You are on page 1of 106

SPRINGER BRIEFS IN

PETROLEUM GEOSCIENCE & ENGINEERING

Vahid Tavakoli

Geological Core
Analysis
Application
to Reservoir
Characterization
SpringerBriefs in Petroleum Geoscience
& Engineering

Series editors
Dorrik Stow, Heriot-Watt University, Edinburgh, UK
Mark Bentley, AGR TRACS International Ltd, Aberdeen, UK
Jebraeel Gholinezhad, University of Portsmouth, Portsmouth, UK
Lateef Akanji, King’s College, University of Aberdeen, Aberdeen, UK
Khalik Mohamad Sabil, Heriot-Watt University, Putrajaya, Malaysia
Susan Agar, Houston, USA
Kenichi Soga, Department of Civil and Environmental Engineering, University of
California, Berkeley, USA
A. A. Sulaimon, Department of Petroleum Engineering, Universiti Teknologi
PETRONAS, Seri Iskandar, Malaysia
The SpringerBriefs series in Petroleum Geoscience & Engineering promotes and
expedites the dissemination of substantive new research results, state-of-the-art
subject reviews and tutorial overviews in the field of petroleum exploration,
petroleum engineering and production technology. The subject focus is on upstream
exploration and production, subsurface geoscience and engineering. These concise
summaries (50–125 pages) will include cutting-edge research, analytical methods,
advanced modelling techniques and practical applications. Coverage will extend to
all theoretical and applied aspects of the field, including traditional drilling,
shale-gas fracking, deepwater sedimentology, seismic exploration, pore-flow
modelling and petroleum economics. Topics include but are not limited to:

• Petroleum Geology & Geophysics


• Exploration: Conventional and Unconventional
• Seismic Interpretation
• Formation Evaluation (well logging)
• Drilling and Completion
• Hydraulic Fracturing
• Geomechanics
• Reservoir Simulation and Modelling
• Flow in Porous Media: from nano- to field-scale
• Reservoir Engineering
• Production Engineering
• Well Engineering; Design, Decommissioning and Abandonment
• Petroleum Systems; Instrumentation and Control
• Flow Assurance, Mineral Scale & Hydrates
• Reservoir and Well Intervention
• Reservoir Stimulation
• Oilfield Chemistry
• Risk and Uncertainty
• Petroleum Economics and Energy Policy

Contributions to the series can be made by submitting a proposal to the responsible


Springer contact, Charlotte Cross at charlotte.cross@springer.com or the Academic
Series Editor, Prof. Dorrik Stow at dorrik.stow@pet.hw.ac.uk.

More information about this series at http://www.springer.com/series/15391


Vahid Tavakoli

Geological Core Analysis


Application to Reservoir Characterization

123
Vahid Tavakoli
School of Geology, College of Science
University of Tehran
Tehran
Iran

ISSN 2509-3126 ISSN 2509-3134 (electronic)


SpringerBriefs in Petroleum Geoscience & Engineering
ISBN 978-3-319-78026-9 ISBN 978-3-319-78027-6 (eBook)
https://doi.org/10.1007/978-3-319-78027-6
Library of Congress Control Number: 2018934945

© The Author(s) 2018


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made. The publisher remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.

Printed on acid-free paper

This Springer imprint is published by the registered company Springer International Publishing AG
part of Springer Nature
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

Cores are the most important and reliable source of subsurface information in
reservoir studies. Despite this, few books have been published on this subject thus
far and fewer on the geological section of core analysis projects. Many students ask
me where we should start and what we should do after data gathering. This book
tries to fill this gap. It starts with an introduction and continues with the preparation
stages, actually, what happens before starting geological analysis of the cores. Also,
this introduction is important because porosity and permeability data are integrated
with the geological results to build a perfect framework for reservoir modeling and
property distribution within the reservoir. The book continues with microscopic
studies. What should be recorded and how? Macroscopic studies are considered and
data of interest are explained. Geochemistry helps to a better understanding of the
geological properties and therefore its principles and applications are considered.
The final chapter covers the integration of geological and petrophysical data to
determine the rock types as the main building blocks of the reservoir.
The book covers both academic and industrial aspects of geological core anal-
ysis and thus is applicable for both groups. Although written for petroleum geol-
ogists and reservoir engineers, the book can be useful as a reference for any student,
researcher, industry professional, or anyone who deals with cores. This book
evolved from my industrial experience on managing and analyzing cores from
various hydrocarbon fields and also my graduate courses on petroleum geology.
I did not include many photos in the text because I think that readers can observe
many of them with a simple search. Instead, I explain more about the fundamental
principles and use schematic illustrations for better understanding of the processes
or mechanisms.
The author is thankful to Dr. Mehrangiz Naderi-Khujin for designing and
illustrating most figures of the book. She also read the first version of the manu-
script and provided useful comments and suggestions for which I am grateful. She
is a talented geologist who is beside me during all stages of my academic and
industrial activities. The preparation of the book was made possible through the
help of my colleagues and students at the University of Tehran. I also appreciate the
cooperation of the private core analysis laboratories and the National Iranian Oil

v
vi Preface

Company (NIOC) who have helped me develop this book. In addition, I want to
thank all my family for their patience during the preparation of this book. I know
that with all care and attention, it is possible that I missed some points. Therefore
any critique, idea, or suggestion is welcome.

Tehran, Iran Vahid Tavakoli


February 2018
Contents

1 Core Analysis: An Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.1 Purpose of the Book . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Core Analysis Role in Reservoir Characterization . . . . . . . . . . . . 2
1.3 Core Analysis Plan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Coring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Core Preservation and Transfer . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.6 Retrieved Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.7 Routine Core Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.8 Special Core Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.9 Wire Line Log Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.10 Core Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2 Preparing for Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1 Core Gamma Logging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Core–Log Depth Matching . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3 Core CT-Scanning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4 Core Opening and Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.5 Dean–Stark Extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.6 Core Cleaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.7 Marking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.8 Plugging and Trimming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.9 Soxhlet Extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.10 Sidewall Coring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3 Microscopic Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1 Thin Section Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1.1 Thin Section Staining . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

vii
viii Contents

3.2 Quantitative Carbonate Petrography . . . . . . . . . . . . . . . . . . . . . . 32


3.2.1 Quantitative Siliciclastic Petrography . . . . . . . . . . . . . . . 40
3.3 Paleontology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.4 XRD Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.4.1 Sample Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.4.2 Bulk XRD Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.4.3 Clay Mineralogy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.5 SEM Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.5.1 Sample Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.5.2 EDX Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.6 Microscopic Uncertainties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4 Macroscopic Studies . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . 53
4.1 Core Slabbing and Resination . . . . . . . .. . . . . . . . . . . . . . . . . . 53
4.2 Core Description . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . 54
4.2.1 Core Photography . . . . . . . . . . .. . . . . . . . . . . . . . . . . . 59
4.3 Fracture Presentation . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . 61
4.4 Core Log Preparation . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . 62
4.5 Sequence Stratigraphy and Reservoir Zonation . . . . . . . . . . . . . . 65
4.6 Macroscopic Uncertainties . . . . . . . . ........ . . . . . . . . . . . . . 67
References . . . . . . . . . . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . 68
5 Geochemical Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.1 Sample Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.2 Isotopes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.2.1 Carbon Isotope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.2.2 Oxygen Isotope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5.2.3 Strontium Isotope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5.3 Elemental Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.4 Uranium Geochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
6 Rock Typing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6.1 Geological Rock Typing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6.2 Hydraulic Flow Units (HFU) . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
6.3 Reservoir Data Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
6.4 Defining Electrofacies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
6.5 Comparison and Final Results . . . . . . . . . . . . . . . . . . . . . . . . . . 97
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
Chapter 1
Core Analysis: An Introduction

Abstract Cores are a fundamental source of information for exploration, evalua-


tion, development, and production of any hydrocarbon field. Cores are a unique
source of some datatypes such as rock textural parameters or permeability. They
can be calculated or estimated based on other data sources and cannot be gained
directly from those data. Some others, such as porosity, are calibrated against core
analysis results. Cores are direct samples from the reservoir rocks that can be tested,
analyzed, and viewed by the researcher. A core analysis project starts from the
coring plan, coring, and core preservation, and continues with three main phases
including routine, geological, and special core analysis. Some additional stages are
also included such as geomechanics or geochemistry. Various experts are involved
in a core planning task. They consider all variables including requirement, cost, and
risk to decide the different aspects of coring and core analysis. After coring, cores
are transferred to the laboratory. Core analysis starts with the core gamma logging
and whole core CT-scanning. Basic petrophysical parameters using single-phase
fluid are measured on the cores in the routine stage. This step also includes core
handling and preparation for routine, special, and geological analysis. Geological
analysis includes thin section preparation, microscopic and macroscopic studies,
and rock typing. Final data are compared with the wire line logs and distributed to
the interwell space. Dynamic data are provided using multiphase fluid tests in a
special section. All data are integrated to reconstruct the rock and fluid distribution
within the reservoir.

1.1 Purpose of the Book

Now, much time has passed from the first coring in a reservoir. A lot of data and
experience have been gathered from coring and core analysis of the hydrocarbon
reservoirs. There are still more uncertainties in petroleum exploration and pro-
duction. Various disciplines such as geology, reservoir engineering, petrophysics,
geochemistry, and geomechanics are involved in a core analysis project, from core
planning to core archiving. Regarding the importance of the cores and core analysis,

© The Author(s) 2018 1


V. Tavakoli, Geological Core Analysis, SpringerBriefs in Petroleum
Geoscience & Engineering, https://doi.org/10.1007/978-3-319-78027-6_1
2 1 Core Analysis: An Introduction

some textbooks have been written (e.g., McPhee et al. 2015) for this purpose. Other
texts cover some aspects of this task, especially on reservoir engineering discipline
(e.g., Tiab and Donaldson 2015). These books have little information on the geo-
logical aspects of core analysis, from microscopic to macroscopic steps. The
methods for geological data integration with basic petrophysical parameters for
better understanding of the geological effects on reservoir behavior are also not yet
understood perfectly.
The purpose of this book is to explain the geological part of a core analysis
project from both scientific and industrial points of view. As a matter of fact,
geology is the base of any reservoir evaluation and thus direct geological data from
cores are very important for the reservoir analysis team. The book explains the
geological core analysis from core preparation to the end of analysis and archiving.
Both microscopic and macroscopic analyses are included. Furthermore, geochem-
ical measurements and analysis complete the previous tasks and the final results are
integrated to construct the building block of the reservoir, the rock types.
Subsequently, these findings are arranged based on sequence stratigraphic concepts
for reservoir zonation.

1.2 Core Analysis Role in Reservoir Characterization

The basic information for reservoir characterization includes porosity, permeability,


water saturation, and net-to-gross ratio. These data are retrieved from two general
sources including direct and indirect measurements. The direct sources of reservoir
information include core and cutting samples and indirect methods include wire line
logging, well tests, and geophysical surveys. Drill cuttings, small pieces of rocks
cut by the bit, give a limited type of information. Porosity and permeability cannot
be retrieved from the cuttings. The problem of lag time between drilling and their
collection time at the well top is a serious error in analyzing these samples. Also, it
is possible that a fragment of rock detaches from the well wall after drilling and
before casing and mixes with the cuttings of the other intervals.
The preliminary coring tools were developed at the end of the 1900s. At the
beginning of the 1920s, the first effective coring tools were introduced to the drilling
industry (Anderson 1975). Considerable improvements were made in the following
years. Currently, there are several types of coring tools. Selecting the coring tool is
dictated by the type of reservoir rocks and core analysis purposes. Regardless of the
coring method, the result is a cylindrical rock sample from the reservoir. These
samples are representative of the reservoir rocks and the most reliable source of
information for reservoir studies. They are direct samples of the rocks that could be
viewed and touched by humans and tested by laboratory instruments. The static and
dynamic data are retrieved from the cores to understand the exact and accurate rock
properties and their behavior. Basic as well as advanced petrophysical parameters
and fluid flow properties using both single- and multiphase fluids can be obtained
from the cores. Wire line logs are calibrated with core data. The most accurate
1.2 Core Analysis Role in Reservoir Characterization 3

Fig. 1.1 Data retrieved from core analysis and their role in reservoir characterization (Courtesy of
M. Naderi)

porosity, for example, is obtained from the cores. Calculating the sonic velocity in a
rock matrix is another example. Geochemical and geomechanical tests on the core
samples reveal many aspects of the reservoir rocks. Most of the geological data such
as rock texture, pore type, sedimentary structures, and clay mineralogy are derived
from the cores. These data could also be retrieved from other sources but cores are
the most accurate source in many cases. Data from cores are compared with the other
sources and their integration helps to determine the most accurate static and dynamic
characteristics of the reservoir rocks (Fig. 1.1).

