Vous êtes sur la page 1sur 119

UNIVERSITY OF CALGARY

SHALE –FROM NANOPORE STRUCTURE INVESTIGATION TO PETROPHYSICS

AND RESERVOIR SIMULATION

by

Peng Wu

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

DEPARTMENT OF CHEMICAL AND PETREOLEUM ENGINEERING

CALGARY, ALBERTA

SEPT, 2014

© Peng Wu 2014
Abstract

Despite the burst of shale gas and shale oil production during the last decade, we are still far

from fully understanding shale reservoirs. A pervasive problem is the scarcity of data for

complete evaluation of wells penetrating these types of reservoirs. This observation leads to the

developments presented in this thesis. Contributions are as follows:

1) Comparison of scanning electron microscopy (SEM), transmission electron microscopy

(TEM) and atomic force microscopy (AFM) for studying the internal structure of shale

reservoirs.

2) Development of a new petrophysical dual porosity model for calculating the cementation

exponent m and water saturation in shale petroleum reservoirs.

3) A sensitivity study using a commercial simulator for studying the effect of the basic

petrophysical properties mentioned above and hydraulic fracture parameters. Results

show that the importance of matrix porosity and hydraulic fracture length vary depending

upon the matrix permeability in shale formations.

ii
Acknowledgements

I would like first thank my supervisor, Dr. Roberto Aguilera. Due to my different background, I

know it’s very difficult to get into the graduate program of Chemical and Petroleum Engineering

Department of University of Calgary. Roberto gave me the chance and opened another door for

me. I’m lucky to have Roberto as my supervisor. He is such a knowledgeable, humble person.

With his full support, I am able to finish my master degree in petroleum engineering. Most

importantly, I love my major now and I believe the knowledge I learned will benefit me for my

whole career life.

Secondly, I want to thank everyone in our GFREE research group. They are all friendly and

kind; answer all my questions patiently. That helped me a lot to catch up with my courses. I also

want to thank the members of my thesis defense committee for their comprehensive review and

objective critique of this manuscript.

Also, I want to acknowledge ConocoPhillips, NSERC and AERI for their funding to this project.

Special thanks to microscopy and imaging facility in University of Calgary for their help with

imaging of shale using their instruments.

Lastly, I want also thank my wife, my parents and my little girl for their support and love on my

journey in the past couple years.

iii
This is dedicated to my loving wife, my angel daughter and my dearest parents.

iv
Table of Contents

Abstract ............................................................................................................................... ii
Acknowledgements ............................................................................................................ iii
This is dedicated to my loving wife, my angel daughter and my dearest parents. ............ iv
Table of Contents .................................................................................................................v
List of Tables .................................................................................................................... vii
List of Figures and Illustrations ....................................................................................... viii
Nomencalture…………………………………………………………………….……......xi

CHAPTER ONE: INTRODUCTION ..................................................................................1


1.1 Shale definition ..........................................................................................................3
1.2 Geological background ..............................................................................................4
1.3 Gas resource triangle .................................................................................................5
1.4 Horizontal well and hydraulic fracturing ...................................................................6
1.5 Economics of shale gas ..............................................................................................9
1.6 Technical publications .............................................................................................11
1.7 References ................................................................................................................11

CHAPTER TWO: LITERATURE REVIEW ....................................................................13


2.1 Porosity ....................................................................................................................15
2.1.1 Basic porosity composition .............................................................................15
2.1.2 Volumetric calculation of gas-in-place ...........................................................16
2.1.3 Porosity from lab measurements .....................................................................18
2.1.4 Porosity from logs ...........................................................................................21
2.2 Adsorbed Gas ...........................................................................................................23
2.2.1 Langmuir adsorption .......................................................................................23
2.2.2 Adsorbed gas from lab measurement: canister desorption and adsorption isotherm
..........................................................................................................................24
2.3 Kerogen and TOC ....................................................................................................27
2.3.1 Definition of kerogen and TOC .......................................................................27
2.3.2 TOC and LOM measurement in laboratory.....................................................29
2.3.3 TOC from well logs .........................................................................................32
2.4 Permeability .............................................................................................................32
2.4.1 Permeability of shale gas, tight gas and conventional gas reservoirs..............32
2.4.2 Permeability from pulse decay measurement ..................................................34
2.5 Water saturation .......................................................................................................36
2.6 Grain density ............................................................................................................37
2.7 Mineralogy ...............................................................................................................37
2.8 Summary of shale petrophysics ...............................................................................38
2.9 Alberta shale gas plays ............................................................................................41
2.10 References ..............................................................................................................48

CHAPTER THREE: INVESTIGATION OF SHALE AT NANOSCALE USING SCAN


ELECTRON MICROSCOPY, TRANSMISSION ELECTRON MICROSCOPY AND
ATOMIC FORCE MICROSCOPY ..........................................................................52

v
3.1 Introduction of instruments of nanometer resolution ..............................................52
3.2 SEM, TEM and AFM of Muskwa and Nordegg shale samples ..............................57
3.2.1 SEM .................................................................................................................58
3.2.2 TEM .................................................................................................................63
3.2.3 AFM ................................................................................................................68
3.3 Summary ..................................................................................................................69
3.4 References ................................................................................................................70

CHAPTER FOUR: PETROPHYSICAL MODEL FOR WATER SATURATION


EVALUATION IN TIGHT AND SHALE RESERVOIRS ..72
4.1 Introduction ..............................................................................................................72
4.1.1 Archie’s equation and Pickett plot ..................................................................72
4.1.2 Pickett plot .......................................................................................................74
4.1.3 Dual porosity model by Aguilera ....................................................................77
4.2 Water saturations in Pickett plot using dual porosity model ...................................79
4.3 Water saturation model for shale formations ...........................................................82
4.4 Triple porosity model including inorganic matrix, kerogen and natural fractures. .93
4.5 References ................................................................................................................94

CHAPTER FIVE: UNCERTAINTY ANALYSIS OF SHALE GAS SIMULATION:


CONSIDERATION OF BASIC PETROPHYSICAL PROPERTIES .....................95
5.1 Simulation model .....................................................................................................95
5.2 Discussion ..............................................................................................................103
5.3 Summary ................................................................................................................104
5.4 References ..............................................................................................................104

CHAPTER SIX: SUMMARY, CONCLUSIONS AND RECOMMENDATIONS ........105

APPENDIX A: DERIVATION OF DUAL POROSITY MODEL INCLUDING ..........107

vi
List of Tables

Table 2-1 Shale gas properties of the four main producing shale basins in the US (Rokosh et
al. 2008) ................................................................................................................................ 26

Table 2-2 Methods to measure shale gas petrophysical properties (modified


from(Sondergeld, Newsham, et al. 2010)). ........................................................................... 40

Table 2-3 Summary of estimates of Alberta shale- and siltstone-hosted hydrocarbon resource
endowment with these formations highlighted in generalized stratigraphic chart of
Alberta (Adapted from (Rokosh et al. 2012)). ...................................................................... 42

Table 2-4 Alberta shale formation lithologies ((Rokosh et al. 2012)). ......................................... 45

Table 2-5 Commonly used lithology terms related to shale ......................................................... 45

Table 2-6 TOC of Alberta shale formations (data in this table are
estimated from histograms in (Rokosh et al. 2012).) ............................................................ 46

Table 2-7 Porosity of Alberta shale resource (data adapted from (Rokosh et al. 2012)) . ........... 47

Table 4-1 Core and log data for Haynesville shale case study ..................................................... 87

Table 5-1 Investigated parameters in sensitivity analysis............................................................. 97

vii
List of Figures and Illustrations

Figure 1-1 Gas resource triangle (Aguilera 2008). ........................................................................ 5

Figure 1-2 North American natural gas production (Russum 2009) ............................................. 6

Figure 1-3 Annual Barnett shale natural gas production by well type
(U.S. Energy Information Administration,2010) .................................................................... 7

Figure 1-4 Horizontal well and hydraulic fracturing (TOTAL,2012) ............................................ 8

Figure 2-1 Schematic of shale matrix and porosity composition ................................................ 15

Figure 2-2 Comparison of three lab results using helium expansion method on core
samples(Sondergeld, Newsham, et al. 2010) ........................................................................ 19

Figure 2-3 Langmuir isotherm ...................................................................................................... 24

Figure 2-4 Hydrocarbon potential of different kerogen types (blue: oil, yellow: gas) (Crain ). .. 28

Figure 2-5 Van Krevelen diagram showing chemical evolution of kerogen of different types
at increasing levels of thermal maturity (Glorioso and Rattia 2012). ................................... 28

Figure 2-6 Scales of organic metamorphism. (Adapted from (Hood, Gutjahr, and Heacock
1975)). ................................................................................................................................... 31

Figure 2-7 Determination of the level of organic metamorphism (LOM) based on knowledge
of vitrinite reflectance (Ro) and the cementation exponent, m. The laboratory data is
taken from Hood et al. (1975) (Adapted from (Yu and Aguilera 2011)).............................. 31

Figure 2-8 Permeability (log scale) range of shale gas, tight gas and conventional gas
(Russum 2009) ...................................................................................................................... 33

Figure 2-9 Permeability vs. Porosity crossplot including tight gas (Nikanassin) and shale gas
data including Horn River (HR), soft shales in Canada, Fayettville(F), Barnet (B),
Huron and Marcellus shales in the United States (Adapted from (Aguilera 2010)). ............ 34

Figure 2-10 Apparatus for pulse decay permeability (Javadpour 2012b)..................................... 36

Figure 2-11 Map of Alberta shale gas formations (data from (Rokosh et al. 2012)) ................... 43

Figure 2-12 Comparison between Montney and Banff/Exshaw showing gross isopatch
thickness, porosity thickness (φ×h), and thermal maturity (data from (Rokosh et al.
2012)). ................................................................................................................................... 44

Figure 3-1 Image of TEM and SEM (Medwow) .......................................................................... 55

Figure 3-2 Image of an AFM instrument and schematic assembly (Kumar, Dao, and Mohanty
2008) ..................................................................................................................................... 56

viii
Figure 3-3 SEM images of Muskwa core sample. Samples scanned as received with no
preparation. (a) Typical flocculated clay microfabric. (b) Open microfracture (3~4 um)
(c) Typical clay microfabric in micron holes. ....................................................................... 59

Figure 3-4 SEM image of shale sample (as received) showing nano-fracture ............................. 60

Figure 3-5 SEM BSE image of polished Nordegg sample (EDS scan results of 3 specified
area shown on the left table) ................................................................................................. 60

Figure 3-6 SEM BSE image of Nordegg core sample .................................................................. 62

Figure 3-7 SEM high magnification image of Muskwa sample showing organic material
structure................................................................................................................................. 63

Figure 3-8 Annular dark field (ADF) TEM image of kerogen in Nordegg core sample .............. 65

Figure 3-9 Annular dark field (ADF) TEM image of kerogen in Muskwa core sample .............. 66

Figure 3-10 Annular dark field (ADF) TEM image of nano structures in Muskwa and
Nordegg core samples. (a) Aggregates with shape and morphology like pyrite
framboids, salt crystals and kerogen in Nordegg sample. (b) Nano crystals in Nordegg
sample. (c) Bright field image of clay flake with sub-nano porosity in Muskwa sample. ... 67

Figure 3-11 AFM image of polished Muskwa sample ................................................................. 68

Figure 4-1 Cementation factor in Pickett plot (adapted from (Davis)) ......................................... 74

Figure 4-2 Importance of saturation exponent (n) in Pickett plot. Green lines represent
hydrocarbons and blue means 100% water for the three schemes shown above the
Pickett plot (upper three schemes were adapted from (Davis)) ............................................ 75

Figure 4-3 Crossplot of dual porosity (fracture + matrix) water saturation lines using
Aguilera’s dual porosity model in Pickett a Plot. Dots are from log data of a well in
Wapiti pool, Falher formation, Western Canada Sedimentary Basin (WCSB). Water
saturation lines equal to 100%, 50%, 10% and 3% are plotted for Archie’s single
porosity model and Aguilera’s dual porosity model (φb = 0.04 and 0.025). ......................... 81

Figure 4-4 Water saturation curves plotted for Haynesville shale using dual porosity model
developed in this thesis (φb = 0.04) ....................................................................................... 88

Figure 4-5 Water saturation curves for Haynesville shale using dual porosity model (φb =
0.027) .................................................................................................................................... 88

Figure 4-6 Water saturation curves plotted for Haynesville shale using dual porosity model
developed in this thesis. Data points from available well logs (φb = 0.04) ........................... 89

Figure 4-7 Picket plots integrated with TOC curves (Yu and Aguilera 2011) ............................. 90

ix
Figure 4-8 Sensitivity analysis of water saturation curves in shale reservoir by changing
kerogen porosity φk. ............................................................................................................... 92

Figure 4-9 Sensitivity analysis of water saturation curves in shalereservoir by changing


kerogen resistivity factor c .................................................................................................... 92

Figure 5-1 Top view of model layout for multi-stage hydraulically fractured horizontal well .... 96

Figure 5-2 Hydraulic fracture permeability as a function of distance from well.......................... 98

Figure 5-3 Tornado plot of cumulative gas production showing sensitivity of tested
parameters (matrix permeability 2.5×10-5~ 4×10-4 md). .................................................... 101

Figure 5-4 Tornado plot of cumulative gas production (matrix permeability 4×10-4 ~ 6.4×10-3
md). ..................................................................................................................................... 101

Figure 5-5 Tornado plot of cumulative gas production (matrix permeability 6.4×10-3 ~ 0.1
md). ..................................................................................................................................... 102

Figure 5-6 Order of importance of tested parameters in sensitivity studies. .............................. 102

Figure A-1 Sketch of dual porosity model for shale ................................................................... 108

x
Nomenclature

GIIPTot: total gas initially in place (scf)


GIIPad: adsorbed gas initially in place (scf)
GIIPfr: free gas initially in place (scf)
TOC: total organic carbon (weigh fraction)
φ: total porosity-fractional
φa: adsorbed gas porosity
φb: matrix block porosity attached to the bulk volume of the matrix system
φm: matrix block porosity attached to the whole composite system
φnc: porosity of non-connected vugs attached to the whole composite system
φ : porosity of natural fractures or connected vugs associated with the whole composite system
2

a: tortuosity factor
m: cementation factor of the whole composite system
mb: cementation factor of the matrix
n: saturation Exponent
ν: partitioning coefficient (fracture porosity ratio)
νnc: vug porosity ratio
F: formation Resistivity Factor
Rt : measured True Resistivity
Rw: brine resistivity
R0 : resistivity of formation rock 100% saturated with brine of resistivity Rw
Sw: water Saturation
A: area (acres)
𝐵𝑔 : initial formation volumetric factor
C: unites conversion factor = 1359.7
c: kerogen resistivity factor
GsL: Langmuir storage capcity (scf/ton)
𝑔𝑐: adsorbed gas content (scf/ton)
h: average net thickness (ft)
κ: kerogen conversion factor
KHF: hydraulic fracture permeability
KHFmax: maximum hydraulic fracture permeability
KHFmin: minimum hydraulic fracture permeability
Grad: hydraulic fracture permeability gradient
L: distance to well
p: pressure (psia)
pl: Langmuir pressure (psia)
ρb: formation/bulk density (g/cm3)
ρr: relative density of kerogen
ρs: adsorbed gas density (g/cm3)
ρko: kerogen density (g/cm3)
Ro: vitrinite reflectance
Vl: Langmuir volume (scf/ton)
vker : kerogen volume percentage (volume fraction)

xi
Chapter One: Introduction

Shale gas has gained a lot of attention in the last decade. Due to quick development in drilling

and fracturing technologies, shale gas, formerly considered very difficult if not impossible to

recover, has become one of the hottest energy topics. Being touted as the fuel of the future, the

estimated global shale gas recoverable resource now exceeds the estimate for conventional

natural gas resources. In North America, shale gas and gas from tight sandstone reservoirs are

contributing a significant portion of the total natural gas production.

While oil is still the major source of the global energy market; concerns, although not justified,

are emerging in some quarters about source depletion and environmental impact (Aguilera and

Aguilera 2011). On the other hand, green renewable energy such as solar, wind, and biofuel are

still far from satisfying a major portion of global energy market needs. Nuclear power plants

have significant potential but also latent possibilities of scariest disasters.

As a result, natural gas becomes the ideal candidate to provide the transition from oil to green

energy. Natural gas has the potential to become the dominant energy source in the not too distant

future thanks to thousands of trillion cubic feet (TCF) marketable resources just in North

American alone. The high hydrogen to carbon ratio of natural gas will also help stabilizing world

carbon emissions (Aguilera and Aguilera 2011).

Hydraulic fracturing and horizontal drilling makes it economically possible to exploit shale gas

basins in North America in places such as the Western Canada Sedimentary Basin (WCSB), the

Appalachian basin (Utica and Marcellus shales) and the Fort Worth Basin (Barnett shale).

1
According to Energy Information Administration (EIA) estimates, technically recoverable shale

gas in the United States is around 862 TCF, in Mexico 681 TCF, and in Canada 388 TCF. Total

global recoverable shale gas estimated 6622 T CF exceeds world proven conventional gas

reserves of 6609 TCF . No wonder shale gas has become a hot topic in petroleum conferences

and meetings. But on the other side, protest movements against shale exploitation and hydraulic

fracturing are also growing because of worries about ground water contamination and potential

earthquake risk (Lunan 2011).

