during hydrocarbon production inuplifted basins like the Barents Sea.

Page 74: Acoustic, elastic and physical properties of

International Workshop on Geomechanics and Energy – The Ground as Energy Source and Storage Lausanne, Switzerland, 26-28 November 2013

Introduction

Hydrocarbon production induced reservoir depletion may result in reservoir compaction, surface subsidence, horizontal stress changes, and potentially, fault reactivation. This can pose serious consequences for production such as wellbore damage. The potential for such impacts can be assessed by integrating geomechanical models with reservoir properties obtained from laboratory measurements to predict/estimate the magnitude of reservoir compaction of depleting reservoirs.

Rock mechanical laboratory testing allows measurements of both dynamic and static moduli. Dynamic moduli obtained by ultrasonic measurements are a well-known application for time-lapse seismic monitoring studies. Static moduli are of the same importance as this information is often used in wellbore stability, and in-situ stress applications. Uniaxial compressibility, in particular, is a key parameter used to estimate reservoir compressibility associated to depletion during hydrocarbon production. How a reservoir responds to pressure depletion depends on the burial history. Mechanical compaction differs between normally compacted sediments and undercompacted or overconsolidated sediments. This study investigates reservoir compressibility and compaction due to depletion from reservoirs of shallow overconsolidated (uplift/erosion) sands compared to normally compacted sands from experimental mechanical compaction.

Reservoir Compaction

When the horizontal extent of a depleted reservoir is greater than its vertical thickness, reservoir compaction and overburden subsidence can be assumed to occur in the vertical direction. The magnitude of vertical displacement (Δh) can be expressed by the following equation: �ℎ/ℎ = ���� (1) where h is the initial reservoir thickness, Cm is the uniaxial compaction coefficient or compressibility of the reservoir or the overburden rocks, and ∆p is the change in reservoir pressure. If the reservoir is disk shaped, the maximum deformation, max Δh, associated with the pressure drop can be described by Geertsma’s concept (Geertsma 1973). Bruno (1992) simplified Geertsma’ expressions as the following equation:

max �ℎ = 2��(1 − �)�� �ℎ − ��� + (� + ℎ)� +√�� + ��� (2)

where Cm is the uniaxial compaction coefficient of the reservoir or the overburden rocks, ∆p is the change in reservoir pressure, � is the Poisson’s ratio of the reservoir rock, r is the average radius, h is the vertical thickness, and D is the burial depth at the top of the formation. The uniaxial compaction coefficient, Cm, appears in both equations 1 and 2. This reservoir attribute depends on rock physical properties such as mineral compositions and textural differences i.e. sorting. This rock property also depends on the reservoir stress history which governs the compaction behaviours of the depleted reservoir. Goulty (2003) showed that pressure drawdown during reservoir depletion causes the effective stress to increase. In poorly consolidated reservoirs, depending on the burial history, normally buried reservoirs prior to production will be further compacted exhibiting plastic (irreversible) behaviour (Figure 1). If the burial history differs in the way that the sediments are precompacted and later unloaded, poro-elastic behaviour during depletion may be expected (Figure 1). Such compaction or overburden subsidence mechanism may be found in either overpressured or uplifted sediments. The latter is the focus of this study.

Page 75: Acoustic, elastic and physical properties of

International Workshop on Geomechanics and Energy – The Ground as Energy Source and Storage Lausanne, Switzerland, 26-28 November 2013

Figure 1 Schematic representation of compaction mechanism. Normal compaction curve exhibiting plastic behaviour reveals steeper trend compared to poro-elastic unloading-reloading curves.

Experimental Data

Experimental data of six natural sand aggregates studied by Narongsirikul et al. (2013) were utilised in this study. The sand samples contained 36-98% quartz and in addition various amounts of feldspar and clays. The sands were compacted, unloaded and reloaded under uniaxial strain condition (K0). Three loading cycles simulating episodes of uplift/erosion with increasing peak axial effective stresses to the maximum stress at 30 MPa were applied to all samples (Narongsirikul et al. 2013). Figure 2 shows uniaxial strain (a) and porosity (b) plotted versus axial effective stress for one representative sample. Porosity has a reverse relation to the uniaxial strain as the stress is applied. Both plots show different deformation behaviours of the normally compacted sands (solid line with filled symbol) compared to unloaded sands (dashed line with open symbols) as expected.