1.3 Core Analysis Plan

The first and main question before starting a coring job is about the importance of
the cores for the reservoir evaluations. Is this really necessary? In most cases the
answer is yes. Coring and core analysis are not expensive compared to the overall
4 1 Core Analysis: An Introduction

budget of well drilling and completion. Nevertheless, cores have vital information
for reservoir evaluations and assessments. A team of geologists, petrophysicists,
reservoir engineers, drillers, and production personnel begin the core planning by
listing the objectives of the job. The ultimate goal is to have more understanding of
the reservoir properties. Constraints in timing and budget should also be considered.
Selecting the coring tools is based on the rock properties of the target zone. The
hole size and environmental conditions such as temperature and pressure also play
an important role in this selection.
Along with a coring plan, a core analysis plan also should be prepared. This is
very important because it determines the future of the cores and type of data
retrieved from them. The period of analysis for any task must be completely
scheduled. A work breakdown structure (WBS) should be prepared for each part of
the analysis. An example of a geological core analysis plan is illustrated in
Table 1.1.
It is worth mentioning that coring and core analysis are designed for each well
individually and are specific for that well. Which part of the analysis, including
routine, geological, special, geochemistry, and geomechanics, is really necessary?
The prepared data should satisfy the purpose.

Table 1.1 An example of a geological core analysis program for 100 m of cores. The number of
samples and time duration are flexible based on available core length, laboratory potential, and
number of personnel
No. Task Quantity Unit Month
1 2 3 4 5 6 7 8
1 Microscopic studies
1.1 Thin section preparation 400 Sample ✖
1.2 Petrographical analysis 400 Sample ✖
1.3 Micropaleontology 400 Sample ✖
1.4 Sequence stratigraphy 100 Meter ✖ ✖
1.5 Reservoir rock typing 400 Sample ✖
2 XRD analysis
2.1 Selection and preparation 20 Sample ✖ ✖
2.2 XRD analysis 20 Sample ✖ ✖
3 SEM analysis and imaging
3.1 Selection and preparation 20 Sample ✖
3.2 SEM photography 20 Sample ✖
4 Macroscopic core analysis
4.1 Macroscopic core description 100 Meter ✖ ✖
4.2 Fracture study 100 Meter ✖ ✖ ✖
4.3 Core photography (as whole 100 Meter ✖ ✖
cores and close‐up)
4.4 Core log preparation 100 Meter ✖ ✖
5 Final report preparation ✖ ✖ ✖
1.3 Core Analysis Plan 5

Project management is very important in a core analysis job. Various experts


from different disciplines are involved in the project. Coordinating a huge team
including coring, transfer, geology, RCAL (routine core analysis), SCAL (special
core analysis), and other personnel is a complicated task. The project manager and
controller are responsible for observing the time and comparing the planned and
actual progress of the core analysis project. Any delay should be monitored care-
fully and compensated with appropriate arrangements. In a geological core analysis
section, the main milestone is thin section petrography. Many other analyses
depend on this task. The average time for study of 20 thin sections is one day for a
two-person team. Thus 100 thin sections are studied in one working week. Using
two teams is not recommended unless they have completely matched before. If time
is really a concern, dividing the study into facies and diagenesis parts is an
appropriate selection.

1.4 Coring

The coring interval is defined in the forecast program of the well and confirmed by
the coring plan team. Drilling is stopped and the string is pulled out of the well. The
coring head is attached to the end of the drill string and it runs in the hole. The
cutters of the head are made of synthetic diamond or tungsten. The bit cuts the rocks
and the cores pass through the hollow center of the bit and enter the core barrel. The
core head is selected based on the rock properties subject to drilling. Routine coring
systems consist of inner and outer core barrels. The outer core barrel is attached to
the drill string. The coring bit is placed at the bottom of the outer barrel. A core
catcher at the bottom of the inner core barrel fixes the core sample and prevents it
from moving out of the barrel. Drilling fluid is pumped through the annulus
between the inner and outer core barrels. Conventional coring tools cut cores with a
diameter between 5 cm (2 in.) and 15 cm (6 in.). Each core barrel is 9 m long and
the length of about a 27 m (90 ft) barrel for one run is common (three barrels). The
cores are cut in 1 m lengths for easy transfer to the laboratory.
The sleeves (barrels) are fiberglass in most cases (Fig. 1.2a, b). The inner smooth
surface eases core entry. Fiberglass is inert and the mud fluid has no considerable
effect on it even during core storage. Aluminum barrels are also used in some cases
(Fig. 1.2c). Fiberglass is lighter than aluminum but it has temperature (about 150 °C)
and pressure limitations and is less resistant to core jamming during the coring
process (McPhee et al. 2015). Some sleeves can be opened longitudinally and allow
core visualization and sampling at the wellsite. These sleeves have a hole that lets the
drilling fluid out and reduces the hydraulic pressure inside the barrel. Each barrel has
two longitudinal parallel lines with two different colors that show the top and bottom
of the core (Fig. 1.2b).
The natural fluid of the rock changes by mud filtrate invasion into the core
sample, gas expansion and expulsion, and fluid vaporization during the coring
process in most cases. The last is negligible if the cores are handled and preserved
6 1 Core Analysis: An Introduction

Fig. 1.2 Fiberglass (a, b) and aluminum sleeves (c)

carefully and the cores are analyzed in a reasonable time after coring (maximum of
six months). The amount of invasion depends on many parameters including coring
method, porosity and permeability of the sample, filtrate and reservoir fluid vis-
cosity, and pressure difference between the mud and net formation pressure. The
water-based mud filtrate also reacts with sensitive clay minerals. New systems such
as sponge coring try to reduce the fluid invasion into the core sample and so provide
more accurate water and oil saturation measurements. Anyway, the direct saturation
measurement on the core samples (Dean–Stark test) needs special attention to the
coring situations. Gas is expelled from the oil immediately after pressure release in
the coring or core opening process. For gas reservoirs, all of the hydrocarbon fluids
are expelled before any analysis. Water saturation can still be measured if other
conditions are ideal.
1.5 Core Preservation and Transfer 7

1.5 Core Preservation and Transfer

During the core handling and preservation process, any physical alteration of the
rock material should be minimized. As mentioned, cores are cut into one-meter
sleeves and transported in wooden boxes. Preserving the cores with wax is more
effective but this is only routinely done for some sensitive parts of the core or when
cores are not to be tested for a long time after the coring process. For example, some
reservoir parts of the cores are selected at the wellsite for SCAL tests. These parts
are wrapped in high-quality, nonreactive plastic films and aluminum foil and
then sunk in molten wax. Cores are extracted from the wax at the time of the
planned test.
It is recommended to analyze the core just after coring. Fluids react with the
barrel and evaporation changes the natural properties of the fluids.

1.6 Retrieved Data

Cores are the main source of information in any reservoir evaluation and charac-
terization project. Significant data from various disciplines are obtained from this
source.
• Geological evaluations include:
– Lithology. The mineral constituents of the rocks and their exact amount
using petrographical studies.
– Depositional environments. Help to reconstruct the geometry of the reservoir
body and determine the possible extent of each facies association. The type
of sedimentary environment and sub-environments determines the facies
distribution pattern in a 3D reservoir model.
– Absolute age dating and chronological sequence establishment. Using fossil
records.
– Regional scale correlation. Using fossils, geochemical proxies, and sedi-
mentological properties.
– Diagenesis. The processes that have affected the rocks after deposition. They
have a major role on reservoir properties in many cases.
– Fracture analysis. These studies have some limitations on cores but still
valuable information can be retrieved.
– Pore typing. Using petrographical studies.
– Geochemistry. Both organic and inorganic geochemistry help in more
accurate and effective analysis and interpretation of the source and reservoir
rocks.
– Geological rock typing. The basis of the heterogeneity reduction in the
reservoir.
8 1 Core Analysis: An Introduction

• Reservoir engineering evaluation:


– Porosity determination
– Permeability measurement
– Reservoir rock typing and hydraulic flow unit determination
– Oil–water or gas–water contacts
– Fluid saturation
– Acoustic velocity
– Gamma radiation
– Calibration of wire line logs using engineering data retrieved from the cores
(such as porosity, acoustic velocity, or gamma radiation)
– Grain density
– Electrical properties
– Wettability
– Relative permeability
– Capillary pressure
– Pore volume compressibility
• Geomechanical properties:
– Compressive strength
– Young’s modulus
– Poisson ratio
– Hardness

1.7 Routine Core Analysis

Routine core analysis or conventional core analysis (CCAL) is the measurement of


basic petrophysical properties of plug samples. The process involves measuring
static parameters with a single-phase fluid sample. Core analysis routinely starts with
starting RCAL and RCAL starts with core gamma logging. Whole core CT-scanning
would be part of the project if it is scheduled in the core analysis plan. Core scanning
could be before or after the gamma logging. The project continues with the depth
matching, core layout, Dean–Stark sample selection and water saturation measure-
ment, core cleaning, routine sample selection, core plugging, Soxhlet extraction,
drying, porosity and permeability measurements, slabbing, photography and imag-
ing, and resination. The sample selection and preparation steps as well as
CT-scanning and gamma logging are considered in Chap. 2. The remaining parts
including porosity and permeability measurements are considered here.
Porosity is defined as the ratio of space available for fluid storage to the bulk
volume of the sample. The bulk volume is routinely determined by mercury using
Archimedes’ law or by geometric calculations. Mercury is a nonwetting phase and
does not wet the surface of the sample. Thus it is completely separate from the
sample after the experiment. Bulk volume measurement using mercury
1.7 Routine Core Analysis 9

Fig. 1.3 Schematic design of a standard porosimeter

displacement is not appropriate for samples with vuggy porosity because mercury
remains in some spaces of these samples. Mercury is a toxic material and therefore
the geometric method is preferred in most cases. The diameter and height of the
plug is measured at three to five points and the average value is used to calculate the
bulk volume. Porosity is measured using Boyle’s law with a porosimeter.
A standard porosimeter consists of a reference chamber, a cell, or sample chamber
and pipes. Two gauges measure the pressure at various stages of the test (Fig. 1.3).
Porosity is measured in the following steps.
1. Helium gas is expanded from a storage capsule into a reference cell of known
volume (V1). Valve 1 is open and valve 2 is closed; thus the gas fills the
reference cell and connector pipes. The valve is closed and the system is dis-
connected from the source. The pressure recorded on gauge 1 is P1. The volume
of gas is V1. This is a known value because the volumes of the reference
chamber and connecting pipes are specified.
2. Valve 2 is opened. The gas expands from the reference cell to the sample cell
and occupies the cell and pipes minus the grain volume. The pressure is
recorded on gauge 2 (P2). The plug sample in the cell reduces the space by grain
volume.
According to Boyle’s law:

P1 V1 ¼ P2 V2 ð1:1Þ

This means that the initial pressure multiplied by the initial volume is equal to
the second pressure multiplied by the total volume of the cells and pipes minus the
space occupied by the solid part of the plug (Vg). Knowing the total volume,
10 1 Core Analysis: An Introduction

including the chambers and pipes, the Vg is calculated. The total volume of the rock
is Vb and therefore:

Vp ¼ Vb  Vg ð1:2Þ

If the Vp is known, the porosity is:

PorosityðUÞ ¼ Vp =Vb ð1:3Þ

Grain density is obtained by dividing the weight of the plug by its grain volume.
Grain density data are valuable for lithology determination and wire line log data
calibration. The important point is that the porosity is included in the bulk density
wire line tool whereas only grain density is involved in photoelectric factor mea-
surements in the well. In these definitions, matrix and grain have the same meaning.
They form the solid part of the rock.
Permeability is defined as the ability of a porous rock to transfer fluids. Only
absolute permeability (permeability of a single-phase fluid) is determined in the
RCAL section of the core analysis project. This variable is measured by passing a
fluid through a porous rock according to Darcy’s law (Eq. 1.4).