Recently, due to low commodity price of natural gas, shale oil and condensate attracted a lot of

interest and capital. While estimated amount of shale oil/condensate resource nowadays is still

volatile. Shale oil/condensate provides great value to companies due to it’s the quality of its

product. Knowledge and experience learned from shale gas can be easily adapted to exploration

and development of these liquid resources from shale reservoir. While this thesis mainly focus

on shale gas reservoir, study results and models in it can be easily extended to shale

oil/condensate reservoir.

Many studies have been carried out aiming at how to access these unconventional reservoirs,

how to improve the drilling rate efficiency, how to optimize hydraulic fracturing procedures,

how to improve cumulative production, how to cut operational cost, and also how to address

safety issues (Aguilera and Harding 2011). However, there is no c ookie-cutter approach when

dealing with unconventional reservoirs. Models proven successfully by data from one or more

reservoirs might fail miserably in other reservoirs.

2
1.1 Shale definition

ERCB defines shale gas as a “lithostratigraphic unit having less than 50% by weight organic

matter, with: less than 10% of the sedimentary clasts having a grain size greater than 62.5

micrometers; and more than 10% of the sedimentary clasts having a grain size less than 4

micrometers” (ERCB). This definition differentiates shale from coal in terms of organic matter

percentage (coal contains more than 50% by weight of carbonaceous material). It also indicates

that shale is defined by particle size, not mineralogy. According to ERCB’s definition, shale

should have less than 10% sand, and more than 10% clay. However, this definition might not be

strictly followed by everyone in industry and academia.

It is important to remember that different people may have different understanding about the

term “shale” (Spencer 2010). It is beyond the scope of this thesis to differentiate various terms

that have been used by petroleum engineers and geologists such as “claystone”, “mudstone”,

“clayey” and “argillaceous”. Sometimes, these terms are used ambiguously and not clearly

defined.

While the exact ERCB percentages might not be widely accepted, the general concept suggested

by ERCB is consistent with most people’s understanding of “shale”. Shale is very fine-grained

detrital rock composed of a mixture of clay minerals, quartz-rich silts and/or fine-grained

carbonates, calcites and dolomites. Other components like pyrite, feldspars and plagioclases also

exist commonly in shale. Shale is also usually considered as fissile, splitting in layers.

3
In this thesis, the shale of interest contains kerogen (a solid mixture of organic compound) from

which hydrocarbons (natural gas, natural gas liquid, and oil) can be produced.

1.2 Geological background

Shales are typically deposited in very slow moving water and are often found in areas of either

marine or lacustrine origins. Clay and/or silt sized particles are carried away farthest from their

point of origin, and deposited with organic matter in anaerobic environments. As more

sedimentary layers are added over great length of time, particles are subjected to overburden

pressure and heat, and become interlocked into a rock with deposition layers. During this stage,

kerogen and bitumen are also produced from organic matter, and hydrocarbons are generated.

With different origins of organic matter, different kerogen types are produced. Freshly deposited

clay has larger than 50% primary porosity. The shrinkage and compaction transform them into a

geological section having as little as one fifth of its thickness, and significantly less porosity.

Different colors of shales are usually associated with minerals containing iron with different

oxidation states. Black shales contain an unusual amount of organic matter (3-15%) compared

with less than 1% for average shales (Hews 2012).

4
Figure 1-1 Gas resource triangle (Aguilera 2008).

1.3 Gas resource triangle

Natural gas resource triangle is shown in Figure 1-1 (Aguilera 2008). Conventional gas, the

easiest energy source were produced the earliest. But these resources are limited and decreasing

rapidly. There are much more unconventional gas resources around the world, but are much

more difficult to produce due to their low permeability and more complex geology and rock

composition. These unconventional gas resources include tight gas, coal bed methane (CBM),

shale gas and gas hydrates. Except for the gas hydrates, others have all been successfully

produced and are contributing a significant portion in natural gas production around the world.

Figure 1-2 (Russum 2009) shows natural gas production profile from 1970 to 2009. Generally, it

clearly shows significant reduction in conventional gas production and corresponding increase in

unconventional gas production. In Canada, the conventional gas production reached the peak in

2000, and decreased after that; whereas unconventional gas exploration and production showed

steady growth. In the US, the unconventional gas production in 2009 almost tripled from

5
production in 2005, t hanks to the explosion of horizontal well and hydraulic fracturing

technologies.

Figure 1-2 North American natural gas production (Russum 2009)

1.4 Horizontal well and hydraulic fracturing

Due to the low permeability of unconventional gas reservoirs, especially those containing tight

gas and shale gas, extraordinary stimulation and completion technology are needed for

commercial successful development and production. Advanced horizontal well drilling allows

operators to drill and set well pipes for two kilometres or more horizontally through the same

rock formation (Figure 1-4). Basically, this tremendously increases the rock contacted by the

well. Here, the net pay becomes the length of the well (typically 1500~ 2500 meter) rather than

just the thickness of the interval (typically 20~100 meter). Figure 1-3 shows the tremendous

shale gas production increase in Barnett shale since 2004. From 2004 to 2009, horizontal wells

6
developed so quickly that in five years it came from emergence to absolute dominance in gas

production contributions. This corresponds to the gas production profile in North America shown

in Figure 1-2.

Figure 1-3 Annual Barnett shale natural gas production by well type

(U.S. Energy Information Administration, 2010)

However, even with horizontal wells, it is not possible to produce hydrocarbons from these

formations without further stimulations. Fracture networks are created by injecting a very high-

pressure mixture of water, sand and additives into the well via an appropriate casing . This is

called hydraulic fracturing or fracking. Starting from the toe, multiple fracs are created for a

single horizontal well. Pressurized water forces the rock to open, creating cracks/fractures,

expanding from the injection points on horizontal wells. Sand or other proppant moves into these

opened cracks/fractures to hold them open once all the water and additives are flowed back. The

volume and distribution of created fracture network will have direct impact on h ydrocarbon

production. Microseismic is used to monitor the hydraulic fracture network volume and

7
distribution, ensuring that the fractures were developed as planned and stayed in the intended

zone. This capability can help to optimize fracturing spacing and well spacing, reduce

completion cost and improve production economics.

The flow back water is treated to remove contaminants it picked up as it flowed through the rock.

Treated water is reused for the next stage fracturing. The drilling and hydraulic fracturing of a

horizontal shale gas well typically require 2 to 4 million gallons of water. So the control of water

flow back and recycle is important for completion cost control and environment conservation.

Figure 1-4 Horizontal well and hydraulic fracturing (TOTAL, 2012)

8
There are hundreds of chemical compounds listed as additives for hydraulic fracturing.

Depending on the conditions of specific wells being fractured, different combinations of these

additives were used, each component serves a specific purpose. Predominant fluids currently

used for fracture treatments are water based with friction-reducing additives, making fracturing

fluid easier to be pumped into the formation. There are also bactericides that contained to protect

the reservoir from bacteria contamination, and oxygen scavengers and other stabilizers to prevent

metal pipe corrosion.

1.5 Economics of shale gas

Natural gas price sunk to a new low level in recent years, making shale gas production

uneconomic. More and more negative opinions regarding shale gas came out all of a sudden.

While most comments are questioning the profitability of shale gas, others are attacking the

environment impact of hydraulic fracturing. It is interesting to see that opinions can shift so

quickly with economics and energy price.

Despite all the negative opinions, it’s hard to argue that shale gas and heavy oil/oil sand are the

two biggest unconventional energy resources. Without successful exploration of these

unconventional resources, it will be difficult for us to face the fact that conventional gas and oil

are running out in the next 50 to 100 years, or even shorter period considering the economic and

population growing speed in the last century. We should feel blessed with these vast

unconventional resources. Development of carbon free energy still has a long way to go, which

will probably take many generations to complete. Unconventional resources will provide us the

9
energy we need in the transition time between conventional energy and the future carbon free

energy.

Due to the poor reservoir quality comparing to conventional oil and gas reservoir (extremely

high viscosity of heavy oil/oil sands, very low permeability of shale rock), much more energy is

needed to produce these unconventional hydrocarbons from the underground. High energy input

means high capital/operation cost and more greenhouse gas emission. It’s not fair to compare

shale gas with conventional gas in terms of profitability and environment impact. It’s more

reasonable to compare shale gas to heavy oil/oil sands. Between these two unconventional

resources, shale gas is the one which is more widely distributed globally and with much less

environmental impact. Recently, due to the low price of natural gas, shale oil/natural gas liquid

have become new targets of natural gas operators, and have been commercially produced

successfully. Technologies and experiences in shale gas industry can all be directly applied to

shale oil/natural gas liquid.

Continuous efforts are still being inputted in unconventional resources exploration and

production, aiming to lower the cost and improving the recovery efficiency. At the same time,

the voice should also be made from industry and academia about the global importance of shale

gas/oil. Comprehensive investigation and research should be made to relieve public speculations

about ground water pollution, earth quake concern and natural gas leaking. Corresponding

precautions and prevention methods should be carefully planned and designed.

10
1.6 Technical publications

Parts of the results of this research have been presented at the following international

conferences of the Society of Petroleum Engineers (SPE):

• "Investigation of Gas Shales at Nanoscale Using Scan Electron Microscopy,

Transmission Electron Microscopy and Atomic Force Microscopy," SPE-159887-PP

presented at the 2012 SPE Annual Technical Conference and Exhibition (ATCE) in San

Antonio, Texas (8 -10 October 2012) by Peng Wu and Roberto Aguilera.

• "Uncertainty Analysis of Shale Gas Simulation: Consideration of Basic Petrophysical

Properties,” paper SPE 167236-MS presented at the 2013 SPE Unconventional Resources

Conference - Canada held in Calgary, Alberta, Canada (5-7 November, 2013) by Peng

Wu and Roberto Aguilera.

1.7 References

Aguilera, R. 2008. Natural gas production from tight gas formations: A global perspective Proc.,
19th World Petroleum Congress, Spain.

Aguilera, Roberto F., Roberto Aguilera. 2011. World natural gas endowment as a bridge towards
zero carbon emissions Technological Forecasting & Social Change.

Aguilera, Roberto, Thomas G. Harding. 2011. GFREE research program in the Schulich school
of engineering at University of Calgary. Proc., SPE 147282, paper presented at Annual Technical
Conference in Denver, CO, USA

ERCB. setion 1.020(2)27.1 of Oil and Gas Conservation Regulations (Reprint).

Hews, Peter. 2012. T itle. Shales, Siltstones and Unconventional Reservoirs (Geological
Workshop).

Lunan, Dale. 2011. Hot summer. Unconventional gas guide, Oct. 2011.

Russum, D. 2009. Evaluating Unconventional Gas, AJM.

11
Spencer, R.J., P.K. Pedersen, C.R. Clarkson et al. 2010. Shale Gas (series of articles) Canadian
Society of Petroleum Geologists, Reservoir 37 (8-10).

TOTAL. 2012. Production techniques, www.total.com.

U.S. Energy Information Administration, http://www.eia.gov/analysis/studies/worldshalegas/.

12
Chapter Two: Literature Review

Despite the burst of shale gas production during the last decade, we are still far from full

understanding of shale gas reservoirs. While advanced drilling, stimulation, fracturing and

completion technologies are nowadays becoming more mature, our limited knowledge of shale

rock properties has prevented us from guiding these technologies to their maximum efficiency

and full potential, and at lower cost if possible. The extreme complexity and high heterogeneity

of shale gas reservoirs make them difficult to evaluate, model and simulate.

Importance of petrophysics is the same for shale gas as for any other types of reservoir. At early

stage of oil and gas exploration, petrophysics provides the basic reservoir property inputs for

volumetric hydrocarbon estimation, which is crucial for initial resource evaluation, financial

decision, and risk management. Then, in the development stage, petrophysics is incorporated into

reservoir simulation for more accurate production forecast, well design, and history match and

production scheme optimization. Petrophysics is sometimes treated as an “art” due to its

complexity and uncertainty. Basic properties such as porosity, water saturation, hydrocarbon

saturation and permeability are extracted from well logs, core analysis and more recently from

drill cuttings when log and core data are scarce (Ortega 2012).

For conventional oil and gas wells, or even in heavy oil and oil-sands wells, there are standard

core analysis procedures and routine well logs. With reliable core data available, petrophysics

models can be calibrated, thus the confidence level for reservoir property input is relatively high.

But there are still many uncertainties due to problems associated with complex lithology, large

13
variation in grain distribution, and also many rock structures which are very difficult to quantify

such as natural fractures, vugs, burrows, laminated and dispersed clay.

As for shale reservoirs, only a small percentage of the wells have cores collected and analysed

due in some cases to the formation depth and corresponding high cost. Unfortunately, even when

cores are available, there are no standard industrial analyses protocols. We face a huge challenge

in shale petrophysics due to complex shale rock structure, very limited information, and less

confidence in core data for correlations.

Other than porosity, permeability and water saturation, we also need to know about kerogen

volume and total organic carbon (TOC) for reasonable volumetric estimation and reservoir

simulation of shale reservoirs. Usually there are not enough wells with both complete well log

sets and core data to establish a good c orrelation database for the target shale reservoir. Less

input available, rare and less reliable core data for correlations, and more output required are

some of the facts of shale petrophysics.

Non-standard lab measurement protocols are applied by different service companies; different

petrophysical models for different reservoirs are used to relate core data with log data; different

fluid flow and numerical models, not designed originally for shale reservoirs, are used for

simulating these types of reservoirs. All of these problems cause huge uncertainties in shale

reservoir evaluation, simulation and forecast. History matching of past production lacks full

meaning if there is no confidence about the most basic reservoir properties.

14
In this chapter, we summarize current available methods to evaluate shale gas petrophysical

properties. By doing so, we highlight that uncertainty of petrophysical evaluation of shale gas

reservoirs must be recognized with extra care and treated correspondingly. Basic petrophysical

properties of Alberta shale formations are also summarized at the end of this chapter.

2.1 Porosity

2.1.1 Basic porosity composition

Figure 2-1 Schematic of shale matrix and porosity composition

Shale gas composition can be divided into kerogen and inorganic matrix. Inorganic matrix can be

further divided into clay and non-clay minerals. The petrophysical model shown in Figure 2-1

can be used as a b asis for understanding the porosity concept in shale gas reservoirs, and for

visualizing total gas content estimation. Natural fractures, which are not shown in this schematic,

can be included in inorganic matrix, or they can be considered as the third component besides

15
kerogen and inorganic matrix. Natural fracture porosity associated with the whole matrix (φ2) is

hard to quantify, and is usually considered to occupy a small percentage of total porosity.

Commonly used porosity measurements with helium on c rushed core samples are not able to

measure natural fracture porosity. Under favourable conditions, estimations can be made from

logs and well testing data. In commercial numerical simulators like GEM (from CMG), the

permeability/conductivity of natural fractures is considered to increase significantly after

hydraulic fracturing. This is reasonable considering reopening of closed or partially cemented

natural fractures. In this case, natural fractures are actually considered as part of the hydraulic

fracture network. The porosity value input for natural fractures become much less important than

the permeability values.

2.1.2 Volumetric calculation of gas-in-place

For shale gas in-place calculations, Equation 2-1 adapted from conventional gas need to be

modified to include adsorbed gas. Equations 2-2 to 2-5 have been used for volumetric calculation

of total original gas-in-place, free gas in place and adsorbed gas in place (Cui, Bustin, and Bustin

2009) (Ambrose et al. 2010, Aguilera 2010). Introduction of adsorbed gas porosity (φa) accounts

for the volume taken by adsorbed gas onto kerogen, which could be a significant factor.

𝟏
𝑮𝑰𝑰𝑷𝑻𝒐𝒕 = 𝟒𝟑𝟓𝟔𝟎 ∗ 𝑨 ∗ 𝒉 ∗ 𝝓 ∗ (𝟏 − 𝑺𝒘 ) ∗ 𝑩 2-1
𝒈

The above equation is good for conventional gas reservoirs but questionable for shale gas

reservoirs.

16
𝑮𝑰𝑰𝑷𝑻𝒐𝒕 = 𝑮𝑰𝑰𝑷𝒂𝒅 + 𝑮𝑰𝑰𝑷𝒇𝒓 2-2

𝑷
𝑮𝑰𝑰𝑷𝒂𝒅 = 𝑮𝒔𝑳 𝑷+𝑷 ∗ 𝝆𝒃 ∗ 𝑨 ∗ 𝒉 ∗ 𝑪 2-3
𝑳

𝟏
𝑮𝑰𝑰𝑷𝒇𝒓 = 𝟒𝟑𝟓𝟔𝟎 ∗ 𝑨 ∗ 𝒉 ∗ 𝑩 [𝝓(𝟏 − 𝑺𝒘 ) − 𝝓𝒂 ] 2-4
𝒈

𝝆 𝑷
𝝓𝒂 = 𝟏. 𝟑𝟏𝟖 × 𝟏𝟎− 𝟔 ∗ 𝑴𝒘 ∗ 𝝆𝒃 (𝑮𝒔𝑳 𝑷+𝑷 ) 2-5
𝒂 𝑳

where:
𝐺𝐼𝐼𝑃𝑇𝑜𝑡 : total gas initially in place (scf)
𝐺𝐼𝐼𝑃𝑎𝑑 : adsorbed gas initially in place (scf)
𝐺𝐼𝐼𝑃𝑓𝑟 : free gas initially in place (scf)
𝐴: area (acres)
ℎ: average net thickness (ft)
𝑆𝑤 : total initial water saturation
𝐵𝑔 : initial formation volumetric factor
𝜙: total porosity
𝜙𝑎 : adsorbed gas porosity
GsL : Langmuir storage capacity (scf/ton)
𝑝 : pressure (psia)
𝑝𝐿 : Langmuir pressure (psia)
𝜌𝑏 : formation density (g/cm3)
𝜌𝑠 : adsorbed gas density (g/cm3)
𝐶 : unit conversion factor = 1359.7

There are also different opinions on how to evaluate shale gas reservoirs. Glorioso and Rattia

(Glorioso and Rattia 2012) suggest that Equation 2-1 should be used for calculating total gas in-

place in shale reservoirs, arguing that logs read “all” the pore space of the rock occupied by any

type of fluid, including those occupied by adsorbed gas. In the author’s opinion, this argument

does not apply as regular logging tools are not able to account for the adsorbed gas. Total

original gas in place calculated using Equation 2-2 to Equation 2-4 can be significantly higher

than using Equation 2-1. The difference has been reported at 20~38% percent in some example

calculations (Ambrose et al. 2010, Aguilera 2010). Data input in Aguilera’s example came from

17
published petrophysical data from some shale formations. The large surface area of kerogen due

to its nanopore structure indicates that volume taken by adsorbed gas can be a significant part of

the total porosity, and should not be neglected.