Figure 2 Deformation behaviours of one sand sample expressed in uniaxial strain (a) and porosity (b) plotted versus axial effective stress. The deformation of normally compacted (solid lines with filled symbols) sands is high compared to overconsolidated sands (dashed lines with open symbols).

Result and Discussions

Uniaxial Compaction Coefficient

Figure 3a shows deformation behaviours plotted for all samples. The compaction of normally compacted sands (solid lines with filled symbols) differs significantly from the overconsolidated sands (dashed lines with open symbols). The overconsolidated sand curves plotted from the final

Increasing effective stress

Dec

reas

ing

po

rosi

ty

0 10 20 300

50

100

150

Axial effective stress (MPa)

Un

iaxi

al s

trai

n (

mS

)

0 10 20 3026

28

30

32

34

36

38

Axial effective stress (MPa)

Po

rosi

ty (

%)

a. b.

Page 76: Acoustic, elastic and physical properties of

International Workshop on Geomechanics and Energy – The Ground as Energy Source and Storage Lausanne, Switzerland, 26-28 November 2013

unloading stage show a small reversal of the uniaxial strain as the stress was released (uplift). Reservoir compaction during hydrocarbon production in an uplifted reservoir caused by an increase in effective stress due to depletion will follow the overconsolidated trends depicted in Figure 3a. Figure 3b shows uniaxial compressibility, Cm, calculated from the slopes of Figure 3a for all samples. The compressibility of the normally compacted sands (filled symbols) is generally higher (greater pore collapse) compared to the overconsolidated sands (open symbols) and varies with quartz contents. However, the overconsolidated sands are not influenced by mineralogical and textural variations. This is seen as the overconsolidated sand uniaxial compressibility is similar for all samples.

Figure 3 Deformation behaviors (a) and uniaxial compressibilities (b) between normally consolidated (filled symbols) and overconsolidated sands (open symbols).

Geomechanical Modelling Example

For simplicity, we assumed the same reservoir parameters for both normally consolidated and overconsolidated reservoirs except for the uniaxial compressibility. This is in order to assess the deformation magnitude between the two reservoir types. The reservoirs have an initial vertical thickness of 200 meters prior to production. Assuming no injection to support reservoir pressure, the pressure continues to decline. The change in reservoir pressure is 15 MPa due to the production. We applied these parameters and the uniaxial compressibilities selected from one of the sand samples with high quartz content (92% quartz) into the equation 1. This results in the reservoir displacement model shown in Figure 4. The uniaxial compressibilities of the sample are 3.18×10-3 and 0.39×10-3

(MPa-1) for normal consolidation and overconsolidation, respectively. The geomechanical model in Figure 4 shows that the same amount of pressure change results in different reservoir displacement magnitudes where the reservoir compaction of the overconsolidated reservoir is significantly less compared to the normally compacted reservoir.

Figure 4 Geomechanical modelling of two reservoirs shows that the reservoir displacement in the overconsolidated (right) is less compared to the normally compacted reservoir (left) due to depletion.

Maximum Burial Depth and Amount of Uplift

In sediments undergoing uplift the preconsolidation stress and the stress reduction are related to the maximum burial depth prior to uplift and the degree of uplift/erosion. Knowing this information allows for a better prediction of the transition from poro-elastic to plastic yielding as pore pressure

0 10 20 300

50

100

150

200

Axial effective stress (MPa)

Un

iaxi

al s

trai

n (

mS

)

40

50

60

70

80

90

40 60 80 1000

1

2

3

4

5

6

7x 10

-3

Quartz content (%)U

nia

xial

co

mp

ress

ibili

ty, (

MP

a-1)

Quartz (%)

a. b.