K ¼ QLl=ADP ð1:4Þ

where K is permeability (Darcy), Q is the flow rate (cm3/s), L is length (cm), l is


fluid viscosity (cP), A is the cross-sectional area (cm2), and DP is the fluid pressure
gradient between the input and output points. The SI unit for permeability is m2 but
Darcy is more convenient in reservoir studies.
A rock sample with a permeability of 1 Darcy, named after Henry Darcy, permits
a fluid with viscosity of 1 cP (1 mPa s) flow through 1 cm3/s under a pressure
gradient of 1 atm/cm in an area of 1 cm2. It is obvious that 1 Darcy is a very high
permeability compared to most reservoir rocks. Therefore the milidarcy (mD) unit is
widely used in the petroleum industry. The range of permeability in most reservoirs
is between 0.1 and 100 mD. More or fewer permeabilities are also possible, but
values lower than 0.1 mD are not sufficient to produce crude oil economically. The
cutoff value of 0.01 mD is also applicable for gas reservoirs. Permeabilities lower
than 0.01 mD could not carry any fluid for production purposes.
The routine part of a core analysis project ends with the SCAL sample selection
based on integrated geological and RCAL rock typing (see Chap. 6).

1.8 Special Core Analysis

Special core analyses are advanced tests on core plugs mainly dealing with fluid
flow and conductivity of more than one fluid. They are expensive tests needing
weeks or months of advanced laboratory measurements and therefore the samples
1.8 Special Core Analysis 11

must be chosen very carefully to obtain the most beneficial data from the reservoir.
It is recommended to select the samples after the geological and RCAL studies and
measurements. The basic units for SCAL sample selection are rock types (see
Chap. 6). All selected samples should be CT-scanned for considering any damages,
fractures, stylolites, or other barriers or carriers in the plug. These tests include but
are not limited to:
• Mercury injection capillary pressure tests (MICP)
• Relative permeability of two- or three-phase fluids
• Wettability
• Reservoir condition petrophysical properties
• Improved oil recovery (IOR, EOR) studies
• Determination of Archie exponents: a, m, n
• NMR core analysis
• Pore volume compressibility
• Formation damage effects

1.9 Wire Line Log Evaluation

Wire line logs are the most available data in reservoir studies. They are accessible
from almost all wells and reservoir intervals. Actually, extending and developing
the core data into the 3D interwell space is based on their comparison with logs.
There are various types of wire line logs but the conventional set of logs includes
natural gamma radiation (GR), neutron porosity (NPHI), bulk density (RHOB),
electrical resistivity (R), and sonic velocity (DT). Data are recorded in LAS format.
This format uses the ASCII (American Standard Code for Information Interchange)
codes for recording the data. The standard depth increment for wire line logs is
0.1524 m or 15 cm. As the plugging is routinely every 30 cm, core and log data do
not have exactly the same depths in the same interval. This fact should be con-
sidered when these data are compared with each other. Upscaling the data solves
the problem in most cases. A simple linear interpolation also is a good solution in
many cases. When the changes in logs and core data are plotted against depth and
compared to each other, actually the software is interpolating both of them to draw
a line of changes.
Wire line log evaluation before core opening is very useful. The log data are
analyzed by two methods including probabilistic and deterministic petrophysics
(see Kennedy 2015). Both methods are applicable but the probabilistic approach
uses all data types and calculates the effect of each variable on the final result.
Therefore the latter is preferred. The results are very useful for understanding core
and fluid properties before core opening (Fig. 1.4). Net pays are almost clear on
logs, the contacts could be defined, and an insight to the core properties is provided.
Defining electrofacies and correlating them with final rock types is one of the main
parts of reservoir studies (see Chap. 6).
12 1 Core Analysis: An Introduction

Fig. 1.4 An example of wire line log data analysis in a Permian–Triassic age gas reservoir
(Kangan and Dalan formations) in the Persian Gulf. Lithology, porosity, and saturations are
calculated before core opening (Courtesy of M. Nazemi)

1.10 Core Storage

After completing all analyses, cores must be archived. Rocks are not very sensitive
to ambient temperature and humidity. The main part of the fluids is vaporized or
cleaned before storage. There are no strict conditions for core maintenance. This is
different for Holocene sediment cores in which the environmental conditions are
very important for their repository. About the reservoir cores, the subject of this
book, the important problems are preventing mildew formation and physical
damage during core storage. An accurate system is necessary to find the desired
cores in a repository easily. There is a fully automated system in many repositories
that can find and move the cores to the study site. A working table suitable for core
viewing and description is recommended. Cores must be available for any future
1.10 Core Storage 13

Fig. 1.5 Core storage frames for the whole core (a, b) and thin sections (c)

work. Strong and secure racks are very useful. As with any other repository, health
and safety problems must be considered.
The core-derived materials such as plugs, trims, and thin sections are also col-
lected and stored after project finalization. Thus a core repository must have the
appropriate space for plugs, trims, thin sections, and also cuttings (Fig. 1.5).

References

Anderson G (1975) Coring and core analysis handbook. Petroleum Publication Company, USA
Kennedy M (2015) Practical petrophysics. Elsevier, Netherlands
McPhee C, Reed J, Zubizarreta I (2015) Core analysis: a best practice guide. Elsevier, United
Kingdom
Tiab D, Donaldson EC (2015) Petrophysics theory and practice of measuring reservoir rock and
fluid transport properties. Gulf Professional Publishing, USA
Chapter 2
Preparing for Analysis

Abstract Routine core analysis starts before geological section. This analysis
includes core gamma logging, depth matching, CT-scanning, saturation determi-
nation, core cleaning, marking, plugging, and trimming the plugs. Natural gamma
radiation is measured from the cores for depth matching. This is a very important
stage for core–log data comparison and correlation. The depth should also be
matched between the various runs of the wire line logs. In the next step, for virtual
viewing of fractures before opening, CT-scan images are prepared from the cores.
The CT-scan images can also be used for evaluation of some petrophysical
parameters. Water saturation of some plugs is measured. Special attention should be
given because water-based mud can change the water saturation of the samples.
Cores are cleaned for the first stage of macroscopic description and the marking
process. Routinely, three horizontal and one vertical plug are prepared from each
meter of core. This could be changed according to the project purpose and the
number of planned tests. Plugs are cleaned by the Soxhlet extraction method. Two
sides of each plug are trimmed. The cleaning process is also applied to the trims in
the case of heavy oil content. These trims are used for thin section preparation and
starting the geological section of the study. In some cases, plugs are prepared using
the sidewall coring method. After sidewall sampling, the other processes are the
same. Sidewall coring is suitable in some exceptional cases.

2.1 Core Gamma Logging

Drilling a well is a multidisciplinary work; each section measures the depth indi-
vidually by its specific tool. Because of using different tools for these measure-
ments, different depths are provided for a distinct point in the well. For example, the
logger depth for the bottom hole may be 2010 m whereas this is 2011 m for the
driller. As all data should be integrated in the reservoir studies, the same depth must
be considered for a specific point. In reservoir modeling, for example, a building
block (a cell) has a unique porosity value that is gained based on wire line log and
core porosity data integration. This is more vital for very thin reservoir layers, such

© The Author(s) 2018 15


V. Tavakoli, Geological Core Analysis, SpringerBriefs in Petroleum
Geoscience & Engineering, https://doi.org/10.1007/978-3-319-78027-6_2
16 2 Preparing for Analysis

as Kazhdumi (equivalent to Nahr Umr) sandstones in the Persian Gulf Basin. For
matching all data in a well, a reference depth is needed, in order that all other
downhole information can be shifted accordingly. This reference is the depth of
natural gamma ray (GR) data. Almost all tools working in a well have a GR
detector that measures the natural GR of the formation. The natural gamma radi-
ation of the reservoir rocks is the result of thorium, potassium, and uranium isotope
decay. These radiations are recorded by a detector and calibrate to an API unit.
The API unit varies from 0 to 200 for clean to completely shaly formations,
respectively. The GR of the cores is also determined before opening. The basic
parts of a core gamma logger are a GR detector and a conveyor belt, both mounted
on a base (Fig. 2.1). The tool is calibrated before any job. The standard Th, U, and
K containers with certain gamma radiation are used for this purpose. As the cores
are transferred to the laboratory in one meter length, they are arranged in order of
depth before starting the task. After starting the machine, the detector records the
background radiation and then the conveyor belt pulls the samples through the
detector. The detector measures the quantity of each radioactive element (Th, K,
and U) based on their different energy levels. These quantities as well as the total
GR count are recorded as a function of depth by the appropriate software connected
to the detector. Most of the core gamma loggers record the results based on count
per second (CPS) of the radiated GR. It is obvious that there is linear relationship
between this count and the natural gamma radiation of the cores. The amount of
counted GRs depends on the radioactive elemental concentration, the volume of
rock that is scanned, the distance between the source and the detector, and the
density of the rock (Ellis and Singer 2007). The depth of investigation for a wire

Fig. 2.1 A core gamma logger and one meter of core on the conveyor belt
2.1 Core Gamma Logging 17

line GR log is about 25 cm (Kennedy 2015). This means that the detector in a well
measures GR from a spherical volume with radius of 25 cm. Thus a core with lower
volume (a cylinder volume with about 5 cm radius) routinely has lower GR
quantities. Ignoring intralayer heterogeneity, both of them measure the GR of the
same rocks and have the same GR trend (peaks and troughs). Other factors
influencing the gamma ray counts are reservoir and environmental conditions such
as fluids, drilling mud, and invasion.
In general, there is a linear mathematical relationship between two points of core
and well GR data and all others could be transferred based on this
relationship. A linear regression line based on all datapoints is a better solution for
converting core to log measurements. Such a transformation is known as horizontal
shifting or log normalization. This horizontal shifting is not necessary in most cases
because the purpose is just marking the variations and matching the depths of cores
and wire line log data. This is only necessary when the user wants to calculate a
petrophysical property using these two measurements.
Because of different measurement conditions such as investigated rock volume,
mud filtrate invasion, and formation pressure and temperature, the radiation is not
the same for core samples and the borehole wall. In fact, the conditions are not the
same even for different runs in one well. Therefore the values are never exactly the
same. Figure 2.2 shows three repeated GR measurements of one core compared
with the corresponding wire line GR data of the well. As seen, the major trends are
similar but the values are not exactly the same.