2.1.3 Porosity from lab measurements

Accuracy of porosity measurements is important as it has a significant impact on initial

economic evaluation of shale gas resources. Crushed-rock GRI methods using helium expansion

are generally used for core analysis of shales. However, significant differences have been found

when comparing measured properties from the same samples sent to various commercial

laboratories (Passey et al. 2010, Sondergeld et al. 2010).

18
Figure 2-2 Comparison of three lab results using helium expansion method on core

samples(Sondergeld, Newsham, et al. 2010)

Lack of standard protocols throughout industry laboratories makes the basic inputs for

petrophysical models and shale gas simulation very uncertain. The least the operator can do is to

understand protocol details behind every important property measured in the laboratory and to

avoid the use of laboratories unwilling to reveal detailed information about the measurement

procedure. One example (Sondergeld et al. 2010) is shown above (Figure 2-2).

19
Plug samples from the same general position in the core were delivered to three commercial

laboratories. The experiments were done on s amples “as received” (AR). All laboratories used

helium expansion method to measure AR porosity. Laboratory 1 crushed and sieved the samples,

whereas laboratory 2 crushed but did not sieve the sample before testing. Laboratory 3 measured

porosity on the core cylinder. More details of how core cylinders extraction, preparation and

storage can be found in the original paper (Sondergeld et al. 2010).

Some will argue that even from the same core, plug samples could have different properties due

to high heterogeneity. But in this case, differences between results from these three laboratories

are due to different measurement protocols. This can be easily observed from the general trend of

measured porosity values: lab 1 > lab 2 > lab 3. These differences are quite significant as shown

in Figure 2-2. In uncrushed core cylinder, helium is not able to enter nanopores and non-

connected vugs. On the other hand, more nanopores networks are exposed in crushed samples.

This explains why lab 3 measurements were lower than the other two. Smaller, sieved off

powder/particles lose inter-particle porosities, so they have less porosity than bigger remaining

particles. That is probably the reason causing lab 1 m easured values to be higher than results

from Lab 2.

Dry porosities can also be measured on shale samples after fluid extraction. Dry porosity is

usually higher than AR porosity as extraction will remove clay bound water, leaving more pore

spaces for helium expansion. As helium expansion method is based on Boyles Law, the gas

adsorption is not considered for porosity and permeability measurement on c rushed samples.

Equation 2-2 to Equation 2-4 should be used to estimate the total gas in the shale sample. High

20
pressure mercury injection capillary pressure (MICP) has also been used for porosity

measurements. Pressures up t o 60, 000 ps ia have been used. With such high pressure, there is

still a limitation which only allows mercury entering pores with throats bigger than 2 nm

(Sondergeld, Ambrose, et al. 2010).

2.1.4 Porosity from logs

Porosity cannot be routinely estimated from conventional well logs due to complex lithological

components and existence of kerogen. Even with introduction of correction factors, sonic,

density and neutron logs are not generally accepted for porosity evaluations. Quite a few studies

have been done using a complicated set of logs to evaluate porosity, TOC, kerogen volume, grain

density, free hydrocarbon and water saturation (Quirein et al. 2010, Ramirez et al. 2011, Quirein

et al. 2012). These logs include standard neutron, sonic, density, resistivity, geochemical, nuclear

magnetic resonance (NMR), dielectric logs and more. The interpreted results were compared

with core measurements to prove their petrophysical models and work flow.

NMR logs have been found to provide good match to core porosity (Quirein et al. 2012, Ramirez

et al. 2011, Sigal and Odusina 2011). NMR is sensitive to hydrogen nuclei, which is abundant in

water and hydrocarbon. There are also hydrogen nuclei present in minerals such as antigorite

(Mg3Si2O5(OH)4), kaolinite Al2Si2O5(OH)4, Biotite (K(Mg,Fe)3(AlSi3)O10(OH)2), and illite

((K,H3O)(Al,Mg,Fe)2(Si,Al)4O10[(OH)2,(H2O)]).

However, hydrogen nuclei in these minerals are locked in a lattice, and have much shorter

relaxation time compared to fluids. As a result, they are not detected by NMR logging tools.

21
NMR signal amplitude is proportional to the number of hydrogen nuclei present and is calibrated

to give porosity, free from lithology effects. More excitingly, the relaxation time of NMR signal

can be interpreted to give water, oil and gas porosity respectively, and even clay bound water and

capillary bound water saturations (Kenyon et al. 1995).

Together with standard logging tools, we are able to interpret total porosity, kerogen volume

percentage, free gas porosity, oil porosity, water saturations and irreducible water saturations.

While NMR log cannot completely substitute core analysis, it provides a compromised

alternative when additional information is needed but the budget does not allow paying for

cutting and analyzing cores.

Geochemical logs measure elemental weight fractions, which can be interpret into mineral

compositions. This together with standard and NMR logging tools provide another workflow to

interpret almost all the necessary petrophysical output of shales.

But the reality is that in most wells there is neither core, nor NMR nor geochemical logs. Thus

the research presented in this thesis concentrates on developing some correlations and models to

estimate all the necessary petrophysical properties when the specialized data are scarce or absent.

Chapter 4 will provide a new developed model to derive water saturations and the cementation

factor of shale based on standard log data. Case studies will illustrate the complete work flow for

evaluating porosity, TOC, cementation factor and water saturation.

22
An important question that needs to be considered carefully is what the porosity values from core

analysis and/or log evaluation really represent. It is important to understand the mechanism and

limitation of various core measurement protocols and logging tools. Petrophysical models should

not only be based on da ta correlations between core and well logs, but also on comprehensive

understanding of shale rock lithology and structure.

2.2 Adsorbed Gas

2.2.1 Langmuir adsorption

A significant portion of the gas in shale reservoirs can be adsorbed to kerogen. Kerogen has a

highly porous structure with pores varying in size from 1~2 nm to sub-micron. These micro-

/nanopores provide a large surface area for gas adsorption. Langmuir isotherm has been widely

accepted to model adsorption in coal bed methane and shale gas reservoirs. A general expression

of Langmuir isotherm used in petrophysics has been presented by Lewis (Lewis et al. 2004)

(Figure 2-3). The deciding factors on t he Langmuir curve (𝑉𝑙 and 𝑃𝑙 ) are dependent on

kerogen/shale properties, gas species/composition and reservoir temperature.

𝑽𝒍 ×𝑷
𝒈𝒄 = 2-6
𝑷+𝑷𝒍

where:
𝑔𝑐: adsorbed gas content (scf/ton)
𝑃 : reservoir pressure (psia)
𝑉𝑙 : Langmuir volume (scf/ton)
𝑃𝑙 : Langmuir pressure (psia)

23
Figure 2-3 Langmuir isotherm

2.2.2 Adsorbed gas from lab measurement: canister desorption and adsorption isotherm

Adsorbed gas can be measured by canister desorption. This standard technique for coal bed

methane has now been widely accepted for shale gas. Fresh core sample from well is directly put

in an airtight sealed container, and gas desorbed is measured as a function of time at reservoir

temperature and atmospheric pressure. Total adsorbed gas content is the summation of three

components: lost gas, desorbed gas and crushed gas. Lost gas is the gas volume desorbed from

core sample before the core is sealed in the canister. Crushed gas is gas remained in the core

sample by the end of desorption measurement. It is measured by crushing the sample into fine

particles to accelerate the desorption process (Shtepani et al. 2010). In real cases, the measured

Langmuir curve may exhibit irregular shapes, making estimation of lost gas from extrapolation

difficult.

24
Following canister desorption analysis, adsorption isotherm experiment is used to determine gas-

storage capacity of shale using crushed core samples. This experiment measures gas storage

capacity related to pressure and could also be used to estimate gas saturation in the reservoir. For

accurate measurement, experiment should be performed using gas mixture with compositions as

detected in canister desorption process. Crushed sample size should also be carefully controlled.

Similarly as in porosity experiment, the importance of crushed core sample size, the control sieve

size need to be carefully investigated to find out the appropriate size to accurately represent

properties of the shale in reservoir condition.

If isotherm is only measured for one core sample, then corrections need to be made for the

formation intervals due to variable TOC levels. For a single well and one formation, the level of

maturity and kerogen type are the same and the adsorbed gas content is proportional to the TOC

(Glorioso and Rattia 2012).

𝑻𝑶𝑪
𝑽𝒍𝒄 = 𝑽𝒍𝒕 𝑻𝑶𝑪𝒍𝒐𝒈 2-7
𝒊𝒔𝒐

where:
𝑉𝑙𝑐 : Langmuir volume corrected for different TOC levels
𝑉𝑙𝑡 : Langmuir volume measured at isotherm temperature (reservoir temperature)
𝑇𝑂𝐶𝑙𝑜𝑔 : TOC level established by logging (w%)
𝑇𝑂𝐶𝑖𝑠𝑜 : TOC level used to measure Langmuir isotherm (w%)

For the United States cases shown in Table 2-1, adsorbed gas has a significant contribution to

total gas volumes. Accurate measurement of adsorbed gas is very important for both volumetric

estimation of total-gas-in place and reservoir simulation. The amount of adsorbed gas also has

influence on the measurement of total gas porosity and free gas porosity. The volume taken by

25
Table 2-1 Shale gas properties of the four main producing shale basins in the US (Rokosh et al.
2008)

adsorbed gas must not be neglected. The influence of adsorbed gas on b oth log readings and

porosity measurements using crushed core samples is still not clear. More comprehensive studies

about porosity, free gas, and adsorbed gas need to be done. Reservoir simulation inputs also need

to be carefully considered. If Langmuir adsorption model has already been added as a component

26
in the simulator, then the porosity to be used should not be the total porosity, instead, free gas

porosity should be used.

2.3 Kerogen and TOC

2.3.1 Definition of kerogen and TOC

Kerogen, a mixture of hydrocarbons, is the insoluble portion of organic matter in sedimentary

rock. It is not soluble in toluene and other normal organic solvent. The soluble portion is known

as bitumen. Soluble bitumen has been found in “oil shale”. Kerogen rich shales with insufficient

grade of thermal maturity to release hydrocarbons give rise to superficial deposits called “oil

shales”, which must be differentiated from “shale oil”. (Glorioso and Rattia 2012). Most

common extraction methods for oil shale is surface mining followed by pyrolysis, which

converts kerogen in oil shale into crude oil. Oil shale, analogous to oil sand, with huge potential

resource but much more difficult to extract is beyond discussion of this thesis. For shale

gas/shale oil discussed here, we can consider all organic material as insoluble kerogen.

Kerogen can be classified according to the source of the material (Figure 2-4). Kerogen in shale

gas is type II, whereas coal mostly contains type III kerogen. Van Krevelen diagram (Figure 2-5)

is the most commonly used diagram to classify the kerogen types in terms of their hydrogen

index and oxygen index. During the process of generating hydrocarbons, both C/H ratio and C/O

atomic ratio decrease during the process of generating hydrocarbons. Oxygen is lost primarily as

CO2 and H2O.

27
Figure 2-4 Hydrocarbon potential of different kerogen types (blue: oil, yellow: gas) (Crain 2010 ).

Figure 2-5 Van Krevelen diagram showing chemical evolution of kerogen of different types at

increasing levels of thermal maturity (Glorioso and Rattia 2012).

TOC, total organic carbon is defined as the amount of carbon bound in organic compounds in

shale. Other elements presented in kerogen such as hydrogen, oxygen, nitrogen and sulphur are

not included in the TOC, so there is a factor (𝜅) to account for this difference. Although this

28
factor can be slightly different for kerogen of different types and maturity, 1.2 is generally a

reasonable value (Lewis et al. 2004).


𝑽𝒌𝒆𝒓 ∙𝝆𝒌𝒐
𝑻𝑶𝑪 ∙ 𝜿 = 2-8
𝝆𝒃

where:
𝑉𝑘𝑒𝑟 : kerogen volume percentage
𝜌𝑘𝑜 : kerogen density (g/cm3)
𝜌𝑏 : formation/bulk density (g/cm3)
𝜅 : conversion factor

2.3.2 TOC and LOM measurement in laboratory

TOC is usually measured by Rock-Eval pyrolysis in the laboratory. The Rock Eval (RE)

pyrolysis method consists of a programmed heating of crushed core samples at a s eries of

temperatures in a pyrolysis oven with an inert atmosphere (helium). Not only TOC can be

measured, but also amounts of free hydrocarbons (S1), hydrocarbons generated through thermal

cracking (S2), and release of trapped CO2 (S3). All can be measured selectively and

quantitatively. From S1, S2 and S3, hydrogen index (HI) and oxygen index (OI) can be derived,

which can help to identify kerogen type and quantify level of organic maturity (LOM).

The LOM scale was developed by Hood et al. (Hood, Gutjahr, and Heacock 1975) to replace

Suggate’s coal rank number (Suggate 1959) used in New Zealand (Tertiary-Cretaceous). LOM

can be obtained by measuring vitrinite reflectance (Ro) in the laboratory. Crushed core or drill

cuttings samples are mounted in cold-setting resin and polished as received, and measured under

a microscopic vitrinite reflectance measuring system. The relationship between LOM and

vitrinite reflectance has been summarized by Hood based on coal data (Figure 2-6). Plots of Ro

and LOM based on data from Figure 2-6 are widely accepted and used for both coal and shale.

29
For example, a vitrinite reflectance (Ro) value of 1.6% from core data is equivalent to a LOM

value of about 11 to 12 calculated from the Equation 2-9 (Yu and Aguilera 2011).

1/𝑚
%𝑅𝑜
𝐿𝑂𝑀 = 8.18 � %𝑅𝑜 28.45
� 2-9
0.59+0.41�1− exp�− ��
0.36

The above equation was developed empirically on t he basis of laboratory data published by

Hood et al(Hood, Gutjahr, and Heacock 1975). An extension has been made by considering that

the exponent m is approximated by the cementation exponent in Archie’s equation. A graphical

solution of Equation 2-9 is presented in Figure 2-7. Preliminary results indicate that, in general,

smaller levels of organic metamorphism in shales lead to larger values of m. Additional research

is being conducted along these lines.

Accurate estimation of LOM is important for TOC quantification from log readings. Crossplots

of hydrogen index (HI) vs. Ro and HI vs. Tmax (the temperature at which maximum rate of

generation of hydrocarbons occurs during pyrolysis) have been used to refine kerogen type and

to assess maturity with respect to the oil and gas “windows”. Both HI and Tmax can be obtained

from Rock-Eval pyrolysis. Then the depth plots of Ro and Tmax can be used to spot the oil and

gas top for specific wells, or to locate sweet spots for horizontal wells.

30
Figure 2-6 Scales of organic metamorphism. (Adapted from (Hood, Gutjahr, and Heacock 1975)).

20

18

16
LOM (Level of Organic Methamorphism)

14

12
Laboratory
10 m = 1.47
m = 1.60
8
m = 1.80
6 m = 2.00

0
0 0.5 1 1.5 2 2.5 3 3.5 4
% Rov

Figure 2-7 Determination of the level of organic metamorphism (LOM) based on knowledge of

vitrinite reflectance (Ro) and the cementation exponent, m. The laboratory data is taken from Hood

et al. (1975) (Adapted from (Yu and Aguilera 2011)).

31
2.3.3 TOC from well logs

There are several ways to estimate TOC from well logs: (1) Advanced geochemical logs can be

used to quantify kerogen volume percentage and TOC. (2) Density logs can be used to establish

a correlation between TOC and density. Sometimes correlations from other types of logs can also

be used. (3) Passey’s ∆LogR method (Passey et al. 1990).

The first method generally should produce more reliable results than the other two. However,

geochemical logs are only performed in a very limited number of wells due to their high cost.

Correlations established from the second method are not universal, and they are usually limited

to nearby analogue wells in the same formation, if geology and petrophysical properties are not

highly heterogeneous. Corrections from laboratory measured density and density logs need to be

carefully considered. The third option, developed by Passey, has been widely tested and

generally provides good results. The method was originally developed for source rocks in the oil

maturity window (Ro=0.5-0.9 and LOM 6-10.5). Later a correction multiplier for shale gas

formations (Ro>0.9/LOM >10.5) was proposed by Sondergeld (Sondergeld et al. 2010). Passey’s

method has also been incorporated into Pickett plots for comprehensive evaluation of TOC,

water saturation and flow region identification (Yu and Aguilera 2011) (Wu and Aguilera 2012).