5

10

15

200

2

4

6

8

Dep

th (

m x

100) Δ

h (m

ete

rs)

Normally compacted reservoir Overconsolidated reservoir

Page 77: Acoustic, elastic and physical properties of

International Workshop on Geomechanics and Energy – The Ground as Energy Source and Storage Lausanne, Switzerland, 26-28 November 2013

continues to drop. The difference between preconsolidation pressure and in-situ pressure may also help predicting how much pore pressure changes will start to trigger plastic deformation resulting in greater porosity loss and lower permeability. Figure 5 explains how porosity loss may occur during production in uplifted basins. Figure 5a shows two reservoirs presently at the same depth but previously buried to different depths. The transition from poro-elastic to plastic yielding may occur faster in Reservoir A compared to Reservoir B as Reservoir A previously was buried at a shallower depth before uplift. Figure 5b shows a similar situation but in this case the reservoirs were previously buried to the same depth but uplifted to different present-day depths. Accelerating pore collapse due to plastic yielding will occur faster in Reservoir A than Reservoir B as Reservoir A experienced lower degree of uplift.

Figure 5 Porosity loss behaviours in uplifted reservoirs varying max.burial depths and uplift degrees.

Conclusions

Static moduli obtained from experimental mechanical compaction of unconsolidated sands show that uniaxial compressibility of the overconsolidated sands is less compared to the normally consolidated sands. The reservoir compaction due to depletion in such a reservoir is therefore less compared to a normally compacted reservoir. The mineral compositions and textural variations do not influence the compressibility in the overconsolidated case. It is important to estimate the preconsolidation stress to delineate when the pore pressure drawdown will change the compaction behaviour from poro-elastic to plastic. The results described herein apply to reservoir compaction during hydrocarbon production in uplifted basins like the Barents Sea.

Acknowledgements

We would like to thank the Norwegian Research Council for the funding of Barents Sea Rock Properties (BarRock) project under the program PETROMAKS. We are also grateful to many NGI personnel for their dedicated help in sample preparation, experimental setup and testing program.

References

Bruno, M. S. [1992] Subsidence-induced well failure. Society of Petroleum Engineers, 149-152. Geertsma, J. [1973] A basic theory of subsidence due to reservoir compaction: the homogeneous case. Trans. Royal Dutch Society of Geologists and Mining Engineers, 28, 43-62. Goulty, N. R. [2003] Reservoir stress path during depletion of Norwegian chalk oilfields. Petroleum Geoscience, 9, 233-241. Narongsirikul, S., Mondol, N. H., Jahren, J. [2013] Density/porosity versus velocity of overconsolidated sands derived from experimental compaction. 75th EAGE Conference & Exhibition, Extended Abstracts, We0610.

Increasing effective stress

Dec

reas

ing

po

rosi

ty

ABPlasticyielding

Increasing effective stress

Dec

reas

ing

po

rosi

ty

BPlasticyielding

A Plasticyielding

a. Different maximum burial depths b. Different amounts of uplift/erosion

Page 78: Acoustic, elastic and physical properties of
Page 79: Acoustic, elastic and physical properties of

Expanded Abstract 5

Velocity anisotropy of unconsolidated sands and its relation to induced stress response

By

Sirikarn Narongsirikul, Nazmul Haque Mondol, and Jens Jahren

76th EAGE Conference & Exhibition (2014)

Amsterdam, The Netherlands, June 2014, Tu P11 16, Volume 2014, p.1 – 5

DOI: 10.3997/2214-4609.20140937

9

Publication

Page 80: Acoustic, elastic and physical properties of
Page 81: Acoustic, elastic and physical properties of

76th EAGE Conference & Exhibition 2014 Amsterdam RAI, The Netherlands, 16-19 June 2014

Tu P11 16Velocity Anisotropy of Unconsolidated Sands andits Relation to Induced Stress ResponseS. Narongsirikul* (University of Oslo), N.H. Mondol (University of Oslo andNorwegian Geotechnical Inst) & J. Jahren (University of Oslo)