2.2 Core–Log Depth Matching

After core GR logging, the GR values of cores and well logs are illustrated beside
each other and depth matching is performed core by core. The values of each
element or the total gamma could be used for this purpose. The total gamma is used
in most cases because this measurement includes the variations of all elements. The
well log data are a continuous record from all intervals or at least from the reservoir
parts, but cores are taken from some reservoir intervals and are discontinuous in
some cases. Therefore the depth of well log GR is routinely considered as the base
reference and the core depths are shifted, based on the comparison of major peaks
and troughs. The sampling interval of the core and well log GR may be different
(Fig. 2.2). This is not a major problem in most cases, as the main variations are the
same. In the depth-matching process, a key bed such as a thin layer of shale could
be a useful guide (Fig. 2.3). When the user shifts a point, all other points are also
shifted. Once the first point is shifted, the others are routinely placed almost in the
correct position but the process continues until all major variations are exactly
matched. This is a time-consuming task that often needs several re-evaluations, but
is certainly necessary.
18 2 Preparing for Analysis

Fig. 2.2 Three core gamma measurements from a carbonate Permian core sample of the Persian
Gulf Basin (left) and the corresponding wire line data (right)

2.3 Core CT-Scanning

The whole cores are routinely scanned before the core gamma logging. Selected
plugs for special core analysis (SCAL) or rock mechanical tests also scan for
detection of any fracture prior to test. Specialized core scanners are used in many
cases but hospital scanners are also functional. In a routine process, many
cross-sectional slices are prepared using an X-ray scanner. The number of slices
depends on the tool resolution, purpose, and budget of the project but it is more
than a routine human body scan. The result of a CT-scan illustrated by the grayscale
spectrum varies from pure white to pure black. Denser material adsorbs more
X-rays and so appears white. Porosities are black to dark gray according to their
fluids. Note that the hydrocarbon gases are expelled from the cores even before
opening. Cement-filled spaces are mostly bright because they are denser than the
surrounding matrix (anhydrite cements, e.g.). Fractures are black if they contain
fluids, and white if they have been filled by cement during diagenetic processes.
The images are also used to evaluate the core heterogeneity. Special software can
integrate all slices and produce a 3D image of the core. The main purpose of
CT-scanning is recognizing any fractures prior to the core opening (Fig. 2.4).
2.3 Core CT-Scanning 19

Fig. 2.3 GR depth-matching log is composed of original and shifted core logs compared with
wire line log data. The table of shifts is also included

Nowadays, high-resolution CT images are used for calculating petrophysical


properties of the samples. These images form the basis of a new developing
analysis called digital petrophysics.

2.4 Core Opening and Layout

After core gamma logging and CT-scanning, the sleeve cap is removed and the core
pushed out from the sleeve. Caps are made from plastic and are bound to the top
and bottom of the core using stainless-steel clamps (Fig. 2.5a). Routinely, cores are
easily pushed out by slightly tilting the core barrel. Hitting the barrel with a plastic
20 2 Preparing for Analysis

Fig. 2.4 An example of anhydrite-filled fractures on CT-scan images. Cross-sectional slices (a),
the longitudinal whole core image (b), and slabbed core photo (c) are also illustrated

hammer is also useful. The hammer must be used with care because it can induce
fractures in the core sample. If it is not possible, the barrel is cut out and the core is
brought out. This is done by fixed or portable power saws. After opening 1 m of
core, the Dean–Stark samples are taken immediately and preserved for later satu-
ration measurement (see Sect. 2.4). The core is then laid out on the appropriate
tables. These tables have some gutter or pyramid shape positions to prevent core
rotation. The tables are normally made of steel and thus putting a half UPVC pipe
with the same core size below the core is recommended to prevent any core damage
(Fig. 2.5b). After opening, two parallel longitudinal lines with different colors are
drawn on the cores using waterproof markers. They show the top and bottom and
prevent core rotation during the plugging process (Fig. 2.5c).
2.5 Dean–Stark Extraction 21

Fig. 2.5 The layout process. Opening the caps (a), laying out the cores on the table (b), and
drawing two parallel lines on the cores (c)

2.5 Dean–Stark Extraction

Dean–Stark extraction is the first evaluation of the water saturation (Sw) of the
samples. A plug is prepared just after opening the core sleeve using an oil-based
lubricating fluid and then is placed in a distillation extraction system (Fig. 2.6). The
boiling solvent (toluene in most cases) vaporizes the water content of the sample.
Both solvent and water condense and collect in a graded tube. The water is denser
and is collected at the bottom of the container and the solvent overflows and returns
to the flask. The process continues until no more water is collected. This takes about
two days depending on core porosity, permeability, and hydrocarbon viscosity. The
water volume is recorded. The plug is cleaned using the Soxhlet extraction method
and the pore volume is measured using Boyle’s law (see Sect. 1.7). After that, the
water saturation of the sample is calculated by dividing the water volume to the
pore volume of the sample. It should be noted that the water-based mud routinely
filtrates some water to the core and thus cores that are drilled with oil-based mud are
preferred for this test.
22 2 Preparing for Analysis

Fig. 2.6 Dean–Stark water saturation measurement apparatus

2.6 Core Cleaning

After core opening and Dean–Stark sampling, the cores are cleaned. This is dif-
ferent from plug cleaning using Soxhlet extraction. The cores are cleaned with
water if there are no water-sensitive minerals in the cores such as clays or halite. If
there are, they are cleaned with only a nylon brush. This process is ignored for the
unconsolidated cores as they disaggregate during this process. This is also not
possible for oil reservoirs, especially cores containing heavy oil. The cleaning
process removes the remaining drilling mud from the core surface (Fig. 2.7).
A preliminary core description and fracture analysis is done after the cleaning
process. The core cleaning process also prepares the cores for the marking stage. It
is possible to cover some selected parts of the core with nylon and aluminum foil
for sensitive tests such as geochemical analysis (Fig. 2.7).

2.7 Marking

After cleaning, the plug locations are determined on the cores. A petrophysical
evaluation of the wire line logs before the marking process is very useful because
the marking policy depends on the reservoir properties of the rocks in most cases.
Routinely, three horizontal and one vertical plug are prepared from each meter of
core. Thus three horizontal plugs are prepared every meter. There is no strict rule
for selecting the location of a vertical plug on 1 m of core. It is recommended to
2.7 Marking 23

Fig. 2.7 Core cleaning process using water and a nylon brush for consolidated cores. Some parts
could be covered for sensitive tests (the T marked part at the center of image)

prepare it wherever there is a horizontal fracture in the core and therefore there is no
need for further core cutting. Each plug type is marked with a special symbol using
a waterproof marker (Fig. 2.8a). Additional plugs such as plugs needed for rock
mechanical or special tests also are marked with different symbols or colors. The
project’s geologist, reservoir engineer, and petrophysicist as well as lab operator
supervise the core marking process.
One of the most important aspects of core marking is how the team looks at the
reservoir and nonreservoir parts and the heterogeneity of the rocks. It is possible to
ignore the nonreservoir intervals or reduce the plug frequency in these parts but the
important point is that such changes should be considered in all aspects of future
studies. As many experts work with the numerical values, ignoring the nonreservoir
sections may cause misunderstanding of the reservoir behavior. If the team ignores
a 2-m thick anhydrite layer, it must be considered in reservoir modeling. This is also
the same for some intervals with no plugs, such as shale layers. In most cases, it is
not possible to prepare the plugs from shaly intervals because they disaggregate in
the plugging process. Anyway, they have a major role in reservoir performance and
compartmentalization. In the core description process all macroscopic properties are
recorded. They should be integrated with any large-scale reservoir studies.
24 2 Preparing for Analysis

Fig. 2.8 The process of marking (a) and plugging (b)

Another important point in the marking process is the location of contacts.


Below the oil–water or gas–water contact, the rocks have less interest in reservoir
studies. Below these contacts, the Sw is 1 and thus no hydrocarbon will flow.
Therefore, the plug frequency is reduced. For example, one plug for each meter of
core is an appropriate selection.

2.8 Plugging and Trimming

Cores are prepared from vertical wells in most cases. Horizontal plugs are prepared
perpendicular to the core length; they are horizontal if you imagine them in their
original position in the reservoir. Vertical plugs are prepared parallel to the core axis
(Fig. 2.8b). In cores with high angle layering, horizontal plugs are prepared parallel
to layering. In a homogeneous rock, porosity is a scalar intrinsic property of the
rock and is not direction-dependent. In contrast, permeability is a vector property
and depends on the flow direction. Absolute permeability is also an intrinsic
property of the rock and does not depend on fluid properties. It is measured in 100%
2.8 Plugging and Trimming 25

saturation of a single-phase fluid. Therefore the direction of the plug has no effect
on porosity or geological properties, but permeabilities are different in various
directions. Routinely, the horizontal permeability is more than the vertical because
the grains are deposited perpendicular to their maximum projection area. The
process of plugging is the same for all plugs except the Dean–Stark samples that are
prepared using an oil-based cooler fluid. The plugs are 2.5–5 cm in diameter and
5–10 cm in length. They are prepared based on job priority. It means that the plugs
of the most important task are taken first. If, for example, the rock mechanical
studies are more important for the reservoir, their plugs are prepared first. After their
plugging, the routine plugging for porosity and permeability studies starts. Double
plugs are taken as close as possible if there is any need for nearly the same plugs for
various tests (such as overburden and relative permeability). A cylindrical drill cuts
the sample using a cooler fluid, mostly water. If there is any water-sensitive mineral
in the rock, an oil-based cooling fluid is used.
After the plugging process, a cylindrical sample with rough up and down sur-
faces is available. In the trimming process, these two surfaces are cut and two chips
of rocks are available (Fig. 2.9). They are put in a bag with their identity infor-
mation and sent for thin section preparation. More trims could be cut if additional
ones are needed. For oil formations, especially heavy oil reservoirs, trims are also
cleaned using Soxhlet extraction. The process of cleaning is exactly the same as for
plugs.

Fig. 2.9 Plugs are used for other petrophysical and reservoir tests whereas trims are used for thin
section preparation and geological studies
26 2 Preparing for Analysis

2.9 Soxhlet Extraction

The Soxhlet extraction method is used for cleaning plugs in a core analysis job
(Fig. 2.10). The plug contains hydrocarbons and salt water; both must be removed
before other analyses such as porosity or permeability measurements. The trims are
also cleaned if the samples contain crude oil, especially heavy oil. A Soxhlet
extraction assembly is composed of a heater, flask that contains the solvent
(methanol or toluene), side arm, siphon arm, sample chamber, and condenser
system. The vaporized solvent travels up the side arm, condensates, and falls down
to the chamber. The distilled liquid solves the plug contaminations. The process
continues until the level of solvent reaches the siphon point. Then all the liquid
containing the solvent, crude oil, and salts refluxes from the chamber, returning to
the flask. With this method, the solvent recycles each time of refluxing and the
distilled clean liquid is in contact with the plugs. The process starts with toluene to
remove the crude oil from the plugs. The cleaning time depends on the sample
properties (such as porosity and permeability) and the oil density. The process may
take about one month for plug samples containing heavy oil. Samples can be
checked with UV light to see any remaining contamination. After cleaning with
toluene, the Soxhlet cleaning continues with methanol to remove the remaining
salts. Samples are dried using an oven. The plugs are ready for other routine core
analysis (RCAL) tests and trims are now suitable for thin section preparation. As
mentioned previously, this process is not necessary for trims of a gas reservoir.
There are also some other methods for plug cleaning (see McPhee et al. 2015).