2.4 Permeability

2.4.1 Permeability of shale gas, tight gas and conventional gas reservoirs

Permeability of unconventional gas (including tight gas and shale gas) reservoirs is 0.1

millidarcy or less. The permeability of shale gas reservoirs is usually much lower than tight gas,

32
and ranges from 10 nanodarcy to 1 microdarcy (Figure 2-8). When we incorporate pore throat

aperture rp35 together with permeability and porosity, a clear trend (Figure 2-9) has been observed

for all reservoirs (Aguilera 2010). The graph shows that reservoir data fall into well-defined

bands with certain ranges of pore throat apertures. Unconventional gas reservoirs have pore

throat radii bigger than about 0.55 m icrons. Shale gas reservoirs have pore throat radii smaller

than 100 nm.

Note in the log scale of permeability that there can be changes of about three orders of

magnitude in a single formation. Heterogeneity is just one of the aspects causing

permeability uncertainty in shale gas reservoirs. Permeability measured in laboratory does

not reflect the permeability after hydraulic fracture stimulations. The unknown geometry of

the stimulated fracture network makes it very difficult to input a permeability value for any

analytical model or numerical simulation.

Figure 2-8 Permeability (log scale) range of shale gas, tight gas and conventional gas (Russum 2009)

33
Figure 2-9 Permeability vs. Porosity crossplot including tight gas (Nikanassin) and shale gas data

including Horn River (HR), soft shales in Canada, Fayettville(F), Barnet (B), Huron and Marcellus

shales in the United States (Adapted from (Aguilera 2010)).

Fracture width and proppant packing can vary significantly from stage to stage in a multi-

stage hydraulic fracturing job. Geomechanical changes during the gas production also have

complex effects on t he whole stimulated reservoir volume. All these problems make the

permeability to be very difficult to estimate and model.

2.4.2 Permeability from pulse decay measurement

The pulse decay apparatus shown in Figure 2-10 can be used for determining permeability. It

consists of an upstream vessel that stores a gas volume V1, a sample holder capable of applying

high-confining stresses, which holds a cylindrical core sample, and a downstream vessel of

volume V2. A differential pressure transducer measures the pressure difference (Δp) across the

34
core and a pressure gauge measures the absolute pressure (p2) in vessel 2. T he pressure pulse

decay data (Δp) is measured as a function of time (Javadpour 2012b).

Pulse decay techniques enable measurement of very low permeabilities, something which is not

possible with traditional steady-flow permeability measurement. Both full core (Dicker and

Smits 1988) and crushed samples have been used for measuring permeability (Luffel, Hopkins,

and Shettler 1993). Mercury injection has also been investigated to measure shale permeability

(Kamath 1992).

Improvements have been developed recently to correct for the impact of adsorption (Cui, Bustin,

and Bustin 2009) by using natural gas (methane) instead of helium and nitrogen (Gao, Xu, and

Jiang 2013). A study has also shown that for crushed samples, the measured permeability is

strongly dependent on t he particle size and measurement procedure (Tinni, Fathi, and Agarwa

2012). Lack of standard protocols make measured permeability values questionable.

Laboratory measured permeability can be considered as matrix permeability km. Permeability

derived from production data analysis is an effective permeability ke which includes both

hydraulic fractures and stimulated natural fractures. Both values are important for analytical

models and numerical simulation. During shale gas production, as reservoir pressure decreases,

both km and ke are influenced by changes in the geomechanical environment. Permeability ke is

more sensitive to pressure changes than km.

35
Δp p2

A
V1 core V2

C B
N2

Figure 2-10 Apparatus for pulse decay permeability (Javadpour 2012b)

2.5 Water saturation

In shale gas reservoirs, it is widely believed that water only exists in the inorganic matrix (Alfred

and Vernik 2012, Glorioso and Rattia 2012, Ramirez et al. 2011). As shown in Figure 2-1, there

are generally two types of water, clay bound water and capillary bound water.

In the laboratory, water saturation Sw is measured by standard Dean Stark or by the retort

method. These methods extract fluid from crushed core samples to measure both water saturation

and dry porosity. Dean Stark combines thermal and toluene extraction while retort method uses

only thermal extraction. The water saturation data from Dean Stark method is generally higher

than that measured in retort method. Studies have been done to try to understand the difference

between these two methods (Handwerger, Willberg, and Pagels 2012). The author strongly

suggests caution against using any derived correlations/corrections for shales globally due to

differences in clay abundance, clay mineralogy and textural aspects.

Water saturations can also be derived from well logs. When advanced logs are available, such as

NMR and geochemical logs, water saturation estimation is more reliable. Models have also been

proposed to estimate water saturation from more conventional logs (Alfred and Vernik 2012)

36
(Wu and Aguilera 2012). However, these methods are not meant to replace detailed log

measurement and petrophysical models, but rather to provide quick estimations of water

saturation with limited input data particularly in those cases where available information is

scarce. The practical use of the simpler methods is still to be tested by more case studies.

Determination of water saturation from logs is discussed in detail in Chapter 4 of this thesis

jointly with the development of an original dual-porosity model for evaluation of shale

reservoirs.

2.6 Grain density

Grain density of shale is influenced by clay, heavy metal minerals and kerogen. It can be

obtained during the process of porosity and water saturation measurements. The measured grain

density suffers from similar uncertainties to the ones described above. Grain density can also be

estimated by correctly interpreting mineralogical analysis, such as X-ray diffraction (XRD) or X-

ray fluorescence (XRF), and by advanced geochemical logging tools. Care must be exercised

with density determined from XRD, as it is based on element analysis, which does not account

for kerogen in the rock.

2.7 Mineralogy

Mineralogy is essential for a complete quantitative evaluation of shale reservoir properties

(Glorioso and Rattia 2012). Accurate measurements of mineral composition provide a good base

for all other petrophysical parameters. XRD is the technique most often used in the laboratory.

Other methods are also available such as XRF, Fourier transform infra-red spectroscopy (FT-IR),

37
laser ionization breakdown spectroscopy (LIBS), elemental analyzer (EA), and scanning electron

microscope-energy dispersive spectroscopy (SEM-EDS). All of them are powerful but all of

them have their own limitations. In the field, geochemical logging tools can be used for

elemental analysis, from which mineral compositions can be derived.

2.8 Summary of shale petrophysics

Petrophysical parameters for shale are not limited to the ones we mentioned above. A more

complete list for shale gas petrophysical property evaluation including direct lab measurement

methods, logging tools and other methods is presented in Table 2-2.

Determination of geomechanical properties such as Young’s modulus, Poisson’s ratio and

brittleness is also very important. These parameters are key for establishing optimum locations of

hydraulic fracture initiation, and for estimating propagation and closure of hydraulic fractures.

High brittleness index is usually associated with relatively high Young’s modulus and low

Poisson’s ratio, which is favourable for hydraulic fracturing.

From the above summary and previous discussion in this chapter, it is possible to appreciate the

challenges and difficulties faced on petrophysical studies of shale reservoirs. Every shale rock is

unique and every shale reservoir is unique. It is a difficult task to establish standard protocols for

laboratory measurement procedures, log interpretation workflows, and correlations between core

and log data. Uncertainties will always be there. It is important to be cautious when adapting

methods/models from analogous shale reservoirs, especially if they are not in the same

38
geological area. More basic studies still need to be done to understand the internal shale rock

structure and the influence of shale rock structure on different measurement techniques.

Most laboratory techniques and logging tools have been used for a long time in conventional

reservoirs, and some of them have been successful in coal bed methane reservoirs. Studies

should be done to pinpoint the specific reasons that cause difficulties when applying these

conventional techniques in shale reservoirs and to provide solutions to improve the accuracy of

measurements and to minimize uncertainties.

Beyond all of the petrophysical parameters mentioned above, there is another major, unsolved

uncertainty: the hydraulic fracture network. Even high-density microseismic data and chemical

tracers are not able to provide a definitive description of how these fracture networks really look

like. For quantification purposes, it is important to know fracture length, fracture width, fracture

density, dominant fracture direction, network distribution and the change of natural fractures

stemming from the stimulation. Reducing uncertainty in the petrophysical parameters discussed

above allows a higher focus on the hydraulic fractures and to get more meaningful results from

reservoir simulation efforts. This is demonstrated in Chapter 5 with a simple conceptual model.

39
Table 2-2 Methods to measure shale gas petrophysical properties (modified from (Sondergeld et al.

2010)).

Reservoir Property Lab measurement (core, cuttings) Log and other methods

Porosity Gas expansion, MICP NMR, derived from log combination

Canister desorption & Langmuir


Free & Sorbed Gas
Adsorption

Log correlation (density, gamma ray, neutron…),


TOC RockEval & Leco combustion Passey's method (sonic and resistivity
combination)

Maturity Visual Ro, RockEval

pulse decay analysis, MICP (mercury


Permeability PDA, NMR
injection capillary pressure)

Water Saturation Dean Stark & Retort derived from log combination

XRD, TS point counts, FTIR, ICP-MS


Rock Composition (inductively coupled plasma mass Geochemical log
spectrometry), SEM-EDS

DSI (dipole shear sonic imager, Schlumberger)


Elastic Properties Core-based compression test (static)
Seismic data (P-wave and S-wave velocity)

Mud log,PDA(production data analysis),


Fluid Properties PVT
pressure gradients

Uncoventional well testing such as DFIT


(diagnostic fracure injection test, Fekete), PITA
Fracture and Closure Stress
(perf inflow test analysis, Fekete), Frac job,
DSI

IFOT (injection fall off test), PDA, log-based,


Pore Pressure
“dip-in”

OHL(openhole liner) & PL


Temperature frac job, & IFOT
(preperforated liner)

40
2.9 Alberta shale gas plays

Alberta is blessed with large resources of heavy oil, tight gas, coal bed methane, shale gas and

shale oil. Most of these unconventional resources were once considered not economically

recoverable. Alberta has come a long way to understand the geology of these resources and to

develop economic exploitation technologies. The natural environment protection and ecosystem

maintenance are always primary concern of Alberta. Now Alberta has become the energy focus

of the entire world. Alberta’s resources are attracting attention and investments from different

places around the globe. It is anticipated that tight gas and shale gas will become another major

contribution to Alberta energy resources comparable to oil sands.

A series of excellent ERCB’s open reports from 2008 t o 2013 have investigated the major

potential shale gas plays in Alberta including Colorado group, Banff/Exshaw, Duvernay,

Muskwa, Wilrich, Nordegg, Rierdon and Montney formations (Rokosh et al. 2008, Rokosh et al.

2012, Rokosh et al. 2009, Beaton et al. 2009a, Pawlowicz et al. 2009a, b, Beaton et al. 2010c,

Anderson et al. 2010b, Anderson et al. 2010a, Beaton et al. 2009c, Beaton et al. 2009b, Beaton et

al. 2010d, a, b, Rokosh et al. 2013).

Hundreds of core and outcrop samples have been systematically mapped, and evaluated from

laboratory measurements and well logs using methods mentioned in the previous section. Data

from these series of studies are valuable and will provide an initial guideline for future shale gas

exploration in Alberta. A summary of results is presented in Table 2-3.

41
The total hydrocarbon endowment from the formations mentioned above (except Colorado

group) is estimated at 3424 TCf natural gas, 59 billion barrels of natural-gas liquids, and 424

billion barrels of oil (Rokosh et al. 2012)(Table 2-3). These volumes were estimated based on all

measured data and geostatistics models. These studies are still ongoing, and data will be updated

once more accurate measurement techniques and models are available.

Table 2-3 Summary of estimates of Alberta shale- and siltstone-hosted hydrocarbon resource

endowment with these formations highlighted in generalized stratigraphic chart of Alberta

(Adapted from (Rokosh et al. 2012)).

A map of Alberta formations with depth to top and stratigraphic order is shown in Figure

2-11(Rokosh et al. 2012). Core sample sites are shown as red dots. Overlapping of these

formations are common, especially on t he west and northwest regions. Thus exploration is not

limited to a single formation. A systematic plan and careful design is needed to maximize

42
production from all these potential organic rich formations. Among them, Montney, Muskwa and

Duvenay are the top three formations with the largest endowment estimated at this time.

Figure 2-11 Map of Alberta shale gas formations (data from (Rokosh et al. 2012))

43
According to ERCB’s estimate, Montney is ranked first in all three hydrocarbon resources:

natural gas, natural gas liquid and oil. This is probably due to two main reasons: (1) the

formation thickness and (2) the high thermal maturity of organic material across the whole

Figure 2-12 Comparison between Montney and Banff/Exshaw showing gross isopatch

thickness, porosity thickness (φ×h), and thermal maturity (data from (Rokosh et al.

2012)).

44
formation. On the other hand, the Banff/Exshaw group shows the lowest hydrocarbon potential

in all three categories. That is due to the relatively small thickness of the formations and the low

maturity level of organic material (Figure 2-12).

Table 2-4 summarizes lithology description from the ERCB report series. These “shale”

formations are composed of many different lithologies. Commonly used lithology terms related

to shales are summarized in Table 2-5. Some of them do not necessarily fall into ERCB’s

definition of shale. TOC and porosity measurements on core samples from researched Alberta

shale formations are listed in Table 2-6 and Table 2-7.

Table 2-4 Alberta shale formation lithologies ((Rokosh et al. 2012)).

Table 2-5 Commonly used lithology terms related to shale

45
Shale permeability measurements are performed on a very limited number of samples. Reported

values range from micro-darcy to milli-darcy. Liquid permeability (kliquid) values are usually

smaller than air permability (kair). Permeability vertical to bedding planes (kv) is smaller than

permeability along bedding planes (kH). TOC, LOM, porosity, adsorbed gas content,

permeability and net pay provide a general idea of shale reservoir quality. Exploration and

exploitation of shale reservoirs also relies on ot her factors such as geomechanic properties of

rocks and natural fracture density.

The above summary of shale formations in Alberta demonstrate that basic petrophysical

properties are important for estimating hydrocarbon endowment, guiding future exploration and

even helping to pinpoint the sweet spots for reservoir development and well drilling. Minimizing

uncertainties in measurements of petrophysical properties modeling is crucial to obtain reliable

data for estimating endowment and simulation purposes.

Table 2-6 TOC of Alberta shale formations (data in this table are

estimated from histograms in (Rokosh et al. 2012))

46
Table 2-7 Porosity of Alberta shale resource (data adapted from (Rokosh et al. 2012)) .

47
2.10 References

Aguilera, R. 2010. Flow Units: From Conventional to Tight Gas to Shale Gas Reservoirs. Proc.,
SPE 132845, paper presented at the Trinidad and Tobago energy resources conference in Port of
Spain, Trinidad.

Alfred, D., L. Vernik. 2012. A New Petrophysical Model for Organic Shales. Proc., paper
presented at the SPWLA 53rd annual loggin symposium, Cartagena, Columbia.

Ambrose, R.J., M. Diaz-Campos, I. Yucel Akkutlu et al. 2010. New Pore-scale Considerations
for Shale Gas in Place Calculations. Proc., SPE 131772, paper presented at SPE Unconventional
Gas Conference held in Pittsburgh, Pennsylvanica, USA.

Anderson, S.D.A., A.P. Beaton, H. Berhane et al. 2010a. Mineralogy, Permeametry, Mercury
Porosimetry, Pycnometry and Scanning Electron Microscope Imaging of the Montney
Formation: Shale Gas Data Release (Reprint).

Anderson, S.D.A., A.P. Beaton, H. Berhane et al. 2010b. Mineralogy, Permeametry, Mercury
Porosimetry, Pycnometry and Scanning Electron Microscope Imaging of Duvernay and Muskwa
Formations: Shale Gas Data Release (Reprint).

Beaton, A.P., J.G. Pawlowicz, S.D.A. Anderson et al. 2010a. Organic Petrography of the
Duvernay and Muskwa Formations in Alberta: Shale Gas Data Release (Reprint).

Beaton, A.P., J.G. Pawlowicz, S.D.A. Anderson et al. 2010b. Organic Petrography of the
Montney Formation in Alberta: Shale Gas Data Release (Reprint).

Beaton, A.P., J.G. Pawlowicz, S.D.A. Anderson et al. 2010c. Rock Eval, Total Organic Carbon
and Adsorption Isotherms of the Duvernay and Muskwa Formations in Alberta: Shale Gas Data
Release (Reprint).

Beaton, A.P., J.G. Pawlowicz, S.D.A. Anderson et al. 2010d. Rock Eval, Total Organic Carbon
and Adsorption Isotherms of the Montney Formation in Alberta: Shale Gas Data Release
(Reprint).

Beaton, A.P., J.G. Pawlowicz, S.D.A. Anderson et al. 2009a. Geochemical and Sedimentological
Investigation of Banff and Exshaw Formations for Shale Gas Potential: Initial Results (Reprint).

Beaton, A.P., J.G. Pawlowicz, S.D.A. Anderson et al. 2009b. Rock-Eval, Total Organic Carbon,
Adsorption Isotherms and Organic Petrography of the Banff and Exshaw Formations: Shale Gas
Data Release (Reprint).

Beaton, A.P., J.G. Pawlowicz, S.D.A. Anderson et al. 2009c. Rock Eval, Total Organic Carbon,
Adsorption Isotherms and Organic Petrography of the Colorado Group: Shale Gas Data Release
(Reprint).