SUMMARYVelocity anisotropy is an important parameter in seismic imaging, AVO analysis, and sonic loginterpretation. Isotropic sands or sandstones can behave anisotropically if subjected to anisotropic stress.The variations in velocity anisotropy in sands can also be due to discontinuity of the boundary betweensand grains influenced by mineral and textural differences. Since the velocity anisotropy can be inducedby stress, the velocity anisotropy in rocks can indicate stress anisotropy. This study investigates velocityanisotropy and its relation to induced stress response (stress anisotropy) affected by mineralogy andtextures of seven unconsolidated natural sands by experimental mechanical compaction under uniaxialstrain condition. The results show that velocity anisotropy and its relation to induced stress response aresignificantly affected by mineral compositions and textural differences. Our findings have potentialapplications in velocity anisotropy and induced stress anisotropy in shallow sands where compaction ismainly mechanical. The close correlation between these two anisotropic parameters (velocity and inducedstress response) can also be used for detecting stress anomalies using velocity anisotropy. By inverting thestress anisotropy, induced horizontal stress changes can be obtained and can be useful for many in-situstress applications i.e. fault seal analysis, wellbore stability, and compaction and subsidence studies.

Page 82: Acoustic, elastic and physical properties of

76th EAGE Conference & Exhibition 2014 Amsterdam RAI, The Netherlands, 16-19 June 2014

Introduction

Velocity anisotropy is an important parameter in seismic imaging, AVO analysis, and sonic log interpretation. Sands or sandstones are generally inherently isotropic. Such isotropic rocks subjected to anisotropic stress (i.e. uniaxial) will become transversely isotropic (TI). Since the velocity anisotropy can be induced by stress, the velocity anisotropy in rocks can indicate stress anisotropy (Vega et al. 2006). Understanding velocity anisotropy and stress anisotropy relations can help detecting stress anomalies or predicting induced horizontal stress changes which are important for many in-situ stress applications i.e. fault seal analysis, wellbore stability, compaction and subsidence, etc. Although the sands are much less intrinsically anisotropic compared to shales (Wang 2002), intrinsic anisotropy in sands may also vary greatly due to the presence of discontinuities within the sand, such as the boundary between sand grains (Sayers 2002). Discontinuities may be induced by variations in mineral compositions and textural differences i.e. sorting and grain size. This study investigates velocity anisotropy and its relation to induced stress response (stress anisotropy) as influenced by mineral and textural variations of seven unconsolidated natural sands by experimental mechanical compaction under uniaxial strain condition (K0).

Samples and methods

Seven brine-saturated natural sand samples with varying mineral compositions and textural variations i.e. sorting (Figure 1) were mechanically compacted in a triaxial cell setup at the Norwegian Geotechnical Institute (NGI). The mineral compositions were previously analysed by XRD (Fawad et al. 2011) and can be classified (Figure 1a) using Dott (1964) sandstone classification. The sand samples contained 36-98% quartz and in addition various amounts of feldspar and clays, except the Volcanic Arenite sample where the quartz content is only 4% (Figure 1b).

Rock fragments

Feldspar

ArkoseLithic

Subarkose

0

15

75

Feldspaticgreywacke

50

525

5

1

2

34

5

6

Quartz Quartz wacke

50

5050VolcanicArenite

PhylloareniteCalclithite

MetamorphicSedimentary

Volcanic

7

Sample No.

0 % 25 % 50 % 75 % 100 %

No. 1 Quartz Arenite 1

No. 2 Quartz Arenite 2

No. 3 Subarkose 1

No. 4 Subarkose 2

No. 5 Akosic Arenite

No. 6 Feldspatic Greywacke

No. 7 Volcanic Arenite

Quartz FeldsparClay Other

a. b.

Figure 1 (a) Sandstone classification (Dott 1964) with the samples marked in red. (b) Mineral compositions from Fawad et al. (2011).

The sand samples were isotropically compacted to 0.52 MPa and further loaded under uniaxial strain condition (K0) from 0.52 to 30 MPa vertical effective stress (Figure 2). P- and S- wave velocities were measured both parallel (Vpv, Vsv) and perpendicular (Vph, Vsh) to the applied stress direction, σ’v, using the pulse transmission technique (Figure 2). As the stress in the vertical direction was applied, the horizontal stress change (σ’h) was recorded to observe horizontal stress development during the K0 application. This allows stress anisotropy to be investigated. We observed that S-waves are highly attenuated at low effective stress levels between 5-10 MPa leading to uncertainties in S-

wave velocity picking. It is also worth noting that since the sands were loose with well-connected pores, very little squirt flow affecting velocity dispersion was expected.