Fig. 2.10 Schematic illustration of Soxhlet extraction system for cleaning the plug (or trim)
samples
2.10 Sidewall Coring 27

2.10 Sidewall Coring

Sidewall is not really a coring method. It is more like a plugging method from the
borehole wall. A percussion or rotary sidewall corer is placed against the interval of
interest using a wire line tool. In the percussion method, hollow tubes with sharp
edges are shot into the wellbore wall. The tube is forced into the formation using
explosive charges and a cylindrical sample, the same as a plug, is picked up from
the formation. It is obvious that this method is not applicable to tight consolidated
formations such as tight carbonates. The high pressure used for tube penetration
changes the texture of the unconsolidated form and fractures the harder rock
samples. Therefore the percussion method has been almost completely replaced by
rotary sidewall coring. In this method, a hollow tube rotates against the borehole
wall. A bit that has been assigned to the head of the tube cuts the sample and
collects it in the tool. The tool is taken back to the surface. The number of samples
collected in each run depends on the tool. It could be between 20 and 60 plugs. The
plugging distance is variable based on the project need. The old sidewall plugs were
smaller than the standard sizes used for routine reservoir studies but today the
standard plugs are prepared by various companies. Many plugs are deformed in
sidewall coring by both percussion and rotary methods.
Sidewall coring is recommended when the routine coring process with full
recovery is not possible or the reservoir has enough data. In the latter case, just a
special test or data from an interval are necessary. For example, if everything is
known about a reservoir from previously cored and logged wells and there is just an
unconventional mud lost in a new well, a sidewall coring is necessary. Sidewall
coring is also useful for soft formations such as loose sands.
Sidewall coring is cheaper and faster than the routine coring method, but there
are many disadvantages. At first, the full recovery of the reservoir is not available.
This is a main problem, as many fluid behaviors in the reservoir depend on some
specific layers. The mud pressures used in drilling and filtrate invasion both change
the nature of the sample. These samples are not appropriate for reservoir mechanical
tests. They are not even eligible for porosity and permeability measurements. The
fluid content also has been changed during the coring process. However, they are
still suitable for geological studies.

References

Ellis D, Singer JM (2007) Well logging for earth scientists. Springer, Amsterdam
Kennedy M (2015) Practical petrophysics. Elsevier, Amsterdam
McPhee C, Reed J, Zubizarreta I (2015) Core analysis: a best practice guide. Elsevier, United
Kingdom
Chapter 3
Microscopic Studies

Abstract Microscopic observations are one of the main sources of information for
geological studies. This is more important in core analysis with limited macroscopic
samples. Routine microscopic studies of a core sample include petrographical
analysis to understand facies properties and diagenetic processes, paleontological
studies for absolute age dating, X-ray diffraction for mineral identification (espe-
cially clays), scanning electron microscopy equipped with energy dispersive
spectroscopy for pore and pore throat determination, mineral identification, and
elemental analysis. The static reservoir properties of a sample depend completely
on primary (facies) or secondary (diagenesis) characteristics of the rocks. The rock
mineralogy, constituents, sedimentary environment, microscopic porosities,
cements, compaction features, and many other parameters are gained by study of a
rock sample under a polarizing microscope. They are recorded on standard sheets
and compared with other rock properties derived from routine or special core
analysis sections. Paleontological studies are used for absolute age dating and make
it possible to correlate the strata in a chronostratigraphic framework. Such a
framework is integrated with other microscopic and macroscopic geological data
and provides the reservoir zonation scheme and understanding of reservoir geom-
etry. Fine-size minerals, especially clays, are identified by the X-ray method. They
play a vital role in reservoir properties and future drilling in the field. Pore types and
pore throats determine the fluid flow properties and major rock types of the
reservoir. They have a major effect on reservoir heterogeneity. Final results are
combined with other sources and characterize the geological role in micro and
regional scale distribution of reservoir properties.

3.1 Thin Section Preparation

Petrography is the basis of any geological core analysis task. After trimming the
plugs, trims are used for thin section preparation. If there is any oil staining,
especially in heavy oil fields, trim cleaning is necessary. Trims are cleaned with the
Soxhlet extraction system, the same as plug samples. There is no need for removing

© The Author(s) 2018 29


V. Tavakoli, Geological Core Analysis, SpringerBriefs in Petroleum
Geoscience & Engineering, https://doi.org/10.1007/978-3-319-78027-6_3
30 3 Microscopic Studies

drilling salts from the trims. They are routinely removed during thin section
preparation stages. It should be mentioned that thin sections could also be prepared
from anywhere in the core. Just a small sample is enough. The trims have at least
one flat surface that is attached to a glass slide using epoxy. If the sample is not a
trim, it has no flat and smooth surface and therefore a first cut is necessary. The
slide must be frosted before attaching the sample. This is to roughen the surface in
order for the epoxy to bind well. The other side is cut using a diamond saw
(Fig. 3.1a, b) and what remains is ground away until the desired thickness is
reached. This is a very important stage because there is a high risk of rock damage
in this step. Go slow and be careful. The user should check the thickness in short
intervals until light passes through the sample. This calls for a great deal of
experience. Most of the core samples are composed of carbonates or siliciclastic
minerals. Checking is more important when working with mud-dominated or mixed
siliciclastic–carbonate samples because they are more prone to damage. The
grinding process starts with 400-grit carborundum on a rotating steel disk
(Fig. 3.1c) and continues with the 600-grit carborundum on a glass plate
(Fig. 3.1d). The final optional stage is smoothing the surface with 1000-grit car-
borundum. Thin section slides are typically 25 mm  45 mm but larger ones are
also produced. They are routinely covered with another glass slide (Fig. 3.1e, f)
using epoxy again. The rock sample is about 30 lm (0.03 mm) thick and light
passes through the rock and glasses. Every mineral has its special properties in
normal or polarized lights. All visible parameters from the rock are recorded.

Fig. 3.1 Thin section preparation steps. Cutting the sample with a saw (a), automatic grinding
disk (b), manual disk (c), and final grinding on glass (d). Final result (e, f). Courtesy of
A. Rezazadeh
3.1 Thin Section Preparation 31

3.1.1 Thin Section Staining

Mineral staining is used for rapid and accurate identification of some common
minerals. New techniques such as scanning electron microscopy (SEM) or
cathodoluminescence can also display such information more reliably than staining,
but due to the availability and low cost of staining, it is still used for this purpose.
There are different techniques for staining various minerals. Carbonate staining is
used for distinguishing calcite from dolomite or aragonite. Mineral staining has a
long history. The most widely used methods for carbonates are a dilute hydrochloric
acid containing alizarin red S (ARS), potassium ferricyanide, or a mixture of both
(Dickson 1965; Evamy 1963). Alizarin is a plant root-derived chemical compound
that produces a red stain on hydrochloric acid-reacted carbonates. The calcite and
aragonite stain red but dolomite and other minerals remain unstained. The reaction
time is very short, from 10 s to about 3 min depending on the acid concentration and
the rate of reaction between the sample and acid solution. Use of 8–10 cc HCL plus
100 cc distilled water is recommended in almost all the literature (Friedman 1959;
Dickson 1965). The acid is cold, at room temperature, in all cases.
Potassium ferricyanide (PF) is used for distinguishing ferroan calcite and
dolomite. A pale to deep turquoise blue is produced in the reaction. Rhodochrosite
(MnCO3) stains pale brown. The treatment time depends on the reaction rate of the
carbonate with the acid solution. As dolomite reacts less vigorously than calcite, the
intensity of color is not indicative of the Fe content. There is a nonlinear rela-
tionship between iron content and color intensity in pure calcite (Reeder 1983).
Hydrochloric acid is used in both methods and thus the sample will react with
the solution. The final result is thinner than the original rock sample. Both ARS and
FP are useful in carbonate staining and a mixture containing these two materials is
recommended (Fig. 3.2).

Fig. 3.2 Staining of carbonate minerals by ARS, PF, and combination [data from Friedman
(1959) and Dickson (1965)]
32 3 Microscopic Studies

3.2 Quantitative Carbonate Petrography

Petrography is a description of the rocks aided by the microscopic studies of thin


sections. This task is fundamental to a geological description of the rocks, and
therefore it should be done very carefully with the help of a trained eye. User
experience is very important especially in complicated textures, overlapped dia-
genetic processes, and optical quantitative estimation of the rock constituents. The
parameters are clearer and easy to recognize in a qualified thin section. In the study
of heavy oil fields, Soxhlet extraction is recommended for the trims, as the oil is not
removed in the thin section preparation process and covers the surface of the thin
section. In reservoir geology, all parameters are compared to each other in order to
understand and interpret the reservoir properties in space and time. To reconstruct
such a distribution, the results should be as quantitative as possible. Qualitative
descriptions are converted to quantitative variations using a numerical scale. For
example, intraclast frequency in a rock could be expressed by none, rare, common,
abundant, and very abundant, qualitatively. They are converted to 0–4 in a quan-
titative description. As this is not exactly the number of intraclasts in the rock, the
word semiquantitative is preferable. The numerical frequency can be plotted in a
sedimentological log against depth and compared with the porosity or permeability
distribution. Digital recording is more applicable as these parameters could be
compared with other numerical data. Prepare a sheet containing the headers.
Separate the sheet into facies and diagenesis parts (Fig. 3.3). Slightly different
parameters are recorded in various projects but the main characters are constant.
These parameters include:
• General information of the sample. This information includes row number in the
database, core number, box number, plug number (if thin section is from the
plug), and depth.
• Lithology. Generally, all samples are carbonate but other constituents are also
present. Minor amounts of quartz and clays are present in most cases. Anhydrite
is one of the ingredients in a carbonate–evaporite reservoir. The main lithology
is marked by its abbreviation (e.g., L for limestone, D for dolomite, A for
anhydrite, etc.). The amount of various carbonate minerals is also important.
Calcite and dolomite are dominant in most cases but siderite also can be seen. It
is better to evaluate the percentage of each mineral individually. Visual com-
parison charts are suitable for this purpose (Fig. 3.4). The lithology determines
the behavior of reservoir properties in many cases. Wettability is one of the
parameters influenced by lithology. Porosity and permeability are controlled
partly by the amount of each mineral component in many cases (calcite and
dolomite, e.g.).
• Allochems. The framework of grain-dominated carbonates is composed of
allochems. They are also present in some mud-dominated samples with less
frequency. Bioclast, ooid, intraclast, pellet, peloid, and oncoid are the most
common allochems. Each represents a distinct environmental condition. Facies
classification is mostly based on the micrite frequency, which reflects the
3.2 Quantitative Carbonate Petrography 33

Fig. 3.3 An example of a quantitative carbonate petrography sheet. Parameters are interchange-
able based on the project objectives

depositional energy, and the amount of allochems. Point counting is a method to


quantify thin section components. A point counter is attached to a microscope
stage and moves the thin section with constant increments along a line each time
the petrographer pushes a button. Each button is reserved for one component
before starting the count (Fig. 3.5). At the end of the line, the operator moves
the slide to the beginning of the next line. A counter counts the number of
34 3 Microscopic Studies

Fig. 3.4 Visual comparison charts for estimating rock constituents under the microscope
[compiled from Bebou and Loucks (1984) and Scholle and Ulmer-Scholle (2006)]

Fig. 3.5 A point counter attached to a polarizing microscope (a) and point counting in
JMicroVision (v. 1.27) software (b). Some parts of the figure (b) changed graphically for better
view
3.2 Quantitative Carbonate Petrography 35

pushes of each button. Using the number of total points, the percentage of each
component is calculated. A total of 300–500 points is recommended. There is
also special software for the point counting based on photos that are prepared
from the slide. The overall process is the same. There are some limitations such
as rotating the stage and seeing the polarizing characters of the minerals in the
software. There are also some advantages such as automatic calculations and
showing the results with various charts and diagrams immediately. The results
of the point counting process are accurate enough to compare with the other
reservoir parameters.
• Sedimentary features. Many sedimentary features in addition to the allochems
and lithology components are used in geological interpretations. These features
are different from project to project and vary based on the research purpose.
Some of these features include bioturbation, opaque minerals, lamination, mud
crack, brecciation, and fenestral fabric of the rock.
• Facies name. In general, all sedimentary rock attributes could be used to define a
facies (Reading 1986). These characters include lithology (lithofacies), fossil
content (biofacies), or even the depositional current (turbidity facies). The
parameters of interest reflect the primary depositional conditions of the sample.
Thus any diagenetic process and feature is ignored before naming the facies. For
example, a pseudomatrix created by disaggregation of mud clasts in a sandstone
is not related to the primary depositional environment and is subtracted from the
overall matrix content of the rock for facies nomenclature. Dunham (1962)
classification modified by Embry and Klovan (1971) is the most employed
classification of carbonate facies (Fig. 3.6).