Crain , E.R. Crain's petrophysical handbook (in. http://www.spec2000.net/11-vshtoc.htm.

48
Cui, X., A.M.M. Bustin, R.M. Bustin. 2009. Measurements of gas permeability and diffusivity of
tight reservoir rocks: different approaches and their applications (in Geofluids 9: 208-223.

Dicker, A.I., R.M Smits. 1988. A practical approach for determining permeability from
laboratory pressure-pulse decay measurements. Proc., SPE 17578, SPE International Meeting in
Petroleum Engineering, Tianjin, China.

Gao, Cheng, Ruina Xu, Peixue Jiang. 2013. T he Shale-Gas Permeability Measurement
Considering the Rarefaction Effect on Transport Mechanism in the Nanopores. Proc., 16944-MS,
International Petroleum Technology Conference in Beijing, China.

Glorioso, Juan C., Aquiles Rattia. 2012. Unconventional Reservoirs: Basic Petrophysical
Concepts for Shale Gas. Proc., SPE 153004, paper presented at the SPE/EAGE European
unconventional resources conference in Vinenna, Austria.

Handwerger, D.A., D. Willberg, M. Pagels. 2012. Reconciling Retort versus Dean Stark
Measurements on Tight Shales. Proc., SPE 159976, paper presented at the SPE annual technical
conference and exhibition in San Antonio, Texas, USA.

Hood, A., C.C.M. Gutjahr, R.L. Heacock. 1975. Organic Metamorphisum and the Generation of
Petroleum (in The American Association of Petroleum Geologists Bulletin 59 (6): 986-996.

Javadpour, F. 2012. Lecture of Advances in Unconventional Shale Gas Resources

Kamath, J. 1992. Evaluation of accuracy of estimating air permeability from mercury-injection


data. (in Society of Petroleum Engineers Formation Evaluation 7: 304-310.

Kenyon, B, R. Kleinberg, C. Straley et al. 1995. Nuclear magnetic resonace imaging-technology


for the 21st century (in Oilfield Review 7 (3): 19-33.

Lewis, R.E., D. Ingraham, M. Pearcy et al. 2004. New Evaluation Techniques for Gas Shale
Reservoirs. Proc., Reservoir symposium, Schlumberger.

Luffel, D.L., C.W. Hopkins, P.D. Shettler. 1993. Matrix permeability measurements of gas
productive shales. Proc., SPE 26633, S PE Annual Technical Conference and Exhibition,
Houston, TX, USA.

Ortega, C.E. 2012. D rill Cuttings-based Methodology to Optimize Multi-stage Hydraulic


Fracturing in Horizontal Wells and Unconventional Gas Reservoirs. MSc, University of Calgary,
Calgary.

Passey, Q.R., W.L. Bohacs, W.L. Klimentidis. 2010. F rom Oil-Prone Source Rock to Gas-
Producing Shale Reservoir - Geologic and Petrophysical Characterization of Unconventional
Shale Gas Reservoirs. Proc., SPE 131350, pa per presented at International Oil and Gas
Conference and Exhibition, Beijing, China, 29.

49
Passey, Q.R., S. Creaney, J.B. Kulla. 1990. A practical model for organic richness from porosity
and resistivity logs (in AAPG Bulletin 74 (12): 1777-1794.

Pawlowicz, J.G., S.D.A. Anderson, C.D. Rokosh et al. 2009a. Mineralogy, Permeametry,
Mercury Porosimetry and Scanning Electron Microscope Imaging of the Banff and Exshaw
Formations: Shale Gas Data Release (Reprint).

Pawlowicz, J.G., S.D.A. Anderson, C.D. Rokosh et al. 2009b. Mineralogy, Permeametry,
Mercury Porosimetry and Scanning Electron Microscope Imaging of the Colorado Group: Shale
Gas Data Release (Reprint).

Quirein, J., J. Witkowsky, J. Truax et al. 2010. Integrating Core Data and Wireline Geochemical
Data for Formation Evaluation and Characterization of Shale Gas Reservoirs. Proc., SPE
134559, paper presented at the SPE annual technical conference and exhibition, Florence, Italy,
18p.

Quirein, John, Eric Murphy, Greg Praznik et al. 2012. A Comparison of Core and Well Log Data
to Evaluate Porosity, TOC, and Hydrocarbon Volume in the Eagle Fort Shale. Proc., SPE
159904, paper presented at the SPE annual techinical conference in San Antonio, Texas, USA.

Ramirez, T.R., J.D. Klein, R.J.M. Bonnie et al. 2011. Comparative Study Formation Evaluation
Methods for Unconventional Shale-Gas Reservoirs: Application to the Haynesville Shale
(Texas). Proc., SPE 144062, pa per presented at SPE North American Unconventional Gas
Conference in Woodlands, Texas, USA.

Rokosh, C.D, J.G. Pawlowicz, H. Berhane et al. 2009. Geochemical and Sedimentological
Investigation of the Colorado Group for Shale Gas Potential: Initial Results. Calgary (Reprint).

Rokosh, C.D. , J.G. Pawlowicz, H. Berhane et al. 2008. What is Shale Gas? An Introduction to
Shale-Gas Geology in Alberta, Vol. Open file report 2008-08 (Reprint).

Rokosh, C.D., C.S. Crocq, J.G. Pawlowicz et al. 2013. Rock Eval and Total Organic Carbon of
Sedimentary Rocks in Alberta (tabular data, tab-delimited format) (Reprint).

Rokosh, C.D., S. Lyster, S.D.A. Anderson et al. 2012. Summary of Alberta's Shale- and
Siltstone-Hosted Hydrocarbons (Reprint).

Russum, D. 2009. Evaluating Unconventional Gas, AJM.

Shtepani, E., L.A. Noll, L.W. Elrod et al. 2010. A New Regression-Based Method for Accurate
Measurement of Coal and Shale Gas Content (in Reservoir Evaluation & Engineering: 359-364.

Sigal, R.F., E. Odusina. 2011. Laboratory NMR Measurements on M ethane Saturated Barnett
Shale Samples (in Petrophysics 52 (1): 32-49.

50
Sondergeld, C.H., K.E. Newsham, J.T. Comisky et al. 2010. P etrophysical Consideration in
Evaluating and Producing Shale Gas Resources. Proc., SPE 131768, paper presetned at the SPE
unconvetional gas conference, Pittsburgh, PA, 34.

Sondergeld, Carl H., Ray J. Ambrose, Chandra S. Rai et al. 2010. Micro-structural studies of gas
shales. Proc., SPE 131771, paper presented at the unconventional gas conference in Pittsburgh,
PA, USA.

Suggate, R.P. 1959. N ew Zealand coals, their geological setting and its influence on their
properties, New Zealand Dept. Sci. Indsutry Researcyh.

Tinni, Ali, Ebrahim Fathi, Rajiv Agarwa. 2012. Shale Permeability Measurements on Plugs and
Crushed Samples. Proc., SPE 162235, SPE Canadian Unconventional Resources Conference in
Calgary, Alberta, Canada.

Wu, P, R. Aguilera. 2012. Investigation of Shale Gas at Nanoscale Using Scan Electron
Microscopy, Transmission Electron Microscopy and Atomic Force Microscopy, and Up-scaling
to a Petrophysical Model for Water Saturation Evaluation in Shales. Proc., SPE 159887, SPE
annual technical conference in San Antonio, TX, USA.

Yu, G., R. Aguilera. 2011. Use of Pickett Plots for Evaluation of Shale Gas Formations. Proc.,
SPE 146948, paper presented at the SPE annual technical conference and exhibition in Denver,
CO, USA.

51
Chapter Three: Investigation of Shale at Nanoscale Using Scan Electron Microscopy,

Transmission Electron Microscopy and Atomic Force Microscopy

Due to quick development in horizontal drilling and fracturing technologies, shale formations

containing gas, oil and natural gas liquids, formerly considered very difficult if not impossible to

recover, have become one of the hottest energy topics. Until now, only a limited number of

studies have been carried out aiming at the investigation of shale structure at nanometer scale

due to limited access to specialized instruments such as scanning electron microscopy (SEM),

transmission electron microscopy (TEM) and atomic force microscopy (AFM).

SEM, TEM and AFM are all capable of revealing nanoscale structures in shale samples. Images

of shale formations in the Western Canada Sedimentary Basin Canada obtained at the

microscopy and imaging facility (MIF) in University of Calgary are compared with images from

other shale formations reported in the literature. As powerful as these nanoscale-capable

microscopes are, they all have limitations due to the high instrument cost, limited access and

time consuming imaging and sample preparation process. This study aimed at a quick

comparison of all these instruments using simple and direct sample preparation techniques,

providing suggestions and advice on how these instruments can be used for petroleum and

geology purposes from a practical point of view.

3.1 Introduction of instruments of nanometer resolution

Fast developing technologies for horizontal well drilling and hydraulic fracturing opened the

door for exploitation of gigantic unconventional petroleum resources in North America such as

52
those found in the Western Canadian sedimentary basin (WCSB), and the Utica, Marcellus,

Bakken and Barnett shales. How to develop these resources responsibly with the lowest possible

cost and the highest efficiency has become one of the major challenges of the oil and gas

industry. Fundamental understanding of shale rocks is essential for stimulation and completion

optimization, cost control and future exploitation of these resources. As this study concentrates

on the structure of shales and determination of water saturation the method has direct application

in the case of all types of shale petroleum reservoirs.

Although advanced drilling, stimulation and completion technologies make these unconventional

resources accessible, our basic understanding of shale structure, flow paths, porosity, water

saturation and permeability are still very limited. Recent introduction of powerful electron

microscopes such as SEM, TEM and AFM provide the capability to observe shale rock at

nanometer scale. Such insights from these nano-scale capable instruments will help us to better

understand shales at a fundamental level.

The resolution (d) of a conventional optical microscope is dependent on the wavelength of light

(λ), refractive materials for the objective lens and the numerical aperture (NA).

𝝀
𝒅 = 𝟐𝑵𝑨 3-1

The highest practical NA is 0.95 w ith air as the external medium, and 1.5 with oil. If we use

wavelength of green light (550 nm), the highest resolution value is around ~200 nm. Both

scanning electron microscopy (SEM) and transmission electron microscopy (TEM) are capable

of imaging at much higher resolution than optical microscope due to smaller de Broglie

53
wavelength of electrons compared to that of visible light. In TEM, an ultra-thin sample is

irradiated with a beam of uniform electrons. Images are formed from the interaction of electrons

transmitted through the sample.

SEM images a sample by scanning it w ith a high-energy beam of electrons in a r aster scan

pattern. The electrons interact with atoms of the sample producing signals of surface topography,

composition, electrical conductivity and morphology. The resolution of TEM can go down to 0.1

to 0.2 nm, which is about an order of magnitude higher than that of SEM. However, because the

SEM image relies on surface processes rather than transmission, it is able to image bulk samples

up to many centimeters in size whereas TEM requires sample thickness in the order of 100 nm.

Both TEM and SEM require the sample to be stabled in a vacuum chamber, to minimize

interaction of electrons with air molecules and improve image quality. Usually higher resolution

requires higher energy for electron beam and higher vacuum (Egerton 2005). Atomic force

microscope (AFM) consists of a cantilever with a tip (probe) at its end that is used to scan the

specimen surface. T he cantilever is typically made of silicon or silicon nitride with tip radius

from several nanometers to hundreds of nanometers. Depending on the situation, forces that are

measured in AFM include many different kinds including mechanical contact force, van der

Waals forces, electrostatic forces and chemical bonding forces (Braet, De Zanger, and Wisse

1997).

High resolution AFM can provide resolution comparable to SEM and TEM. AFM can work

perfectly well in ambient air or even a liquid environment, which makes AFM less expensive as

54
it does not need an ultra-high vacuum component in instrumentation (Binning, Quate, and

Gerber 1986) (Javadpour 2012a).

It is well known that nanometer scale pores exist in shale. These nanopores have a big impact on

permeability, porosity, pore pressure and fluid flow characteristics. Due to limited resolution

(micron) of conventional electron microscopes, the information we can get for pores and the

internal structure in shale is minimal. SEM and TEM opened the possibilities to look at these

nano-sized pores and structures. So far only a limited number of studies have been carried out to

investigate the structure and pore system in shale at nanometer scale using SEM and TEM

(Curtis et al. 2011, Loucks et al. 2009, Wang et al. 2009, Sondergeld et al. 2010). Images of

nanopores in shale using AFM were first presented by Javadpour (Javadpour 2009). The study

showed both the potential and the challenges of using AFM to study detailed nano-scale

structures in shale.

Figure 3-1 Image of TEM and SEM (Medwow)

55
Figure 3-2 Image of an AFM instrument and schematic assembly (Kumar, Dao, and Mohanty 2008)

Until now most of these imaging analyses have not been used to quantify porosity, kerogen

content and pore distribution. Even though quantitative information can be estimated from a

specific imaging area, it is not convincing to conclude that these extremely small two

dimensional areas (µm × µm) can represent the whole core sample. Considering the number of

core/cutting samples, the number of images to be taken for a single core/cutting piece, and the

time consuming imaging and sample preparation process; generation of statistically significant

data seems to be an exhausting and expensive task. Rather than being a quantitative tool at macro

scale, these instruments provide valuable information at micro and nano-scale. High

magnification images clearly showed the existence of nanopores in kerogen and their distribution

(Curtis et al. 2011, Javadpour 2009, Loucks et al. 2009, Sondergeld et al. 2010).

Element composition can also be obtained from energy-dispersive X-ray spectroscopy (EDS)

and wavelength dispersive X-ray spectroscopy in SEM and STEM, giving us qualitative or semi-

56
quantitative mineralogy compositions. Many different pore types could also be easily

differentiated under SEM and TEM (Slatt and O'Brien 2011). These instruments provide the

most intuitive way to understand shales including composition, structure, kerogen distribution,

heterogeneity of different components (i.e. matrix, kerogen and fracture), and structure and

connectivity of different pore types.

Systematic investigation will help us to better understand the relationship between nano-

structures and the parameters associated with fluid flow such as permeability, porosity, pore

pressure, and storage. With fundamental understanding of the rock, we can then develop more

accurate methods for measuring important parameters for shale petroleum evaluation, and

establish better models encompassing nanostructures to bigger scale properties. Appropriate

changes should be made for individual reservoirs with different shale and pore structures.

In practice, models need to be established that require a minimum amount of input data, if

possible. In horizontal wells, in most cases, data are scarce and might include only a few

standard logs (sometimes only a gamma ray log) and drill cuttings (if drilled with tri-cone bits).

Core is usually not available, but high quality cuttings could be used for lab measurements to

determine for example porosity and permeability (Ortega 2012), for NMR analysis, and SEM,

TEM, AFM imaging.

3.2 SEM, TEM and AFM of Muskwa and Nordegg shale samples

In this study, two core samples from the Western Canada Sedimentary Basin (WCSB) were

imaged by SEM, TEM and AFM. The Muskwa is a sub-unit of the Horn River formation while

57
Nordegg is a sub-unit of the Fernie formation. Nordegg shale has relatively high composition of

phosphates and limestone. FEI XL30 SEM, Technai F20 TEM and Nanowizard II AFM

instruments were used for imaging these shale samples. First, samples were imaged under SEM

as received with no further preparation. Then all samples were highly polished for SEM and

AFM imaging.

For SEM experiments, a thin layer of gold coating was deposited onto the sample surface, and

conductive carbon was applied to both sides of the sample to make them electric conductive. In

AFM experiment, imaging was performed with a Tap300Al probe in intermittent contact in air.

The resonant frequency for the probes is 300 k Hz with a force constant of 40N/m. In TEM

experiments, particles (similar to powder) from core samples were scrapped off using a thin

blade and deposited onto TEM grids from diluted ethanol solution.

3.2.1 SEM

Figure 3-3 shows images of the Muskwa with floccules that are similar to those commonly

present in several other shales such as Permian Clear Fork formation, Barnett and Woodford

shales (Slatt and O'Brien 2011). Pores (indicated by arrows) in (a) and (c) are of similar sizes

(<1µm to 2µm) as those in Barnett and Woodford gas shales reported by (Slatt and O'Brien

2011). However, we are not sure if these pores are real or grain pull-out due to break during core

cutting process. A 3~4 µm micro-fracture is shown clearly in (c). Wherever micro-fractures exist,

they will provide important pathways for fluid flow. At 100,000 magnification, we can see the

presence of some nano-fractures in Figure 3-4. These fractures are not very clear but they are still

58
visible. Their presence in these images show they would have width bigger than 0.38 nm

(diameter of the methane molecule) and would contribute to gas flow but at very low speed. All

images in Figure 3-3 and Figure 3-4 were taken using unpolished samples as received.

Figure 3-3 SEM images of Muskwa core sample. Samples scanned as received with no preparation.

(a) Typical flocculated clay microfabric. (b) Open microfracture (3~4 um) (c) Typical clay

microfabric in micron holes.

59
Figure 3-4 SEM image of shale sample (as received) showing nano-fracture

Figure 3-5 SEM BSE image of polished Nordegg sample (EDS scan results of 3 specified area shown

on the left table)

60
A back-scattered electron (BSE) image with EDS scan in a chosen area of highly polished

Nordegg core sample is shown in Figure 3-5 with atomic % of four major elements listed out for

relative comparison. BSE signals are sensitive to the atomic number of the sample, so they can

provide information about the distribution of lithology and organic materials. In this image, the

brightest area is pyrite framboids, dark grey area represents the organic material, and light grey

area represents the inorganic matrix (rock/clay). This is confirmed by energy dispersive

spectroscopy.