σ'v , Vpv, Vsv

K0 conditionσ'v >σ’h εh = 0

σ'h , Vph, Vsh

Figure 2 Schematic of applied stress (σ’v) and induced stress change (σ’h) under uniaxial strain (K0) condition and measurement directions of the vertical and horizontal P- and S-wave velocities.

Page 83: Acoustic, elastic and physical properties of

76th EAGE Conference & Exhibition 2014 Amsterdam RAI, The Netherlands, 16-19 June 2014

Result and discussion

Stress-velocity relationship

Compressional and shear wave velocities and stress relationship can be expressed as a function of quartz content in Figure 3. The horizontal P-and S-wave velocities (Vph, Vsh) are generally lower than the vertical velocities (Vpv, Vsv). Varying mineral compositions resulted in wide velocity range (Both Vp and Vs). The samples with low quartz content (Volcanic Arenite, Akosic Arenite,) show higher velocities compared to the samples with higher quartz content (Quartz Arenite and Subarkosic samples). This can be attributed to more ductile minerals and clays presented in low quartz samples resulting in better packing and alignment of the grains. Higher velocities in such samples also correspond with greater porosity reduction. This can be explained similarly by the presence of ductile minerals i.e. clays and feldspar in such samples permitting more compressibility.

0 10 20 30

1800

2000

2200

2400

2600

Eff. vertical stress (MPa)

P-w

ave

velo

city

(m/s

)

VerticalHorizontal

0 10 20 30400

600

800

1000

1200

Eff. vertical stress (MPa)

S-w

ave

velo

city

(m/s

)

20

40

60

80

100VerticalHorizontal

Quartz (%)

Figure 3 Crossplots of vertical and horizontal P- and S- wave velocities as functions of vertical effective stress and quartz content. Velocity Anisotropy

P- and S-wave velocity anisotropy parameters (Vph/Vpv and Vsh/Vsv) are plotted as functions of effective stress and quartz content demonstrating the influence of stress and mineral constituent differences on velocity anisotropy (Figure 4). The results clearly show that both P- and S- wave anisotropy increase with increasing effective stress. As noted previously S-waves are highly attenuated at low stress levels. Hence, the uncertainty of S-wave anisotropy at low stresses is high. The strongest anisotropy is seen in the Volcanic Arenite sample and the anisotropy becomes weaker systematically with increasing amount of quartz. An exception is observed for the Subarkose 2 sample which contains a high amount of quartz but shows relatively high anisotropy. This may be explained by the fact that the Subarkose 2 sample is poorly sorted. Hence, the degree of discontinuity at the boundaries between the sand grains is enhanced. This shows that the velocity anisotropy can be attributed to not only induced stresses and mineral compositions but also textural differences i.e. sorting.

0 10 20 30

0.85

0.9

0.95

1

Eff. vertical stress (MPa)

Vph/

Vpv

0 10 20 30

0.85

0.9

0.95

1

Eff. vertical stress (MPa)

Vsh/

Vsv

20

40

60

80

100

Picking uncertainty at low stresses

Quartz (%)

Subarkosic 2poorly sorted

Volcanic Arenite

Figure 4 P- and S- anisotropy parameters plotted as functions of effectives stress and quartz content.

Page 84: Acoustic, elastic and physical properties of

76th EAGE Conference & Exhibition 2014 Amsterdam RAI, The Netherlands, 16-19 June 2014

Figure 5 shows the Thomsen’ parameters (Thomsen 1986) calculated for Epsilon (ε) and Gamma (γ). The ε varies from -0.15 to -0.03 but the γ is slightly lower with the range between -0.12 to -0.02. The best fit regression for our data is γ = -0.0052+0.883ε (black line) shown comparison with two published relations. Wang (2002) proposed a functional relationship between ε and γ for the sedimentary rocks compiled from sandstones, shales, coals and carbonates to be γ = -0.01049+0.9560ε (Red line). Tsuneyama and Mavko (2005) found a similar relation γ = -0.0282+1.2006ε for sandstone and shale datasets (Blue line). We find that our best fit is very close to that of Wang (2002). The ε and γ in our tested sand samples are valid for sands buried at shallow depth where compaction is mostly mechanical. The overall ranges of ε and γ are significantly lower compared to clays at similar