Fig. 3.6 Dunham (1962) classification of carbonates modified by Embry and Klovan (1971)
36 3 Microscopic Studies

This classification is well applicable in the petroleum industry because it is based


on rock texture which reflects the reservoir properties of the samples. The point is
that actually the microfacies are classifying but it is routine to use the word facies.
In carbonates, microfacies include all microscopic and macroscopic features of the
rock (Flugel 2010). After core description, the macroscopic features also are
attributed to the microscopic facies. The classification is based on the energy of the
depositional environment. Thus the frequency of micrite versus allochems is the
key factor for facies determination.
Classification based on rock texture is not adequately enough in most cases,
because allochems play an important role in interpretation of sedimentary envi-
ronments. The most frequent allochems (more than 10%) are also used in the name
of facies. For example, an ooid grainstone is a mud-free carbonate rock with ooids
as the most frequent allochem. The two most frequent allochems are also used in
order of frequency. For example, a bioclast peloid wackestone is a mud-dominated
carbonate sample and the main allochem is peloid but bioclasts are also present with
more than 10% frequency. Carbonates are very heterogeneous and this type of
classification yields a lot of names. Most of these rocks have been deposited in the
same depositional conditions. Bioclast wackestone and peloid wackestone both are
deposited in a lagoon of the ramp environment (Flugel 2010). Therefore they are
merged in one facies group, bioclasts/peloid wackestone. The problem may be
different for the intraclast and ooid grainstones that deposit in shoal and leeward
shoal environments, respectively. The purpose of facies determination is interpre-
tation of the primary depositional setting and this goal should be kept in mind in
facies merging to build facies groups. An example of facies merging in Permian–
Triassic carbonates of the Persian Gulf Basin is illustrated in Table 3.1.
Plotting the appropriate charts for illustrating the distribution of various facies in
a facies group is a good guide for further interpretations (Fig. 3.7). The final result
is used for defining facies associations that belong to one environment. For
example, all facies deposited in a ramp peritidal setting are integrated to build a
peritidal zone environment.
It is recommended to assign a code number to each facies. This code is used in
drawing charts, columns, and any comparison with other reservoir properties. They
are arranged based on related depositional environments, routinely from land to sea.
• Sedimentary environments. As mentioned previously, facies groups are merged
to build a facies association. Each association is related to a particular deposi-
tional setting. The facies and sedimentary environments are some of the most
useful data in sequence stratigraphy of a reservoir.
• The amount of porosity and pore types. The porosity value is determined with
visual estimation, point counting, or image analysis using various software. In
visual estimation, porosity value is determined using comparison charts for
visual estimation under the microscope. The results of porosity values derived
from the point counting are comparable to laboratory tests. It is obvious that the
visual and laboratory values are not exactly the same in most cases. Commonly,
visual estimates of porosity from thin sections are lower than the routine core
3.2 Quantitative Carbonate Petrography 37

Table 3.1 An example of facies merging in Permian–Triassic carbonates of the Persian Gulf
Basin
Facies code Facies groups Included Environment
F1 Anhydrite Anhydrite Supratidal
F2 Claystone Claystone Peritidal
F3 Mudstone often Mudstone/dolomudstone Peritidal
with evaporites
F4 Stromatolite Stromatolite boundstone Peritidal
boundstone
F5 Thrombolite Thrombolite boundstone Peritidal
boundstone
F6 Fossiliferous Fossiliferous mudstone/fossiliferous Peritidal/
mudstone dolomudstone Lagoon/
open marine
F7 Peloid, bioclast Peloid wackestone, bioclast wackestone, Lagoon
wackestone peloid bioclast wackestone, bioclast peloid
wackestone, intraclast bioclast wackestone,
intraclast wackestone, ooid wackestone,
ooid peloid wackestone, peloid ooid
wackestone, intraclast ooid wackestone,
intraclast peloid wackestone, oncoid
bioclast wackestone, oncoid ooid
wackestone, oncoid wackestone, peloid
intraclast wackestone, peloid oncoid
wackestone
F8 Oncoid/peloid Peloid packstone, peloid bioclast Lagoon
packstone packstone, bioclast peloid packstone,
bioclast oncoid packstone, intraclast oncoid
packstone, oncoid bioclast packstone,
oncoid intraclast packstone, oncoid
packstone, oncoid peloid packstone, peloid
oncoid packstone, oncoid ooid packstone
F9 Ooid, bioclast Bioclast packstone, bioclast ooid Leeward
packstone packstone, intraclast bioclast packstone, shoal
intraclast ooid packstone, ooid bioclast
packstone, ooid packstone, ooid peloid
packstone, peloid intraclast packstone,
peloid ooid packstone, intraclast ooid
packstone, intraclast packstone, oncoid
ooid packstone, ooid intraclast packstone,
ooid oncoid packstone
F10 Oncoid, bioclast/ Oncoid grainstone, oncoid bioclast Leeward
ooid grainstone grainstone, oncoid ooid grainstone, shoal
Bioclast oncoid grainstone, ooid oncoid
grainstone, peloid oncoid grainstone
F11 Peloid ooid/ Ooid peloid grainstone, peloid ooid Shoal
bioclast grainstone grainstone, bioclast peloid grainstone,
peloid bioclast grainstone, peloid
grainstone
(continued)
38 3 Microscopic Studies

Table 3.1 (continued)


Facies code Facies groups Included Environment
F12 Bioclast/ooid Ooid grainstone, ooid bioclast grainstone, Shoal
grainstone bioclast ooid grainstone, bioclast
grainstone
F13 Intraclast bioclast/ Intraclast grainstone, intraclast bioclast Seaward
ooid grainstone/ grainstone, bioclast intraclast grainstone, shoal
packstone bioclast intraclast packstone, intraclast ooid
grainstone, ooid intraclast grainstone, ooid
intraclast packstone, intraclast peloid
grainstone, intraclast peloid packstone,
peloid intraclast grainstone, intraclast
packstone, intraclast oncoid grainstone
F14 Crystalline Crystalline carbonate None
carbonate

Fig. 3.7 An example of facies distribution within a facies group (F7 of Table 3.1, Peloid, bioclast
wackestone). The main allochems in about 75% of facies are peloid and bioclasts. Other facies
have been deposited in the same sedimentary environment

analysis (RCAL) tests. This is because there are microporosities in the sample
that cannot be seen in optical microscope scale. These include microporosity
between the micrite particles, calcite cements, dolomite crystals, and also within
the micritic envelopes of the grains. There are also different methods for eval-
uating porosity from thin sections using automated image analysis techniques.
Image binarization is recommended to separate the matrix and porosity effec-
tively before analysis. Binarization is the process of dividing the whole pixels of
the image into two classes, usually black and white. The problem is complicated
3.2 Quantitative Carbonate Petrography 39

as minerals have various colors with different intensities under polarizing light.
Preparing thin sections with blue-dye stained resin facilitates porosity recogni-
tion. The one-side flat sample is impregnated with the mixture of blue dye and
epoxy resin using a vacuum chamber. The colored resins fill the porosities.
Pore types determine some petrophysical properties of the reservoir samples,
such as porosity–permeability relationships. Pore typing is also an appropriate
method for rock typing and grouping the samples to reduce reservoir heterogeneity
(e.g., Ahr 2008). Various types of macropores, pores that are visible under the light
microscope, are determined at this stage. The results are comparable with laboratory
measurements such as mercury injection capillary pressure (MICP) or nuclear
magnetic resonance (NMR) data.
• Fractures. The most important point is that fracture study based on thin sections
or even plugs is not an accurate method. Plugs and thin sections are prepared
from the nonfractured intervals and thus the fractures seen on thin sections or
plugs are not good indicators of the fractures in the reservoir. However, fracture
intensity (frequency), size, and filling are recorded. The fracturing mechanism is
also recorded, if possible (see Sect. 4.3).
• Cementation. Type and frequency of cements have major effects on reservoir
properties. In carbonates, isopachous, blocky, bladed, drusy, and anhydrite
cements are more common.
• Compaction features. Stylolites and solution seams represent chemical com-
paction. They have major effects on reservoir properties in some cases (e.g.,
Mehrabi et al. 2016). Concave–convex contacts, suture surfaces between the
grains, and microfracturing within the grains can be combined in a physical
compaction item. The compaction rate shows the integrated effect of com-
paction indicators.
• Dolomites and the dolomitization process. These change the porosity and per-
meability, their relationships, rock density, pore throat size distribution, and
wettability of the reservoir. Types and frequency of various dolomites are very
important. Dolomites are divided based on their size and shape. One of the early
classifications was made by Friedman (1965). Sibley and Gregg (1987) repre-
sented a more complete classification based on size and crystal boundary rela-
tionships. Various classifications could be used but generally dolomites are
divided into planner and nonplanner according to their shape, and sucrosic and
dolomicrite based on their size. Sucrosic dolomites are coarse-grained, usually
euhedral rhombs with crystals between 20 and 120 lm (Warren 2000). Smaller
dolomites (less than 20 lm) are dolomicrites. The rate of the dolomitization is
also recorded. Compromise boundaries with many crystal-face junctions rep-
resent the planar dolomites. Nonplanar dolomites are characterized by irregular
boundaries. Dolomites can also be classified based on their relationships with
the precursor fabric. If they follow the original fabric, they are fabric-retentive
(mimic) and if the depositional fabric is obscured by dolomitization,
fabric-destructive (nonmimic) dolomites are formed.
40 3 Microscopic Studies

There are also some other features that cannot be related to a group. For
example, neomorphism, micritization, micrite envelope, and anhydrite nodules
could be recorded. They are used for recognizing sedimentary environments, dia-
genetic processes, sequence stratigraphic boundaries, and interpreting reservoir
quality variations within a reservoir. Selecting samples for any other analysis such
as SEM, XRD, blue-dye thin section preparation, geochemical studies, or any other
tests are recorded in separate columns. A tick mark can be recorded for every
selected sample in its column. The geologist easily filters the data and sees the
various characteristics of the selected samples. This also results in saving time when
the analyst wants to know about the various analyses, number of samples, their
facies characteristics, and diagenetic processes. The laboratory tests of the plugs are
recorded beside the thin section parameters. Porosity, permeability, and grain
density are common parameters recorded in the thin section study sheet. This is
important because the next step is comparing laboratory results with the geological
characters. Recording all parameters in one database facilitates such comparisons.
Porosity and permeability distribution within every geological group (such as facies
or rock type) can be observed easily. For example, a simple filtering based on facies
code and drawing a scatterplot with porosity and permeability data constitute a first
look at the reservoir quality of that facies (see Sect. 6.3). It is more convenient to
record the results of various rock typing methods such as Lucia, Winland, FZI, and
Lorenz (see Chap. 6) beside the results of thin section studies. With this method,
the geological parameters of each rock type are easily recognizable and comparable
with the rock types resulting from the geological methods. There are always some
uncommon parameters in any reservoir. Therefore the last column is “Remarks.”
Properties that are not included in the database and also are not a routine charac-
teristic of the reservoir are recorded here. This column is used for better under-
standing and interpretation of other variations.