As shown on t he left table in Figure 3-5, pyrite has relative high percentage of iron; organic

material is dominated by carbon, and inorganic matrix area has high percentage of calcium.

Figure 3-6 is another BSE image of highly polished Nordegg sample. In this image, the sub-

micron pores appear to be dark, the organic material appears grey, and the inorganic matrix

appears bright.

Using appropriate thresholds, we were able to quantify the TOC (~ 9%) and porosity (~7%) of

this specific image area. These estimates provide a link for upscaling purposes and can be

correlated with other lab and log measurements. While both BSE image mode and EDS scan

provide quantization possibilities for porosity, TOC and local mineralogy distribution, the result

might not be consistent with lab measurements and geochemical analysis. The discrepancy is due

to several reasons. As we mentioned earlier, it is very difficult to obtain the quantification data

with statistical significance. Additionally, the image analysis is 2-dimensional, while traditional

lab measurements are based on 3-dimensional bulk volume. Even though the EDS scan can

penetrate the sample down to several microns, the data are still considered as 2-dimensional.

61
Despite absolute quantification difficulties, SEM image analysis can be very useful for relative

comparison of shale core samples from different wells, reservoirs and formations. Semi-

quantitative analysis could be carried out using samples with reasonable batch size. A high

magnification SEM image of highly polished Muskwa sample clearly showing the detailed

structure of organic material is shown in Figure 3-7. This image shows three different regions

with different textures. In the middle, the organic material has high porosity, soft texture and

sponge-look like structure. On the left side, the inorganic matrix has low porosity, hard and

smooth texture. On the right side, the inorganic matrix is covered with a layer of organic

material.

Figure 3-6 SEM BSE image of Nordegg core sample

62
Figure 3-7 SEM high magnification image of Muskwa sample showing organic material structure

3.2.2 TEM

The preparation technique used for TEM experiment might cause components (i.e. clay lamella,

kerogen) to be damaged or separated. But it has proved to be effective in revealing different

nanostructures in core samples in our study (Figure 3-8 to Figure 3-10). Annular dark field (ADF)

TEM images of Muskwa and Nordegg core samples (Figure 3-8, Figure 3-9) reveal the existence

of nanopores in kerogen. ADF is sensitive to crystallinity in the sample. As a result, images in

which crystalline materials are present appear bright, amorphous materials appear dark, and

vacant spaces are black. This imaging mode shows the nanopore morphology more clearly than

bright field imaging. The sponge like internal structure in organic material and nanopore size

63
range are similar to those reported in a previous study by Curtis (Curtis et al. 2011). The huge

surface area provided by these nanopore networks contribute to significant adsorbed gas volume.

Some other interesting nano-features are shown in Figure 3-10. Figure 3-10 (a) shows kerogen

together with some inorganic crystals. From the shape and morphology, we conclude they are

salt crystals and pyrite framboids. The dimension of pyrite framboids is smaller compared to

previous studies (Slatt and O'Brien 2011). Figure 3-10 (b) shows a sharper image of nano salt

crystals. In Figure 3-10 (c), a zoomed-in picture of clay lamella is shown with fine texture. Sub-

nano meter porosities are visible in the clay matrix.

With advanced FIB/STEM (focused ion beam/scanning transmission electron microscope) dual

beam instrument, the image resolution can be further improved (Curtis et al. 2011, Sondergeld et

al. 2010). Even 3-dimensional image analysis becomes possible using sequential etching process

(Sondergeld et al. 2010). But the imaging process is extremely time-consuming and the access to

these advanced instrumentations is very limited. Generation of statistically significant data is

impractical.

64
Figure 3-8 Annular dark field (ADF) TEM image of kerogen in Nordegg core sample

65
Figure 3-9 Annular dark field (ADF) TEM image of kerogen in Muskwa core sample

66
Figure 3-10 Annular dark field (ADF) TEM image of nano structures in Muskwa and Nordegg

core samples. (a) Aggregates with shape and morphology like pyrite framboids, salt crystals and

kerogen in Nordegg sample. (b) Nano crystals in Nordegg sample. (c) Bright field image of clay

flake with sub-nano porosity in Muskwa sample.

67
3.2.3 AFM

Figure 3-11 shows the height and phase images of Muskwa sample. These two images correlate

with each other. The smooth area in phase image corresponds to bright area in height image,

representing hard inorganic material with higher topography. On the other hand, the rough area

in phase image corresponds to the dark area in height image, representing soft organic material

with lower topography. These results are consistent with observation from SEM imaging. During

polishing process, softer organic material was easier to remove. So the area with organic material

usually appears as concave. Highly polished shale samples are still too rough for high resolution

AFM experiment. Sample preparation using focused ion beam is a must to produce AFM images

revealing detailed nanopore network structures. AFM scanning experiment takes much longer

time than SEM and TEM imaging in our case.

Figure 3-11 AFM image of polished Muskwa sample

68
3.3 Summary

SEM, TEM and AFM provide intuitive observations and imaging tools for visualizing the shale

structure at nano-scale level. Although with limitations in reliable quantification of macro

characteristics, these powerful microscopes can provide very valuable insight into fundamental

understanding of shale rocks, which is essential for development of robust models for

petrophysical and reservoir engineering evaluations.

In our study, we use a facility available in the Faculty of Medicine at the University of Calgary

to study shale samples from WCSB. We provide direct evidence on t he existence of nanopore

networks in Nordegg and Muskwa shale samples using all three tested instruments (SEM, TEM

and AFM). It is impractical to obtain quantitative information considering imaging time and cost.

However, with well log and core data available for specific imaging samples, these high

resolution images provide visual guidance and basic structure understanding for establishing

reasonable petrophysics and fluid flow models.

TEM requires samples with less than 200 nm, so only scrapped particles from core samples were

used for this test. Obviously these particles do not show bigger scale structures. AFM imaging

time is much longer than that of SEM and TEM. Highly polished samples are not smooth enough

for producing high quality images. Plasma etching with dual beam integration is highly desirable

for TEM and AFM to obtain high quality images while maintaining the original structure of the

sample. However, this sample preparation process is very time consuming and expensive.

Currently, there are only couple facilities have this capability in Alberta.

69
Among the instruments we have tested, SEM is the best one we found for practical and routine

use by petroleum engineers and geologists. We were able to get SEM images with nanopore

network structure details at different scale using routinely highly polished samples. Under

favourable conditions quantification of porosity, TOC, mineralogy distribution is possible.

3.4 References

https://sjhsrc.wikispaces.com/AFM

Binning, G., C.F. Quate, Ch. Gerber. 1986. Atomic force microscope (in Phys. Rev. Lett. 56:
930-933.

Braet, F., R. De Zanger, E. Wisse. 1997. Drying cells for SEM, AFM and TEM by
hexamethyldisilazane: a study on hepatic endothelial cells (in Journal of Microscopy 186 (1): 84-
87.

Curtis, Mark E., Ray J. Ambrose, Carl H. Sondergeld et al. 2011. Transmission and Scanning
Electron Microscopy Investigation of Pore Connectivity of Gas Shales on the Nanoscale. Proc.,
SPE 144391, paper presented at the SPE North American Unconventional Gas Conference,
Woodlands, Texas, USA.

Egerton, Ray F. 2005. Physical principles of electron microscopy: an introduction to TEM, SEM
and AEM. New York, Springer Science+Business Media, Inc. (Reprint).

Javadpour, F. 2009. Nanopores and apparent permeability of gas flow in mudrock (shales and
siltstone) (in Journal of Canadian Petroleum Technology 48 (8): 16-21.

Javadpour, F. 2012. Atomic-Force Microscopy: A New Tool for Gas-Shale Characterization (in
Journal of Canadian Petroleum Technology 51 (4): 236-243.

Kumar, K., E.K. Dao, K. K. Mohanty. 2008. Atomic force microscopy study of wettability
alteration by surfactants (in SPE Journal 13 (2): 137-145.

Loucks, Robert G., Robert M. Reed, Stephen C. Ruppel et al. 2009. Morphology, genesis and
distribution of nanometer-scale pores in siliceous mudstone of the mississippian barnett shale (in
Journal of Sedimentary Research 79: 848-861.

Medwow. http://www.medwow.com/med/scanning-electron-microscope/hitachi/field-emission-
sem/33693.model-spec (in.

70
Ortega, C.E. 2012. Drill Cuttings-based Methodology to Optimize Multi-stage Hydraulic
Fracturing in Horizontal Wells and Unconventional Gas Reservoirs. MSc, University of Calgary,
Calgary.

Slatt, R.M., N.R. O'Brien. 2011. Pore types in the Barnett and Woodford gas shales:
Contribution to understanding gas storage and migration pathways in fine-grained rocks (in
AAPG Bulletin 95 (2): 2017-2030.

Sondergeld, Carl H., Ray J. Ambrose, Chandra S. Rai et al. 2010. Micro-structural studies of gas
shales. Proc., SPE 131771, paper presented at the unconventional gas conference in Pittsburgh,
PA, USA.

Wang, F.P., Robert M. Reed, John A. Jackson et al. 2009. Pore networks and fluid flow in gas
shales. Proc., SPE 124253, Paper prestend at annual technical conference and exhibition in New
Orleans, LA, USA.

71
Chapter Four: Petrophysical Model for Water Saturation Evaluation
in Tight and Shale Reservoirs

Water saturation is among the most basic petrophysical parameters considered in petroleum

reservoir engineering. In shale reservoirs, water saturations cannot be derived from standard well

logs using Archie’s equation. The model proposed in this thesis is inspired by a dual porosity

model developed by Aguilera (Aguilera and Aguilera 2003). In this thesis organic material has

been treated similarly to non-connected vugs in the model derivation. Triple or quadruple

porosity models could be built upon this model if different types of pores constitute a significant

portion of the total porosity. But more input data would be required for more complicated

models.

A new method has also been developed to display water saturation curves in Pickett plots using

the model developed in this thesis for shale formations. This method can also use for tight

reservoir using Aguilera’s dual porosity model. This plotting method is proven useful to estimate

average water saturation or an overall water saturation range in case studies of shale and tight

reservoirs. More accurate water saturation results were obtained with this plotting method than

with traditional Pickett plots.

4.1 Introduction

4.1.1 Archie’s equation and Pickett plot

As the cornerstone of petrophysics, Archie’s equation has been used since the 1940s. Archie’s

equation relates water saturation, Swn, to brine resistivity, Rw, and measured true resistivity, Rt.

(Archie 1942 ).

72
𝑭𝑹𝒘
𝑺𝒘 𝒏 = 4-1
𝑹𝒕

𝑹𝟎 𝒂
𝑭= �𝑹 = 𝝓𝒎 4-2
𝒘

Combining above two equations we obtain,

𝒂𝑹
𝑺𝒘 𝒏 = 𝝓𝒎 𝑹𝒘 4-3
𝒕

where,

n: Saturation Exponent
m: Cementation Factor
φ: Porosity-fractional
F: Formation Resistivity Factor
Rt : Measured True Resistivity
Rw: Brine resistivity
R0 : Resistivity of formation rock 100% saturated with brine of resistivity Rw
a: Tortuosity factor
Sw: Water Saturation

Water saturation can be estimated using Archie’s equation as shown above. Porosity φ and

resistivity Rt can be measured or calculated from log data. Parameters a, m and n are usually

assumed empirically, or obtained from core data if it is available; a=1, m=2, n= 2 are typical

assumed values for clean sand formations and they give reasonable approximations of water

saturation in many cases. However, m and n are not always close to 2 in all formations.

Consequently, without any geological and petrophysical knowledge of the formation, using

Archie’s equation or Pickett plot by assuming typical used values of a, m and n may produce

very misleading estimations of water saturation and relevant formation characteristics.

73
4.1.2 Pickett plot

Pickett plot was introduced back in 1973 to graphically solve Archie’s equation. By plotting

porosity and resistivity in log-log coordinates, we get straight lines at constant values of water

saturation. Not only this gives a more intuitive presentation of correlations between porosity,

resistivity and water saturation, but it also provides a quick estimate of formation quality via

pattern recognition, and even provides estimates of cementation factor and saturation exponent.

Figure 4-1 Cementation factor in Pickett plot (adapted from (Davis))

As shown in a sample Pickett plot in Figure 4-1, the cementation factor m (the reciprocal of the

slope) significantly affects water saturation. In core analysis, m is determined by a log-log plot of

measured formation resistivity factor F (which equals R0/Rw) versus φ. R0 is the resistivity of the

74
core 100% saturated with brine of resistivity Rw. The three scheme images at the top of Figure

4-1 show formations with smooth, well sorted grains (m<2), average grains (m= 2) and

irregular/cemented grains (m>2). Provided that these formations have the same porosity, Davis

(Davis) explained that smoother grains provide easier pathway for electrons than irregular

shaped grains, thus having smaller m. The value of m is dependent upon how easily electrons can

flow through the interstitial formation water around rock grains with high resistivity (red arrow

Figure 4-2 Importance of saturation exponent (n) in Pickett plot. Green lines represent

hydrocarbons and blue means 100% water for the three schemes shown above the Pickett plot

(upper three schemes were adapted from (Davis))

75
in Figure 4-1 shows the conductive pathway). If we consider formations with fractures and non-

connected vugs, fractures provide very good p athway for electric current whereas there is no

pathway through non-connected vugs. Formations with fractures have m values smaller than 2,

and formations with non-connected vugs have m values bigger than 2.

The saturation exponent controls the distance (spread) between parallel water saturation lines in

the Pickett plot. The bigger the n value, the wider is the spread (Figure 4-2). As shown in upper

three schematics in Figure 4-2, given the same porosity and water saturation, if the formation has

some micro-porosity short circuit, for example dispersed clay/shale as ion-selective membrane,

electric current is easier to pass through the formation, giving smaller n values. On the contrary,

oil wet rocks lose conductive pathways, causing bigger n values. Saturation exponent can also be

measured in cores. Instead of a single measurement on a single core plug sample, ideally a series

of measurements are taken on each core plug sample.

In the Pickett plot, m is assumed to be constant to plot straight water saturation lines. This is only

true for regular and homogenous distributed porosity (a clean sandstone formation for example).

In formations with complex structures such as natural fractures, non-connected vugs, shaly

sands, or shale reservoirs, m will not be constant any more. As a result, using straight water

saturation lines assuming constant m will not give reasonable water saturation estimates in these

complicated formations.

76
4.1.3 Dual porosity model by Aguilera

The model proposed in this study is inspired by a dual porosity model for formations containing

natural fractures and non-connected vugs (Aguilera and Aguilera 2003). The dual porosity model

for formations containing natural fractures and matrix porosity is given by Equation 4-7; the dual

porosity model for non-connected vugs and matrix porosity is given by Equation 4-8 (Aguilera

and Aguilera 2003). It is important to use proper scaling when we dealing with matrix porosity.

The matrix system does not include natural fracture or non-connected vugs; φb is the matrix

porosity attached to the bulk volume of the matrix (Equation 4-4), whereas φm is the matrix

porosity attached to the whole composite system (Equation 4-5). The partitioning coefficient (ν)

and vug porosity ratio (νnc) are given by Equations 4-11 and 4-12, respectively. The magnitude

difference between φm and φb is very small in most cases due to small volume percentage of

natural fracture/non-connected vugs. However, these small differences will transfer to larger

differences in derived water saturations and cause unacceptable errors in some cases.