depth as expected from data reported by Mondol (2012), ε and γ of silt-clay mixtures range from 0-0.15 to 0-0.62, respectively. However, it is clear that the differences in mineral compositions and textural variations have an effect on the anisotropy in so-called isotropic sandstones. Induced stress response (stress anisotropy) Figure 6 shows that as the applied stress increases in the vertical direction (σ’v), it induced changes of stress in the horizontal direction (σ’h). We plotted the horizontal stress versus vertical stress in Figure 6a. The resulting changes in horizontal stress can be shown in Figure 6b as the ratio between horizontal and vertical stress (σ’h/ σ’v, stress anisotropy). The plot shows that stress anisotropy in all samples are more or less unchanges with the change in applied vertical stress. However, the variations in the induced stress anisotropy deviate clearly with differences in mineral compositions, range from 0.41 to 0.56 (Figure 6b). The samples with high quartz contents (i.e. Quartz Arenite and Subarkosic) exhibit weaker induced stress anisotropy compared to the lower quartz samples. This mineral effect on stress anisotropy also corresponds with the effect on the velocity anisotropy (Figure 4) where high quartz content shows weakly anisotropy in both P- and S-velocities and induced stress.

0 10 20 300

5

10

15

Eff. vertical stress (MPa)

Eff.

horiz

onta

l str

ess

(MPa

)

10 20 30

0.4

0.45

0.5

0.55

Eff. vertical stress (MPa)

Hor

izon

tal/V

ertic

al s

tres

s ra

tio

20

40

60

80

100

Quartz (%)

a. b.

Figure 6 (a) Horizontal and vertical stress relation colour coded with quartz content (b) Ratio between horizontal and vertical stress (induced stress anisotropy). Strong stress anisotropy is found in low quartz content sample.

To investigate velocity and induced stress anisotropy relations, we plotted the relative changes in velocity anisotropy for both P- and S-waves (ΔVp, ΔVs) versus relative induced stress anisotropy (Δσ) shown in Figure 7. Such changes can be defined the same way Vega et al. (2006) applied during their studies in detecting the stress-induced velocity anisotropy of sands where the ΔVp, ΔVs, and Δσ are the differences between the anisotropic parameters in vertical and horizontal directions (σ’v-σ’h, Vpv-Vph, and Vsv-Vsh) relative to the parameters in the direction of the applied stress (σ’v,Vpv,Vsv). For

Figure 5 Comparisons of Thomsen’ parameters, Epsilon (ε) and Gamma (γ), with previous studies (Wang 2002; Tsuneyama and Mavko 2005).

-0.15 -0.12 -0.09 -0.06 -0.03 0-0.15

-0.12

-0.09

-0.06

-0.03

0

Epsilon

Gam

ma

10

20

30

40

50

60

70

80

90

100Our dataOur bestfitWang (2002)Tsuneyama and Mavko (2005)

Quartz (%)

γ = - 0.0052 + 0.8838ε

Page 85: Acoustic, elastic and physical properties of

76th EAGE Conference & Exhibition 2014 Amsterdam RAI, The Netherlands, 16-19 June 2014

example, ΔVp = (Vpv-Vph / Vpv)×100. The resulting plots of the anisotropic parameters color-coded with quartz content (Figure 7) show that the Vp and Vs anisotropy closely correlate with the induced stress anisotropy when considering the effect of mineral differences. This is seen in both plots where low relative Vp and Vs anisotropy of the high quartz samples also show low relative induced stress anisotropy and increasing proportionally with decreasing quartz content, with the exception of one sample marked poorly sorted. This relationship shows that velocity anisotropy can be indicative of stress anisotropy. The plots also show that increasing effective stress (Increasing data point size) in each sample will increase the velocity anisotropy greatly but only change the stress anisotropy slightly. This can be summarised that variations in mineral compositions affect both velocity anisotropy and induced stress anisotropy. But the changes in applied stress, for this particular loading path, will only change velocity anisotropy but not the induced stress anisotropy. The effect of applied stress on induced stress anisotropy will be more significant if different loading paths are applied i.e. unloading as documented in Vega et al. (2006).