3.2.1 Quantitative Siliciclastic Petrography

Many parameters are the same in carbonate and siliciclastic petrography. Some
other parameters can be applied with some modifications (Fig. 3.8).
• General information and depth are exactly the same.
• Lithology is composed of three main components including quartz, feldspars,
and rock fragments (lithics). These three components represent the energy and
duration time of the transportation. Quartz is the most stable constituent. High
amounts of quartz are evidence of high energy and longtime transportation of
the sediments. Lithics are good representations of the provenance rocks. They
are the most unstable particles and easily disaggregate in a high-energy trans-
portation system. Feldspars are also sensitive to environmental situations. They
transform to other minerals, especially clays, in humid conditions. Such change
also depends on time. In coastal sediments with a long history of wave erosion
3.2 Quantitative Carbonate Petrography 41

Fig. 3.8 An example of a quantitative siliciclastic petrography sheet. Parameters are interchange-
able based on the project objectives

and deposition, the main component is quartz. In a meandering river with humid
environmental conditions, feldspars are altered into clays. In some cases, car-
bonate grains (mostly bioclasts) are also present. These grains are formed in a
depositional environment in most cases. Thus carbonate grains represent a
mixed siliciclastic–carbonate setting if they are present in considerable amounts.
Evaluating the amount of each component is exactly the same as the carbonates.
• The size of grains in a siliciclastic rock characterizes the energy of the trans-
porting media and the depositional environment. In carbonates, the size is
mostly affected by the dimension of the organism and therefore it is not as
42 3 Microscopic Studies

important as clastics. The grain size is generally divided into gravel (>2 mm),
sand (0.063 to <2 mm), and mud (<0.063 mm) (Wentworth 1922). The high
energy of the transporting system and depositional environment expel the fine
grain sediments and increase both grain size and maturity of the rocks. Fine
grain sediments are deposited in low-energy environments.
Siliciclastic reservoirs are often sandstones. A conglomerate or muddy reservoir
is not common. Sieve analysis is the best method for size measurement in loose
unconsolidated cores. Loose sands of the Ahwaz Sandstone unit of the Asmari
Formation, the first discovered carbonate reservoir of the world, is a good
example of such reservoirs. In consolidated rocks, the size is determined by
visual estimation, microscopic graded slides, or by special software. In the first
case, the size is estimated compared with the diameter of a microscopic field of
view. In the second case, a graded slide is used for size determination. Software
must be calibrated before any measurement. In all three methods, a specific size
must be known before measurement. After size determination of grains, their
statistical analysis and interpretations are the same as loose sediments (see
Tucker (2001) for more explanation).
• Minor components are seen in low amounts and are not used for rock classi-
fication including but not limited to glauconite, bioclast debris, carbonate par-
ticles, siderite, and organic materials.
• Textural parameters include sorting, roundness, and maturity. Sorting is the
degree of size similarity of the grains (Fig. 3.9). Roundness is the sharpness of
the grain corners. In other words, it is the curvature of the particle edges
compared to an equal size circle (Fig. 3.10a). Wadell (1932) proposed using the
ratio of the average radius of curvature of the particle corners to the radius of the
maximum inscribed circle as the roundness (Fig. 3.10a).
X
Ro ðroundnessÞ ¼ r=NR ð3:2Þ

where r is the radii of curvature of the grain corners, N is the number of corners,
and R is the radius of the largest inscribed circle within the particle.

Fig. 3.9 The concept of sorting


3.2 Quantitative Carbonate Petrography 43

Fig. 3.10 Wadell roundness and Riley sphericity (a) and visual scale for roundness and sphericity
(b) (modified from Powers 1953)

A circle would have the roundness of 1; the others include: well rounded
(roundness value between 0.60 and 1.00), rounded (0.40–0.60), subrounded
(0.25–0.40), subangular (0.15–0.25), angular (0.00–0.15), and very angular
(particles with extremely sharp edges). The exact roundness value can be cal-
culated using modern software but in an industrial core analysis job, a visual
estimation of the roundness (from 1 to 6) is recommended (Fig. 3.10b).
Various methods have been devised for calculating grain sphericity. Wadell
(1932) and Sneed and Folk (1958) proposed using the 3D properties of the
grains to calculate the sphericity that is not applicable in thin section studies of
the rocks. Riley (1941) defined the projection sphericity as the ratio of the
diameter of a circumscribed circle to that of the diameter of a circle that
inscribes the grain (Fig. 3.10a).

S ðsphericityÞ ¼ ðdi =dc Þ0:5 ð3:3Þ

• Facies and facies coding classification of sandstones is based on Pettijohn


(1975) in most cases (Fig. 3.11). Sandstones are classified based on the pro-
portion of matrix and the grain composition. These clastic rocks are divided into
three groups including arenites, graywackes (commonly abbreviated as wackes),
and mudstones based on their matrix content. Here, the matrix is defined as the
particles finer than the main framework of the sample. This is about 30 lm in
sandstone (Dott 1964). Three main components that are used in classification are
quartz, feldspar, and rock fragments. Other components are not sufficiently
stable and dissolve, disaggregate, or transform in significant transportation. The
term arenite is attributed to those rocks with less than 15% matrix. Samples with
more than 75% matrix are called mudstone. Between these two limits, rocks are
termed wackes. Sandstones are further classified based on the proportion of
three main constituents (excluding matrix and other minerals). Note that these
44 3 Microscopic Studies

Fig. 3.11 Pettijohn classification of sandstones (figure from Wiranto 2013)

three components are normalized to 100% to plot on a ternary diagram. As with


Dunham, this classification is also based on textural parameters and is applicable
in reservoir geology. This nomenclature not only provides a systematic classi-
fication system but also represents the transportation and depositional envi-
ronment energy. Clastic rocks are more divided based on their grain
composition. For example, a lithic arenite is divided into sedimentary, meta-
morphic, and igneous lithic arenite. These detailed classifications are not rec-
ommended in core analysis projects unless a research program is in progress.
The facies coding process is the same as the carbonates.
• Other features such as laminations, mud cracks, brecciation, and the frequency
of opaque minerals are recorded. It is worth mentioning that this part is strongly
dependent on project objectives.
• The sedimentary environment process is the same as the carbonates.
• Porosity types and amount, including the main types of porosity in sandstones
which are intergranular, dissolution, and fracture porosities.
• The fracture process is the same as the carbonates. Sandstones are generally less
fractured than carbonates.
3.2 Quantitative Carbonate Petrography 45

• Cementation where the main cements are ferron, carbonate, and silica deposited
as rim, syntaxial, and pore filling between the grains.
• Compaction is the same as the carbonates.
• Sample selection, laboratory porosity, permeability and grain density, rock
types, and comments are also the same as carbonates.

3.3 Paleontology

Fossils like other grains form part of the rocks. Flora and fauna that were living in
the sedimentary environments (marine and nonmarine) are deposited after death
simultaneously with sediments and fossilized during the diagenesis processes of the
sediments. Thus they are studied along with other components of the rocks in core
studies. They are used for absolute age dating, stratigraphical correlation, sequence
stratigraphy and unconformity, paleoenvironment, paleoechology, and paleo-
bathymetry recognition.
Macrofossils such as gastropods, cephalopods, rudists, pelecypods, and the like
can be studied by the naked eye. Due to the small scale of the core diameter,
macrofossils are often absent in the core with the exception of fossils that form the
framework of facies including framestone and rudstone.
Microfossils with calcareous tests are studied during petrographical analysis and
their relevant data such as classification (genus, species), frequency (such as allo-
chems, from 0 to 4), and their appearance or extinction are recorded. Based on the
recorded data, interpretations are performed and the necessary results are collected.
Petrographical studies on thin sections are not applicable for all microfossils. Some
of the microfossils (e.g., spores, pollens, acritarchs, chitinozoans, conodonts, and
nanofossils) need special processes for extracting fossils from the rocks. In addition
to fossilized organisms, traces produced by an organism are also studied on the
cores (e.g., Knaust 2017). These fossils are called trace fossils or ichnofossils.
Burrows, borings, bioerosion, trails, and trackways are some types of common
ichnofossils that are observable on the core scale. Trace fossils are powerful tools
when the paleoenvironment and paleoechology of deposited rocks are the subjects
of core studies.
After collecting the necessary data from paleontological studies, the geologist is
able to determine the age of the cored section, perform the biozonation based on
studied fossils, and correlate cored intervals with other wells or sections with the
same period of geological time. An example of fossils in core, plug, and thin section
scale can be seen in Fig. 3.12.
46 3 Microscopic Studies

Fig. 3.12 Transversal section of a gastropod in core scale (a), transversal section of a green algae
(b) and orbotolinids (c) in plug scale, photomicrographs of foraminifera Langella sp. (d) and
Orbitolina sp. (e) in thin sections

3.4 XRD Analysis

X-ray diffraction (XRD) is used to identify minerals through recognizing their


crystalline structure. Fine-size minerals that are not distinguishable under the
microscope are detected by this method. Thus it is the best method for clay typing.
The clay-sized fraction (usually less than 4 lm) is separated from the whole
specimen and the powdered sample is illuminated with X-rays of a fixed wave-
length. The intensity and angle of the reflected radiation are used for mineral
identification through matching against a database.
3.4 XRD Analysis 47

3.4.1 Sample Selection

Selecting samples for any analysis has a major effect on final results and inter-
pretations. Routinely, the XRD samples are selected in two stages for a core
analysis project. Round one is at the time of the core opening and the first
macroscopic core description stage before thin section preparation. Samples are
selected based on gamma log and also primary macroscopic core description. The
purpose of this selection is improving mineralogical knowledge before petro-
graphical observations. Also, shale intervals are routinely ignored in core plugging
because it is not possible to prepare any plug sample from shaly intervals. They
disaggregate when they are in contact with water and delaminate under the plugging
drill.
The main stage of XRD sample selection is after the microscopic studies. At this
stage, there are some ambiguities in mineralogical composition of some samples,
especially clays. These samples are good candidates for clay mineralogy. The clay
cement also is important for sandstone reservoirs.
There is no strict rule for the number of samples. This completely depends on the
nature of the formations and project budget. Clays are more frequent in siliciclastic
rocks and thus more samples are analyzed in such a reservoir.

3.4.2 Bulk XRD Analysis

This type of analysis is suitable when the amount of all components of a sample is
nearly the same. They will all be detectable and the amount of each mineral is
determined.

3.4.3 Clay Mineralogy

If the clay volume is not enough in comparison to the other minerals or the entire
sample is composed of clays, clay mineralogy is preferable. If the clays are not
frequent, their peak is covered by the other frequent minerals. Therefore their
concentration should be increased prior to analysis.
XRD analysis is also used for quantifying clay content in sedimentary rocks,
especially sandstones. There are two common approaches to quantitative analysis of
clays. Clay fraction can be separated by physical methods such as Stokes’ law of
settling. This is a relationship between grain size and settling velocity. Practically,
this relationship depends on many parameters such as particle shape, density,
acceleration due to particle contacts, fluid density, and viscosity. Considering some
assumptions for these variables, Stokes’ law of settling is expressed as
48 3 Microscopic Studies

D ¼ ðV=C Þ1=2 ð3:1Þ

where D is the particle diameter and V is the settling velocity.