𝒗𝒐𝒊𝒅 𝒗𝒐𝒍𝒖𝒎𝒆 𝒘𝒊𝒕𝒉𝒊𝒏 𝒎𝒂𝒕𝒓𝒊𝒙


𝝓𝒃 = 𝒎𝒂𝒕𝒓𝒊𝒙 𝒗𝒐𝒍𝒖𝒎𝒆
4-4

𝒗𝒐𝒊𝒅 𝒗𝒐𝒍𝒖𝒎𝒆 𝒘𝒊𝒕𝒉𝒊𝒏 𝒎𝒂𝒕𝒓𝒊𝒙


𝝓𝒎 = 𝒕𝒐𝒕𝒂𝒍 𝒗𝒐𝒍𝒖𝒎𝒆
4-5

𝒕𝒐𝒕𝒂𝒍 𝒗𝒐𝒍𝒖𝒎𝒆 = 𝒎𝒂𝒕𝒓𝒊𝒙 𝒗𝒐𝒍𝒖𝒎𝒆 + 𝒗𝒐𝒍𝒖𝒎𝒆 𝒐𝒇 𝒇𝒓𝒂𝒄𝒕𝒖𝒓𝒆 𝒐𝒓 𝒏𝒐𝒏 − 𝒄𝒐𝒏𝒏𝒆𝒄𝒕𝒆𝒅 𝒗𝒖𝒈𝒔 4-6

𝒏𝒂𝒕𝒖𝒓𝒂𝒍 𝒇𝒓𝒂𝒄𝒕𝒖𝒓𝒆 𝒗𝒐𝒍𝒖𝒎𝒆


𝝓𝟐 = 𝒕𝒐𝒕𝒂𝒍 𝒗𝒐𝒍𝒖𝒎𝒆
4-7

𝒏𝒐𝒏−𝒄𝒐𝒏𝒏𝒆𝒄𝒕𝒆𝒅 𝒗𝒐𝒍𝒖𝒎𝒆
𝝓𝒏𝒄 = 𝒕𝒐𝒕𝒂𝒍 𝒗𝒐𝒍𝒖𝒎𝒆
4-8

For matrix and fractures:

𝐥𝐨𝐠�𝝂𝝓+(𝟏−𝝂𝝓)𝝓𝒃 −𝒎𝒃 �
𝒎= 4-9
𝐥𝐨𝐠 𝝓

77
For matrix and non-connected vugs:

𝐥𝐨𝐠�𝝊𝒏𝒄 𝝓+(𝟏−𝝊𝒏𝒄 𝝓)𝝓𝒃 −𝒎𝒃 �


𝒎= 4-10
− 𝐥𝐨𝐠 𝝓

𝝓𝟐 𝝓−𝝓 𝝓−𝝓
𝒗= �𝝓 = 𝝓 𝒎 = 𝝓(𝟏−𝝓𝒃 ) 4-11
𝒃

𝝓𝒏𝒄 𝝓−𝝓𝒎 𝝓−𝝓𝒃


𝒗𝒏𝒄 = �𝝓 = = 4-12
𝝓 𝝓(𝟏−𝝓𝒃 )

Where:
φ: Total porosity-fractional
φb: Matrix block porosity attached to the bulk volume of the matrix system
φm: Matrix block porosity attached to the whole composite system
φnc: Porosity of non-connected vugs attached to the whole composite system
φ : Porosity of natural fractures or connected vugs associated with the whole composite system
2

m: Cementation factor of the whole composite system


mb: Cementation factor of the matrix
ν: Partitioning coefficient (fracture porosity ratio)
νnc: Vug porosity ratio

As mentioned in the previous section, natural fractures and non-connected vugs will have very

different and significant influence on t he electron conductivity of the formation. Aguilera took

these influences into consideration and developed equations for the cementation factor m of the

formation, which is crucial for water saturation evaluation. The cementation factor mb of the

matrix (not including fractures or non-connected vugs), is usually assumed to be constant for a

given interval.

78
4.2 Water saturations in Pickett plot using dual porosity model

To plot water saturation lines in Pickett Plot, we assume m to be constant for the formation,

giving us straight lines for constant water saturation values. However, from Aguilera’s equations

4-9 and 4-10, we know that m is dependent upon other parameters such as φ2 (fracture porosity),

φnc (non-connected vug porosity) and φb (matrix porosity). It is unlikely that m will be constant

throughout the interval being evaluated. As m is an exponential component in Archie’s equation,

even a small change can cause significant difference in water saturation values. Considering that

the variability of matrix porosity φb in a given interval is usually smaller than other parameters,

φb is assumed to be constant in this work to plot a series of water saturation lines (from 100 to

3%).

When we have very different facies and particle size distributions in the zones of interest, water

saturation lines can be plotted using different values of φb. Matrix porosity φb for different facies

can be estimated from sonic logs or obtained from core analysis from other wells in the reservoir.

Water saturation lines plotted using this method are not straight lines because m is not constant.

Furthermore, because fracture and vug porosity are usually a small portion of total porosity, from

porosity well logs, we can have an approximate idea about what φb values (usually a r ange) to

use for estimation of water saturation.

In this plotting method, m for dual porosity system containing natural fractures and matrix

porosity is calculated using Equation 4-13, which is simply insertion of Equation 4-11 into

Equation 4-9.

79
𝝓−𝝓𝒃 𝝓−𝝓𝒃
𝐥𝐨𝐠� +�𝟏− �𝝓𝒃 𝒎𝒃 �
𝟏−𝝓𝒃 𝟏−𝝓𝒃
𝒎= 𝐥𝐨𝐠 𝝓
4-13

m for dual prosity system containing non-connected vugs and matrix porosity is calculated using

equation 4-14, which is simply insertion of Equation 4-12 into Equation 4-10.
𝝓−𝝓𝒃 𝝓−𝝓𝒃
𝐥𝐨𝐠� +�𝟏− �𝝓𝒃 −𝒎𝒃 �
𝟏−𝝓𝒃 𝟏−𝝓𝒃
𝒎= −𝐥𝐨𝐠 𝝓
4-14

Water saturations are calculated using Archie’s equation with m calculated from the above

equations.

A case study from a gas well in the Wapiti pool of Alberta is used to demonstrate this new

plotting method. The production zone corresponds to the Falher formation. Figure 4-3 compares

water saturation lines plotted using Archie’s traditional single porosity model and Aguilera’s

dual porosity model for a system with natural fractures. Sonic log porosity is used to estimate φb.

φb has a mean of 0.04 with a coefficient of variation of 0.207, w hereas fracture porosity φ2

calculated from logs has a mean of 0.016 and a coefficient of variation of 0.514. S maller

coefficient of variation corresponds to relatively smaller variation.

Assuming fixed φb for plotting water saturation in a Pickett plot in unconventional low porosity

reservoirs is more reasonable than assuming fixed ν or φ2 in this type of application. From the

plot, water saturation for most samples ranges between 3 and 10% (green curves), which is the

range of water saturations measured from core samples. On the other hand, the traditional

Archie’s models using single porosity gives poor water saturation estimates of 25% to 50% (blue

80
straight lines). Dots with porosity lower than 0.04 should not be estimated using curves of φb

=0.04 (φb ≤ φ), water saturation for those low porosity intervals can be estimated from curves of

φb =0.025 (yellow curves).

Figure 4-3 Crossplot of dual porosity (fracture + matrix) water saturation lines using Aguilera’s

dual porosity model in Pickett a Plot. Dots are from log data of a well in Wapiti pool, Falher

formation, Western Canada Sedimentary Basin (WCSB). Water saturation lines equal to 100%,

50%, 10% and 3% are plotted for Archie’s single porosity model and Aguilera’s dual porosity

model (φb = 0.04 and 0.025).

From the dual porosity water saturation curves, we can see that m is a r eally an important

parameter for water saturation calculation. As we move into tight formations with low porosity,

the curves get closer to each other, meaning that water saturation is more sensitive to porosity

changes in these formations. For high water saturation formations, the distance between the

curves is even narrower (In Figure 4-3, the distance between 100% and 50% water saturation

81
curves is smaller than the distance between 10% and 3% water saturation curves). Water

saturations are significantly overestimated in Wapiti case using Archie’s single porosity model.

4.3 Water saturation model for shale formations

In shales, the clay content of the matrix greatly influences the matrix cementation factor. Acting

like an ion-selective membrane, clay makes electric conductivity easier. More clay content will

result in smaller values of m. The cementation factor for shale is usually smaller than 2, and

values as low as 1.45 and 1.54 have been reported in previous studies (Yu and Aguilera 2011,

Aguilera 1978). Organic material density may also change within a formation due to different

total organic carbon (TOC) content and different levels of maturity. All these unknown factors

make it difficult to evaluate the shale formation even with advanced logging tools and core

measurements.

Furthermore, all the hydraulic fractures created after logging form a complicated fracture

network, which adds another dimension of difficulty for characterization and modeling. The

matrix and organic material properties might also be influenced by the hydraulic fracturing job.

Studies based on m ineral-based formation evaluation of shale gas reservoir showed successful

results (Quirein et al. 2010); (Ramirez et al. 2011). However, these complicated evaluation

packages need extensive input data sets including some advanced tools such as XRD analysis of

core and geochemical measurements, which are not usually available.

The purpose of our study is to provide a simplified model to estimate the cementation exponent

m and water saturation based on minimum input data available from standard well logs. If more

82
input data from SEM/TEM image analysis, core/cutting analysis and advanced logs are available

these parameters can be incorporated into the proposed model for more accurate evaluation. As

we demonstrate in our study, SEM imaging can be useful to differentiate kerogen porosity and

inorganic matrix porosity. With reasonable sample size, we can obtain useful porosity values,

providing an additional method to correlate with log/core porosities, and providing valuable

input data for the models proposed in this study.

Organic material and kerogen have been used interchangeably in published literature regarding

shale evaluation without clear differentiation. There might be some other types of organic

materials other than kerogen with very different structures and characteristics, but kerogen is still

the major component in solid organic materials in shale. In this study, we consider organic

material to being the same as kerogen.

The cementation exponent of shale is expressed by the following equation (see Appendix A for

model derivation):

𝐥𝐨𝐠�𝒄𝒗𝒌𝒆𝒓 +(𝟏−𝒗𝒌𝒆𝒓 )𝝓𝒃 −𝒎𝒃 �


𝒎= − 𝐥𝐨𝐠 𝝓
4-15

𝑻𝑶𝑪∙𝜿 𝝆𝒌𝒐
𝒗𝒌𝒆𝒓 = 𝝆𝒓
𝝆𝒓 = 𝝆 4-16
𝒃𝒖𝒍𝒌

Where:
TOC = total organic carbon (weigh fraction)
vker = kerogen volume percentage (volume fraction)
ρ r = relative density of kerogen
κ = kerogen conversion factor
c = kerogen resistivity factor

83
Kerogen volume can be estimated directly from lab measurements. Here we use a commonly

reported kerogen conversion factor. Because TOC does not account for other elements that may

occur within kerogen (H, O, N, S), a conversion factor κ is used to account for these elements.

Although this factor can be slightly different for kerogens of different types and maturity, 1.2 is

generally considered to be a reasonable value (Lewis et al. 2004).

Similarly to the previous discussion, to plot water saturation lines in a Pickett Plot, we need to

assume that either φb or kerogen volume percentage vker is constant. In this work φb is assumed to

be fixed for plotting water saturation lines due to a smaller variability of φb compared to kerogen

percentage. The relationships between kerogen volume percentage and φb is shown in equation

4-19.

Total porosity is the sum of matrix porosity φm and kerogen porosity φkt. Both porosities are

associated with the whole composite system.

𝝓 = 𝝓𝒎 + 𝝓𝒌𝒕 4-17

𝝓𝒌𝒕 = 𝝓𝒌 ∙ 𝝊𝒌𝒆𝒓 4-18

From Equation 4-17 and 4-18, we can obtain:

𝝓−𝝓𝒃
𝑣𝑘𝑒𝑟 = 𝜙 4-19
𝑘 −𝝓𝒃

Inserting Equation 4-19 into 4-15, we obtain Equation 4-20. Equation 4-20 is inserted into

Archie’s equation to calculate the corresponding Rt for plotting water saturation lines in Pickett

Plot.

84
𝝓−𝝓𝒃 𝝓−𝝓𝒃
𝐥𝐨𝐠�𝒄 +�𝟏− �𝝓𝒃 −𝒎𝒃 �
𝝓𝒌 −𝝓𝒃 𝝓𝒌 −𝝓𝒃
𝒎= − 𝐥𝐨𝐠 𝝓
4-20

A case study from the Haynesville shale gas reservoir (Ramirez et al. 2011) is used to

demonstrate the validity of the dual porosity shale model developed in this thesis and for plotting

water saturation lines in the Pickett Plot. Core analysis data from 10570 t o 10619 f eet is

available for the studied well including porosity, water saturation, standard log data and a

geochemical log from 10524 t o 10619 f eet. A water resistivity Rw of 0.048 ohm-m at 25°C is

calculated based on r eported formation water salinity (Yu and Aguilera 2011) and then is

converted to Rw at reservoir temperature.

The pore space within kerogen φk is assumed to be 50%, which is based on image analysis data

of a backscattered SEM image of a kerogen body (Sondergeld et al. 2010). Our TEM study also

shows 40~60% porosity within kerogen of the Nordegg sample. Constant c, the kerogen

resistivity factor, is dependent on φk. The bigger φk, the smaller is c. When φk = 100%, kerogen is

assumed to be the same as a vug, then c=1. Constant c is assumed to be roughly proportional to

1/φk, so we assume c=2.

The cementation exponent of the matrix mb is determined by combining water saturation curves

with TOC curves in a Pickett Plot. TOC curves were plotted on the Pickett Plot based on

Passey’s theory and Yu’s methodology (Passey, Creaney, and Kulla 1990, Yu and Aguilera

2011). By overlapping 80% water saturation lines with 0% TOC curves at lower porosity, we can

derive mb to be 1.63 (80% is the highest water saturation from average of logging data in the

85
formation; this corresponds to 0% TOC). Available data from cores and well logs are listed in

Table 4-1.

Matrix porosity φb was calculated from total porosity and kerogen volume percentage assuming

50% φk. Coefficient of variation (Cv) is a normalized measure of dispersion in a probability

distribution. The Cv of φb is lower than the Cv of kerogen volume percentage. A similar trend

was also found in a tight gas case study shown in Appendix A. Less variability in φb makes it

reasonable to plot water saturation lines assuming fixed φb instead of fixed TOC. We used an

average value of φb equal to 0.04 to plot water saturation lines in Figure 4-4 and Figure 4-6.

TOC curves were plotted using Yu’s (2011) method to derive the cementation exponent of the

matrix mb which is an important parameter for developing and plotting water saturation lines.

These TOC curves, based on Passey’s model, match really well the TOC measurements from

core data (average ~2%). By combining them, we can also see the relationship between TOC and

water saturation.

The water saturations (25% to 40%) in Figure 4-4 are consistent with core data and water

saturations (20%~70%) shown in Figure 4-6 calculated from logs using in-house ConocoPhillips

interpretation models (Ramirez et al. 2011). The dotted horizontal line at porosity 0.04 i s the

cutoff for these water saturation lines (φ ≥ φb). Water saturations of the two intervals with

porosity lower than 0.04 cannot be accurately estimated with 0.04 of φb. For these two intervals

86
an average φb equal to 0.0265 gi ves water saturations very close to the core data presented i n

Table 1 (Figure 4-5).

Table 4-1 Core and log data for Haynesville shale case study

φ (cal. DT
TOC
c φ (core) from (μ R90(Ωm) Vker (%) φb Sw (core)
(wt%)
log)
sec/ft)
3222.6 0.068 0.068 80 20 4.4 1.83 0.046 0.31
3223.9 0.071 0.073 82 25 4 1.67 0.051 0.34
3225.4 0.078 0.076 83 33 7.2 3.00 0.042 0.36
3226.9 0.073 0.065 79 23 4.4 1.83 0.051 0.36
3228.3 0.064 0.065 79 19 5 2.08 0.039 0.34
3230.0 0.065 0.081 85 21 5.6 2.33 0.037 0.33
3231.3 0.067 0.084 86 21 5.4 2.25 0.04 0.32
3233.1 0.062 0.038 69 18 8.4 3.50 0.02 0.30
3234.5 0.034 0.032 67 45 1 0.42 0.029 0.27
3236.1 0.034 -0.011 51 25 2 0.83 0.024 0.46
Ave. 0.06 0.06 4.74 1.98 0.04 0.34
Cv 0.46 0.46 0.28 0.15

87
Figure 4-4 Water saturation curves plotted for Haynesville shale using dual porosity model

developed in this thesis (φb = 0.04)

Figure 4-5 Water saturation curves for Haynesville shale using dual porosity model (φb = 0.027)

88
Figure 4-6 Water saturation curves plotted for Haynesville shale using dual porosity model

developed in this thesis. Data points from available well logs (φb = 0.04)

With this model we can also explain the difficulties found using traditional straight lines from

Archie’s single porosity model to fit water saturations in Pickett Plots (as shown by circled spots

in Figure 4-7 with water saturations bigger than 100%). T he model also helps to clarify the

confusing area associated with low porosity and more than 100% water saturation (shaded area

in Figure 4-7) in Yu’s study (Yu and Aguilera 2011).

Sensitivity studies showed that c and φk have very small influence in the plotted water saturation

curves (Figure 4-8 and Figure 4-9). The most important factors are Rw, mb and φb. Rw data are

usually available and/or can be obtained at low cost. Matrix porosity φb and its distribution can

89
be obtained from standard logging suits preferentially calibrated with SEM image analysis and

core data. The porosity exponent of the matrix mb can be derived by overlapping the highest

water saturation lines with zero percent TOC in the Pickett Plot.

Figure 4-7 Pickett plots integrated with TOC curves (Yu and Aguilera 2011)

From the integrated Pickett Plots and sensitivity studies, we conclude that the resistivity is not

sensitive to porosity changes, but it is sensitive to water saturation changes. Higher water
saturation corresponds to lower resistivity. Also, higher TOC corresponds to lower water

saturation. These observations from integrated Pickett Plots are consistent with commonly

accepted hypothesis. Water generally does not exist in kerogen. All the water remains in the

inorganic matrix (rock/clay). With total porosity at roughly a similar level; the higher the TOC,

the higher is the kerogen percentage. This means lower inorganic matrix porosity and lower

water saturation. When TOC is close to 0, the shale formation is probably close to the boundary

to other non-shale formation.

If the hydrocarbon in the shale has not migrated to those formations, the non-shale formation will

have extremely high water saturations. Recently, one petrophysical study presented a new model

for water saturation evaluation based on t he premise that the hydrocarbon phase occupies the

kerogen-created porosity while water occupies the non-kerogen matrix porosity (Alfred and

Vernik 2012). This premise seems to agree with the trend we observe in our integrated Pickett

Plots. However, their model does not consider that hydrocarbon could also present in inorganic

matrix.

91
Figure 4-8 Sensitivity analysis of water saturation curves in shale reservoir by changing

kerogen porosity φk.