40 45 50 55 60-5

0

5

10

15

20

Relative stress anisotropy (%)

Rel

ativ

e Vp

ani

sotr

opy

(%)

40 45 50 55 60-5

0

5

10

15

20

Relative stress anisotropy (%)

Rel

ativ

e Vs

ani

sotr

opy

(%)

20

40

60

80

100

Poorly sortedPoorly sorted

Increasing effective stress

Increasing effective stress

Decreasing quartz contentDecreasing

quartz content

Quartz (%)

Figure 7 ΔVp, ΔVs plotted versus Δσ as functions of quartz content and applied vertical stress.

Conclusions

Velocity anisotropy and its relation to the induced stress anisotropy are significantly affected by mineral compositions and textural differences. As the samples are moving towards mudrock (Dott 1964), indicated by decreasing quartz content, the velocity and stress anisotropy increase. Our findings have potential applications in velocity anisotropy and induced stress anisotropy in shallow sands. The close correlation between these two anisotropic parameters can also be used for detecting changes in stress anisotropy using velocity anisotropy. By inverting the stress anisotropy, induced horizontal stress changes can be obtained and can be useful for many in-situ stress applications i.e. fault seal analysis, wellbore stability, and compaction and subsidence studies.

Acknowledgements

We would like to thank the Norwegian Research Council for the funding of Barents Sea Rock Properties (BarRock) project under the program PETROMAKS. We are also grateful to many NGI personnel for their dedicated help in sample preparation, experimental setup and testing program.

References

Dott, R. L. [1964] Wacke, greywacke, and matrix –What approach to immature sandstone classification? Journal of Sedimentary Petrology, 34, 625-632. Fawad, M., Mondol, N. H., Jahren, J. and Bjørlykke, K. [2011] Mechanical compaction and ultrasonic velocity of sands with different texture and mineralogical composition. Geophysical Prospecting, 59, no. 4, 697- 720 Mondol, N. H., [2012] Velocity anisotropy in experimentally compacted clay-silt and clay-clay mixtures. 82nd SEG Annual Meeting, Las Vegas, USA. Expanded abstract. Sayers, C. M. [2002] Stress-dependent elastic anisotropy of sandstones. Geophysical Prospecting, 50, 85-95. Thomsen, L. [1986] Weak elastic anisotropy. Geophysics, 51, 1954–1966. Tsuneyama, F., and Mavko G. [2005] Velocity anisotropy estimation for brine-saturated sandstone and shale. The Leading Edge, 24, 882–888. Vega, S., Mavko, G., and Prasad, M. [2006] Detection of stress-induced velocity anisotropy in unconsolidated sands. The Leading Edge, 253-256. Wang, Z. [2002] Seismic anisotropy in sedimentary rocks: Part 2 — Laboratory data. Geophysics, 67, 1423–1440.

Page 86: Acoustic, elastic and physical properties of
Page 87: Acoustic, elastic and physical properties of

Conference Paper 1

Experimental insight into uplift effects on seismic velocities and petrophysical properties of sandstones: Implication for

the Barents Sea area

By

Sirikarn Narongsirikul, Nazmul Haque Mondol, and Jens Jahren

NGF Winter Conference (2013)

Oslo, Norway

1 0

Publication

Page 88: Acoustic, elastic and physical properties of
Page 89: Acoustic, elastic and physical properties of

Experimental insight into uplift effects on seismic velocities and petrophysical properties of sandstones: Implication for the Barents Sea area

Sirikarn Narongsirikul1, Nazmul Haque Mondol1,2, and Jens Jahren1 1Department of Geosciences, University of Oslo, Norway