Figure 4-9 Sensitivity analysis of water saturation curves in shalereservoir by changing kerogen

resistivity factor c

92
As with all new methods it will take some time to test and recognize the value of the procedure

developed in this paper. However, the case study presented above supports the use of the method

for those cases with limited amounts of wireline data. A word of caution, however, is warranted:

The method is not meant to replace detailed studies where all types of data are available. It is

meant for quick analysis in the vast majority of wells where, in practice, there is very limited

amount of data.

4.4 Triple porosity model including inorganic matrix, kerogen and natural fractures.

The inclusion of natural fractures in the proposed water saturation model leads to a similar form

of the triple porosity model developed by Al Ghamdi et al (Al-Ghamdi et al. 2011)

(𝟏−𝝂𝑲𝒆𝒓 )𝟐
−𝒍𝒐𝒈�𝒄𝝂𝑲𝒆𝒓 + �
𝝓𝟐 +(𝟏−𝝓𝟐 −𝝂𝑲𝒆𝒓 )⁄𝝓𝒃 −𝒎𝒃
𝒎= 𝒍𝒐𝒈𝝓
4-21

Derivation of this equation is very similar to the triple porosity model composed of non-

connected vugs, natural fracture and matrix(Al-Ghamdi et al. 2011). The main problem with the

application of this model is the difficulty in securing reliable fracture porosity values in shales.

This problem is currently under investigation.

93
4.5 References

Aguilera, M.S., R. Aguilera. 2003. Improved models for petrophysical analysis of dual porosity
reservoirs (in Petrophysics 44 (1).

Aguilera, R. 1978. Log Analysis of Gas-bearing Fracture Shales in the Saint Lawrence Lowlands
of Quebec. Proc., SPE 7445, pa per presented at the 53rd SPE annual fall technical conference
and exhibition in Houston, Texas, USA.

Al-Ghamdi, A., B. Chen, H. Behmanesh et al. 2011. An Improved Triple-Porosity Model for
Evaluation of Naturally Fractured Reservoirs (in SPE Reservoir Evaluation & Engineering 14
(4): 377-384.

Alfred, D., L. Vernik. 2012. A New Petrophysical Model for Organic Shales. Proc., paper
presented at the SPWLA 53rd annual loggin symposium, Cartagena, Columbia.

Archie, G.E. 1942 The Electrical Resisitivity as an Aid in Determining Some Reservoir
Characteristics (in J. Pet. Tech. 5: 1-8, Trans. AIME, Volume 146, 19 42, pages 54-62. SPE-
942054-G.

Davis, G. http://www.g-davis.com/.

Lewis, R.E., D. Ingraham, M. Pearcy et al. 2004. New Evaluation Techniques for Gas Shale
Reservoirs. Proc., Reservoir symposium, Schlumberger.

Passey, Q.R., S. Creaney, J.B. Kulla. 1990. A practical model for organic richness from porosity
and resistivity logs (in AAPG Bulletin 74 (12): 1777-1794.

Quirein, J., J. Witkowsky, J. Truax et al. 2010. Integrating Core Data and Wireline Geochemical
Data for Formation Evaluation and Characterization of Shale Gas Reservoirs. Proc., SPE
134559, paper presented at the SPE annual technical conference and exhibition, Florence, Italy,
18p.

Ramirez, T.R., J.D. Klein, R.J.M. Bonnie et al. 2011. Comparative Study Formation Evaluation
Methods for Unconventional Shale-Gas Reservoirs: Application to the Haynesville Shale
(Texas). Proc., SPE 144062, pa per presented at SPE North American Unconventional Gas
Conference in Woodlands, Texas, USA.

Schlumberger. Schlumberger chart Gen-9. In.

Sondergeld, Carl H., Ray J. Ambrose, Chandra S. Rai et al. 2010. Micro-structural studies of gas
shales. Proc., SPE 131771, paper presented at the unconventional gas conference in Pittsburg,
PA, USA.

Yu, G., R. Aguilera. 2011. Use of Pickett Plots for Evaluation of Shale Gas Formations. Proc.,
SPE 146948, paper presented at the SPE annual technical conference and exhibition in Denver,
CO, USA.

94
Chapter Five: Uncertainty Analysis of Shale Gas Simulation:
Consideration of Basic Petrophysical Properties

Uncertainties in basic petrophysical parameters in shale gas reservoirs are discussed in Chapter

2. These parameters include porosity, permeability, gas adsorption, water saturation and TOC.

Current available methods to evaluate shale gas petrophysical properties are summarized. These

methods include core laboratory measurements and well logs. Sources of uncertainty in these

measurements are discussed.

In this chapter, uncertainty analysis using commercial simulation software is performed to

understand how these uncertainties influence reservoir simulation results and predictions.

Sensitivity and uncertainty studies in this thesis are based on typical ranges of shale gas

properties extracted from published evaluation reports and open data. R esults from this study

will help guide the focus of future research and technology development for shale gas reservoirs.

5.1 Simulation model

A simple shale gas model (Figure 5-1) has been used for a sensitivity study of basic

petrophysical parameters and hydraulic fracture properties. Commercial simulator GEM and

CMOST from CMG is used for this study. The approach (CMG 2012) fuses a dual permeability

model meaning that gas can flow from matrix to fractures and from fractures to matrix.

Adsorption is handled with the use of Langmuir isotherm. Gas diffusion is handled by specifying

tortuosity and the diffusion coefficient. Non-darcy flow is modeled with Forchheimer equation.

Logarithmic refinement is used near the hydraulic fracture (Rubin 2010).

95
Figure 5-1 Top view of model layout for multi-stage hydraulically fractured horizontal well

Tested parameters including basic petrophysical properties and some hydraulic fracture

parameters included in the sensitivity analysis are listed in Table 5-1. These parameters and

tested range were chosen to represent typical data range measured or estimated for shale

formations in the WCSB (Rokosh et al. 2012). Close to 500 simulation cases were chosen using

an optimized algorithm in CMOST to minimize the number of simulation runs needed.

Sensitivity results are based on these tested ranges and specific simulation model used in this

study. Results from this study may not be directly adapted to other cases without careful

investigation of specific data range and the simulation model to be used. It is better to use the

96
results developed in this thesis as a general guide/reference for exploitation of shale formations

in Alberta.

Table 5-1 Investigated parameters in sensitivity analysis

Parameter Abbrev. Description Range Unit

MatrixPerm Matrix permeability 2.5×10-5 ~ 0.1 md

GradientHFPerm Hydraulic fracture permeability gradient 0 ~ 0.15 md/ft

PermNatFrac Natural fracture permeability 10-5 ~ 1.6×10-4 md

MinHFPerm Minimum hydraulic fracture permeability 10 ~ 30 md

MaxAdsMass Maximum methane max adsorption mass percentage 0.05 ~ 0.15 fraction

Matrix Porosity Matrix porosity 0.01 ~ 0.06 fraction


Pressure dependence of permeability due to fracture
FracCloseMult 0 ~ 1.2 N/A
compressibility (a multiplier)
Initial_Sw_Matrix Initial matrix Sw 0.1 ~ 0.3 fraction

LangmuirPres Langmuir pressure (PL) 250 ~ 750 psi

ConWaterHF Connate water in hydraulic fracture 0 ~ 0.06 fraction

ConGasHF Connate gas in hydraulic fracture 0 ~ 0.06 fraction

Initial_Sw_HF Initial hydraulic fracture Sw 0.2 ~ 0.6 fraction

The permeability and length of the hydraulic fracture are controlled by hydraulic fracture

permeability gradient, and maximum and minimum hydraulic fracture permeabilities. As the

hydraulic fracture goes from well to tip, the permeability decreases away from the wellbore as

fracture width narrows down. In this model, a linear permeability gradient is assumed to include

the change of hydraulic fracture permeability as a function of distance from the well as shown in

Equation 5-1. If the minimum hydraulic fracture permeability is set equal to the matrix

97
permeability, the hydraulic fracture length is decided by maximum hydraulic fracture

permeability and the gradient of hydraulic fracture permeability (CMG 2012) as shown in Figure

5-2.

𝐾𝐻𝐹 = 𝑀𝑎𝑥(𝐾𝐻𝐹𝑚𝑎𝑥 − 𝐺𝑟𝑎𝑑 ∗ 𝐿, 𝐾𝐻𝐹𝑚𝑖𝑛 ) 5-1

where:

𝑲𝑯𝑭 : hydraulic fracture permeability

𝑲𝑯𝑭𝒎𝒂𝒙 : maximum hydraulic fracture permeability

𝑲𝑯𝑭𝒎𝒊𝒏 : minimum hydraulic fracture permeability

𝑮𝒓𝒂𝒅 : hydraulic fracture permeability gradient

𝑳 : distance to well

Figure 5-2 Hydraulic fracture permeability as a function of distance from well

98
Tornado plots for sensitivity analyses are shown in Figure 5-3, Figure 5-4 and Figure 5-5. These

three cases include different ranges of matrix permeability (from 10-5 to 10-1 md). Such a big

permeability range should not be covered in a s ingle case because it tends to cause simulator

running errors and masking out effect of other parameters. In all these studies, similar ranges are

used in terms of their relative percent change for reasonable comparison of sensitivities between

different parameters. For all three cases, parameters that are sensitive to cumulative gas

production include matrix permeability, matrix porosity, hydraulic fracture permeability

gradient, maximum hydraulic fracture permeability, natural fracture permeability and maximum

gas adsorption. Other parameters showed relative small effect on c umulative gas production

based on the simulation results. However, these parameters should not be neglected in other real

reservoirs without careful analysis. They might also have big impact in other aspects.

Matrix permeability is among the top two sensitive parameters in all three cases. This is

reasonable considering that the matrix permeability is the deciding parameter on t he gas flow

from matrix to fractures, which is important for gas production in all cases. Matrix porosity and

maximum gas adsorption are more important in shale gas reservoirs with relative higher

permeability. In the shale gas reservoir with 10-3 ~ 10-1 md matrix permeability (Figure 5-5), the

matrix porosity becomes the most sensitive factor.

As the matrix permeability increases, the deciding factor of cumulative gas production is more

related to OGIP, which is dependent on porosity and adsorbed gas; at the same time, the

hydraulic fracture properties become less important. When permeability decreases to the micro-

99
darcy level, the hydraulic fracture permeability gradient becomes the most sensitive parameters

as shown in Figure 5-3. Equation 5-1 shows that the hydraulic fracture half-length is directly

related to hydraulic fracture permeability. This means that the properties of hydraulic fractures

including half-length and permeability are crucial in very low permeability shale gas reservoirs.

The order of important parameters in sensitivity studies are shown in Figure 5-6.

From ERCB’s shale gas report on A lberta shale formations including Duvernay, Montney and

Banff/Exshaw shales {Anderson, 2010 #86; Anderson, 2010 #87), permeability measurements

range from 10-4 to 1 md with most of them around ~10-2 md. The sensitivity analysis presented in

this thesis suggests that for shale gas reservoirs with this range of matrix permeabilities, the three

most sensitive parameters leading to the largest uncertainties in cumulative gas production

volumes are matrix porosity, matrix permeability and maximum gas adsorption. While

sensitivity studies presented in this paper are not applicable to every actual case, the results

highlight how important it is to secure the best possible estimations of basic parameters such as

porosity and gas adsorption.

100
Figure 5-3 Tornado plot of cumulative gas production showing sensitivity of tested parameters (matrix

permeability 2.5×10-5~ 4×10-4 md).

Figure 5-4 Tornado plot of cumulative gas production (matrix permeability 4×10-4 ~ 6.4×10-3 md).

101
Figure 5-5 Tornado plot of cumulative gas production (matrix permeability 6.4×10-3 ~ 0.1 md).

Figure 5-6 Order of importance of tested parameters in sensitivity studies.

102
5.2 Discussion

Careful sensitivity studies are important for every field case simulation. This is not only for use

in history match, but also as a guide for focus points of data collection and future research.

Matrix permeability, half-length and permeability of hydraulic fracture, matrix porosity, gas

adsorption, and effective natural fracture permeability are among the most important parameters

according to the sensitivity analysis discussed in this thesis. Among these parameters, porosity is

the most basic petrophysical parameter of shale reservoirs, which can be measured directly and

accurately in the laboratory. More research should be focused on t he optimization and

standardization of porosity measurements via a detailed core analysis protocol. This will permit

more valuable correlations between log porosity and porosity values from core.

Permeability is probably the most difficult parameter to quantify. It is important to understand

the relationship between measured permeability and the input permeability value in the

numerical simulation model. Permeabilities of natural fractures and hydraulic fractures cannot be

measured directly but can be reflected by the effective permeability determined from well

testing, minifracs and/or history matching.

Hydraulic fracture half-length and fracture network distribution can be estimated from

microseismic data if available. Otherwise they can be estimated from details of the hydraulic

fracturing job and geomechanical properties of the formation.

103
Having high confidence in the porosity and gas adsorption input data, the history match can

focus on parameters that are much more difficult to obtain including permeability and hydraulic

fracture properties.

5.3 Summary

Petrophysics modeling of shale gas reservoirs is very difficult due to limited amount of

experimental and log data, and lack of standard industry experimental protocols. Uncertainties in

basic petrophysical properties are summarized and discussed in this chapter. Sensitivity analysis

performed using a commercial numerical simulator demonstrates the importance of minimizing

the uncertainty of basic parameters utilized as input data. Sensitivity analysis provides a quick

method to figure out the most important deciding factors for commercial production of a given

well, with the possibility of extrapolating results to the whole reservoir. Sensitivity analysis can

help the operator to choose the most cost effective tools for data collection, and to guide the

direction for future research in shale gas petrophysics.

5.4 References

CMG. 2012. Shale gas simulation using GEM. Proc., computer modelling group 2012 training,
Calgary, AB, Canada.

Rokosh, C.D., S. Lyster, S.D.A. Anderson et al. 2012. Summary of Alberta's Shale- and
Siltstone-Hosted Hydrocarbons (Reprint).

Rubin, B. 2010. Accurate Simulation of Non Darcy Flow in Stimulated Fractured Shale
Reservoirs. Proc., SPE 132093-MS, SPE Western Reginall Meeting in Anaheim, California,
USA.

104
Chapter Six: Summary, Conclusions and Recommendations

The research discussed in this thesis is focused on the understanding of the basic petrophysical

properties and the fundamental structure of shale at nanoscale. The study started with image

analysis using advanced instrumentations with nanoscale resolution to understand the basic

nanopore structure of shale. This was followed by petrophysics modeling of shale petroleum

reservoirs, and ended up with sensitivity studies by numerical simulation.

Shale provides a promising and relatively clean future energy source. But much more research is

needed to tap this resource economically.

Major conclusions are as follows:

• SEM, TEM and AFM provide useful imaging tools for visualizing the shale structure at a

nanoscale level. Although there are limitations in reliable quantification of macro

characteristics, these powerful microscopes can provide very valuable insight into

fundamental understanding of shale rocks.

• SEM is the most practical tool for image analysis of cores and drill cuttings, as it

provides direct, fundamental structure image at nanoscale resolution. It also has

quantitative capabilities for estimating porosity, TOC and elemental composition, which

provide a useful supplement to standard well log and core analysis. Furthermore the SEM

imaging time is comparable to TEM and much lower than AFM.

• A new petrophysical dual porosity model has been developed for calculating the

cementation exponent m and water saturation in shale petroleum reservoirs.

105
• The cementation exponent m changes continuously because of variability of fracture and

kerogen porosity across the shale formation. As a result, the water saturation lines are not

straight in Pickett plots; instead they show up as curves.

• The combination of porosity, resistivity, cementation exponent, water saturation and

TOC curves on Pickett Plots helps to understand the relationship between these important

characteristics in shale reservoirs. The curves show that water saturation is much more

sensitive to resistivity change but not so much to porosity change.

• A sensitivity study using a commercial simulator has been done on s hale formation for

studying basic petrophysical and hydraulic fracture parameters. Results show that the

importance of matrix porosity and hydraulic length varies depending upon t he matrix

permeability of shale formation.

Major recommendations are as follows:

• Continue research into fundamental characteristics of shale rocks with extremely low

permeability. It is anticipated that this will help future evaluation and exploitation of

shale formations,

• Corroborate with additional field data the theoretical triple porosity petrophysics model

presented in Chapter 4.

• Extend the petrophysical models proposed in this thesis to the case of multi-porosity

shale reservoirs.

106
Appendix A: Derivation of dual porosity model including

kerogen for shale petroleum reservoirs

Matrix and kerogen are considered as two resistors in series (Figure A-1).

L A
r=R R=r
A L

(R = resistivity r = resistance A = area of cross-section L = length)

rtotal = rmatrix + rko

At 100% water saturation,

L L L
Rt = R0 1 + cRw 2
A A A
L1 L2
Rt = R0 + cRw
L L

Rt = Ro (1 − v ker ) + cRw v ker

The basic formation evaluation equations for the composite system of inorganic matrix and

kerogen are,

R t = Ft R w , Ft = ϕ-m

The basic formation evaluation equations for only the inorganic matrix are{Aguilera, 2003 #28},

R o = FR w , F = ϕb -mb

Ft R w = FR w (1-νker ) + cR w νker

φ − m Rw = φb
− mb
Rw (1 − vker ) + cRwvker

107
φ − m = φb
− mb
(1 − vker ) + cvker

m=
[
log cvker + (1 − vker )φb
− mb
]
− log φ

Figure A-1 Sketch of dual porosity model for shale

108

Vous aimerez peut-être aussi