2Norwegian Geotechnical Institute, Oslo, Norway

We present P- and S-wave velocities and corresponding petrophysical properties (total porosity and bulk density) of eight laboratory tested unconsolidated natural sands with different mineralogical compositions and textural variations. The samples were tested at effective stresses from 0.2 up to 30 MPa corresponding to approximately 3000 m subsidence in a sedimentary basin at hydrostatic pore pressure. Three loading cycles were applied to study the influence of pressure reduction on seismic velocities and rock physical properties simulating episodes of uplift in a complex burial history. The results show significant differences in rock physical properties between uplifted (unloaded) and reburied (reloaded) sediments compared to normally compacted sediments. Total porosity, bulk density, P- and S-wave velocities deviate from normal compaction trends during stress release. This can be explained by considering that compaction process is primarily inelastic and that release of stress will release the elastic part of the deformation. The degree of deviation is dependent on the maximum stress that the sediments previously experienced. Moreover, the magnitude of total porosity and bulk density rebound compared to P- and S-wave velocities is less during unloading/uplift. This can be explained by the fact that porosity and density are bulk properties, while velocities are wave propagation phenomena and more sensitive to the changes in microfabric during the release of stress. The results also show that the degree of velocity reduction and porosity increase is a function of uplift for sediments previously buried to the same burial depth. On the other hand, if the sediments experienced the same degree of uplift, porosity loss and velocity increase as a function of past maximum burial depth. This means that empirical P- and S-wave velocity relations for different uplift settings can be established. Such relationships can be used for Vs prediction for sandstones that have experienced uplift/reburial if the burial history is known. The experimental results obtained in this study significantly improve the understanding of velocity anomalies found in sandstones in uplifted basins like those in the Barents Sea area.

NGF Winter Conference 8-10 January 2013

Page 90: Acoustic, elastic and physical properties of
Page 91: Acoustic, elastic and physical properties of

Conference Paper 2

Rock Physics aspects of uplifted sediments - Experimental compaction study

By

Sirikarn Narongsirikul, Nazmul Haque Mondol, and Jens Jahren

NFiP Seminar (2013) Stavanger, Norway

1 1

Publication

Page 92: Acoustic, elastic and physical properties of
Page 93: Acoustic, elastic and physical properties of

Exploration and field development challenges in uplifted Barents Sea – Insights from experimental mechanical compaction study

Sirikarn Narongsirikul1,3, Nazmul Haque Mondol2, and Jens Jahren1

1University of Oslo, Department of Geosciences 2University of Oslo and Norwegian Geotechnical Institute

3Corresponding author: [email protected]

Successful exploration leading to field development in uplifted basins like the Barents Sea requires a good understanding of rock properties. The effect of uplift makes interpretation more difficult since the seismic and petrophysical properties of uplifted rocks differ from the properties commonly observed in normally compacted basins. This study reports experimental mechanical compaction performed on seven brine-saturated natural sand aggregates with varying mineralogical and textural variations. Complex loading, unloading and reloading cycles under uniaxial strain condition (K0) with an effective vertical stress range from 0.5 – 30 MPa were applied to all samples to simulate several episodes of uplift/erosion. P- and S- wave velocities, porosity and density were measured and monitored throughout the experiments. The study outcomes reveal three significant findings that suggest challenges related to interpretation that petroleum companies may encounter during exploration and field development study in the Barents Sea. 1) Rock physics model approach, friable sand model in particular, gives an ambiguous description of rock microtexture (i.e. sorting) when the model is used for uplifted sand data. We found that the uplifted sand data plotted along the model line may not only originate from sorting changes but also by variations in the preconsolidation stress associated with maximum burial depth before uplift. 2) Fluid sensitivity analysis employed using Gassmann fluid substitution on our experimental data shows that for reservoir sands observed at the same current effective stress the fluid sensitivity in uplifted reservoirs is low compared to the normally subsided reservoirs. The sensitivity in uplifted reservoirs decreases with increasing preconsolidation stress related to the maximum burial depth or degree of uplift. This raises awareness during time-lapse seismic monitoring study in the area. 3) Uniaxial compressibility data combined with geomechanical modeling to study effects of pressure depletion on reservoir compaction of producing fields in uplifted reservoirs shows that the reservoir displacement magnitudes of the overconsolidated reservoirs is significantly less compared to the normally compacted reservoirs when considered an equal amount of pore pressure drawdown. The small displacement found in the uplifted reservoir as a result of depletion reveals no cause for alarm. However, petroleum production from uplifted reservoirs may introduce additional complexity in understanding and predicting compaction and subsidence. This is because the initial state of stress prior to production (preconsolidation stress) may be difficult to predict due to a complex burial history (uplift/erosion, reburial) in the Barents Sea.

Page 94: Acoustic, elastic and physical properties of
Page 95: Acoustic, elastic and physical properties of