The variable C depends on grain and fluid density, fluid viscosity, and gravity.
Fluid properties change with temperature. Therefore assuming spherical quartz
grains and distilled water, C varies only with temperature.
Practically, a known weight of a sample (preferentially fine particles with less
than 0.063 mm diameter) is suspended in 1000 cc distilled water in a graduated
cylinder. The diameter of the clay minerals (D) is less than 4 lm. The variable C is
also known for various temperatures. Knowing the diameter and C coefficient, the
velocity is calculated. The sampling depth (h) in the graduated cylinder is optional
(10 or 20 cm is recommended). As velocity is equal to h/t, the time of sampling
(t) is known. A known volume of suspension (20 cc is preferable) is gathered using
a graduated burette from the sampling depth. Actually, at this time all particles with
less than 4 lm diameter are still suspended or particles larger than 4 lm have been
deposited. The sample is put in an oven and the sediment is weighed after drying.
This is the clay weight in the sampled volume. The clay content of the primary
sample is calculated using a proportion between this weight and all volume
(1000 cc). For example, for a 20 cc sample, the weight is multiplied by 50.
This sample can be used for clay typing using XRD analysis. Clay size frac-
tionation is also accomplished by centrifugation. The latter is recommended for
finer grain sizes, commonly less than 2 lm. Some types of clay minerals are larger
than the defined clay size (4 lm) and thus use of a coarser fraction (about 16 lm) is
recommended.
Clay fraction XRD analysis increases the resolution of the resulting peaks.
Therefore the type of clay is determined more accurately. Calculating each clay
volume in the whole sample is not possible in this method. The whole-rock XRD
analysis is capable of providing reliable data on the amounts of clay minerals
present in the sample. However, depending on various types of clay abundance, it
may not be possible to distinguish, or even identify, certain types of clay minerals
in a whole rock sample. Clay minerals have a major role in reservoir properties if
they are present. Shale layers are effective barriers for fluid flow. Kaolinites reduce
the porosity of the sandstones whereas illite reduces permeability. Smectites swell
with adsorbing water and cause many drilling problems as well as formation
damage.

3.5 SEM Analysis

Scanning electron microscopy (SEM) magnifies the sample sending focused beams
of electrons and detecting secondary or backscattered electron signals. In contrast to
petrographical analysis, SEM shows the actual three-dimensional minerals,
cements, pore structures, and relationships. Advanced SEM instruments have a
3.5 SEM Analysis 49

Fig. 3.13 The effect of scale in SEM analysis. The highlighted rectangle in a is magnified in
b and the highlighted rectangle in b is illustrated in c. Moldic porosities (a), their connections (b),
and microporosity between micrite particles (c) in one sample

significantly higher magnification range than any other geological laboratory


instruments. The common magnification range for analyzing reservoir rocks is
between 150 and 10,000. For example, the first one is used for viewing the moldic
porosities in an ooid grainstone whereas the second detects the framboidal pyrites.
The ability of the tool to continuously change the magnification in a wide range
provides valuable information on sample properties (Fig. 3.13). An experienced
geologist is able to distinguish various minerals based on their morphological
properties. Most SEM instruments are equipped with an energy-dispersive X-ray
analyzer (EDX or EDA) that specifies the elemental concentration of the sample.
Comparing morphological properties with elemental concentration leads to the
accurate identification of rock constituents.

3.5.1 Sample Selection

Selecting the sample for SEM analysis and imaging depends on the purpose of the
study but in most cases they are selected after the petrographical studies. At this
stage, there is an estimation of geological rock types. Pore types are the main factor
for sample selection (e.g., Tavakoli and Jamalian 2018). Samples with the same
pore types have almost the same pore throat size distribution (PTSD). According to
the project purpose and available budget, some samples are selected from each pore
type, which is the same in each rock type (see Chap. 6). It is ideal to use samples
with the same depth for both SEM and mercury injection capillary pressure (MICP)
tests. The PTSD is double-checked and the image of the quantitatively measured
pore throats will be available.
50 3 Microscopic Studies

SEM analyses have other purposes, too. Some minerals, especially clays such as
kaolinite, illite, and chlorite, have distinct forms when analyzed with SEM. The
instrument also shows microporosity in the sample. If there is any need for ele-
mental analysis in the specific point or line, a SEM equipped with the EDX is the
best choice.

3.5.2 EDX Analysis

Elemental composition of the samples shows their geological characteristics in


many cases. Such an analysis is used as a proxy for interpreting depositional
environments and their conditions, paleoclimate, diagenesis, exposure surfaces, and
many other parameters. EDX is used to provide quantitative elemental identification
and chemical characterization of the sample. A high-energy beam composed of
charged particles or a beam of X-rays is focused into the sample being studied. The
beam may excite and eject an electron from an inner shell. The vacant place is then
filled with an electron from an outer, higher energy shell. The energy difference
between the higher and lower shell electrons is emitted in the form of X-rays. An
energy-dispersive spectrometer measures the number and energy of the released
X-rays. These energies are characteristics of the atomic structure and the type of
element is thereby determined.
Line analysis is also possible in a combined SEM–EDX analysis. A specified
number of points are determined on a line and elemental concentration is measured
on each point. Changes in elemental concentrations show the variation of the
depositional or diagenetic (depending on the sample) conditions. This is useful for
cement stratigraphy and interpreting changes in diagenetic environments. Elemental
map analysis (EMA) is also possible. In this analysis, the frequency of each element
is specified with a predefined color. The result is like a map.

3.6 Microscopic Uncertainties

Petrography is a semiquantitative task and depends heavily on the experience of the


petrographer. Error in estimation of grain and porosity percentage is common for
beginners. Point counting is more accurate but it is still possible to push a wrong
key or make a mistake in mineral, grain, cement, or matrix identification. Many
parameters are recorded from one thin section and there are a lot of thin sections in
any core analysis task. Just imagine that one percent of the input data is wrong. This
could easily happen when the petrographer is recording the number in the prede-
fined data sheet. For example, if it is necessary to record 30 parameters from each
thin section and 1000 thin sections are present in the project, 30,000 records will be
transferred from the expert mind to the digital data sheet. One percent of this
3.6 Microscopic Uncertainties 51

number is 300 records! It is not possible to double-check the results in most


projects, as time is very important. It is important to reduce the level of errors and
follow the general trend of the data in any project instead of one single record.

References

Ahr WM (2008) Geology of carbonate reservoirs: the identification, description, and character-
ization of hydrocarbon reservoirs in carbonate rocks. Wiley-Interscience, Malden
Bebou DG, Loucks RG (1984) Handbook for logging carbonate rocks. Bureau of Economic
Geology, Texas
Dickson JAD (1965) A modified technique for carbonates in thin section. Nature 205:587
Dott RH (1964) Wacke, greywacke and matrix; what approach to immature sandstone
classification? J Sed Res 34:625–632
Dunham RJ (1962) Classification of carbonate rocks according to depositional texture. In:
Ham WE (ed) Classification of carbonate rocks. AAPG Memoir 1, pp 108–121
Embry AF, Klovan JE (1971) A Late Devonian reef tract on Northeastern Banks Island, NWT.
Bull Can Petrol Geol 19:730–781
Evamy BD (1963) The application of a chemical staining technique to a study of dedolomitisation.
Sedimentology 2:164–170
Flugel E (2010) Microfacies of carbonate rocks, analysis, interpretation and application. Springer,
Berlin
Friedman GM (1959) Identification of carbonate minerals by staining methods. J Sediment Petrol
29:87–97
Friedman GM (1965) Terminology of crystallization textures and fabrics in sedimentary rocks.
J Sediment Petrol 35:643–655
Knaust D (2017) Atlas of trace fossils in well core—appearance, taxonomy and interpretation.
Springer International Publishing, Cham
Mehrabi H, Mansouri M, Rahimpour-Bonab H, Tavakoli V, Hassanzadeh M (2016) Chemical
compaction features as potential barriers in the Permian-Triassic reservoirs of southern Iran.
J Petrol Sci Eng 145:95–113
Pettijohn FJ (1975) Sedimentary rocks. Harper & Row, New York
Powers MC (1953) A New Roundness Scale for Sedimentary Particles. SEPM J of Sediment Res
Vol. 23
Reading HG (1986) Sedimentary environments and facies. Blackwell Scientific Publications,
Oxford
Reeder RJ (1983) Crystal chemistry of the rhombohedral carbonates. In: Reeder RJ
(ed) Carbonates: mineralogy and chemistry. mineralogical society of America reviews in
mineralogy, vol 11. Mineralogical Society of America, pp 1–47
Riley NA (1941) Projection sphericity. J Sed Res 11:94–95
Scholle PA, Ulmer-Scholle DS (2006) Color guide to petrography of carbonate rocks. AAPG
memoir 77, AAPG, Tulsa
Sibley DF, Gregg JM (1987) Classification of dolomite rock textures. J Sediment Petrol 57:967–
975
Sneed ED, Folk RL (1958) Pebbles in the lower Colorado River, Texas, a study in particle
morphogenesis. J Geol 66:114–150
Tavakoli V, Jamalian A (2018) Microporosity evolution in Iranian reservoirs, Dalan and Dariyan
formations, the central Persian Gulf. J Nat Gas Sci Eng 52:155–165
Tucker ME (2001) Sedimentary petrology: an introduction to the origin of sedimentary rocks.
Wiley-Blackwell, Oxford
Wadell H (1932) Volume, shape, roundness of rock particles. J Geol 40:443–451
52 3 Microscopic Studies

Warren J (2000) Dolomite: occurrence, evolution and economically important associations.


Earth-Sci Rev 52:1–81
Wentworth CK (1922) A scale of grade and class terms for clastic sediments. J Geol 30:377–392
Wiranto RS (2013) FOBEX 1: All about sandstone. Available via AAPG. http://aapg.ft.ugm.ac.id/
fobex-1-all-about-sandstone. Accessed 5 Feb 2018
Chapter 4
Macroscopic Studies

Abstract Macroscopic core studies investigate large-scale core properties that are
not detectable by any other methods. A macroscopic description of the core is the
final step in geological observations and reveals the remaining rock properties
between the plugs. As plugging is not possible or vital in some important lithologies
such as shales or anhydrites recording these variations and involving them in any
interpretation is crucial. Building both conceptual and numerical reservoir models
without involving macroscopic core properties leads to vital mistakes in reservoir
characterization. Routinely, there are two stages of macroscopic core description.
The first stage is after core cleaning for the first evaluation of the rocks, helping
microscopic considerations and selecting primary samples for various analyses. The
second stage is at the end of geological studies. Cores are slabbed prior to
description. The fresh uncontaminated surface of the rock is available at this time.
Core-scale rock properties are recorded on standard sheets that were prepared prior
to description. These include but not limited to missed rock layers, rock color,
sedimentary structures, compaction-related features such as stylolites and solution
seams, bedding and stratal surfaces, intraclasts and rip-up clasts, large and trace
fossils, visual porosities, fractures and their properties, fining and coarsening
upward trends, and oil staining. The recorded parameters are illustrated on a core
description log with appropriate scale. The results are integrated with microscopic
geological studies as well as routine core analysis data to achieve the best inter-
pretations and build an accurate reservoir model.

4.1 Core Slabbing and Resination

After finishing the routine core analysis (RCAL), geological studies and selecting
the samples for special core analysis (SCAL) tests, cores are slabbed longitudinally.
It is recommended to cut the sample into 1/3 and 2/3 ratios, making preparation of
additional plugs possible from the thicker part, if necessary. Water is used as cooler
fluid in most cases. Regarding core cleaning and plugging, the interaction of the
core and water should be considered. Using salt water as cooler fluid is very

© The Author(s) 2018 53


V. Tavakoli, Geological Core Analysis, SpringerBriefs in Petroleum
Geoscience & Engineering, https://doi.org/10.1007/978-3-319-78027-6_4
54 4 Macroscopic Studies

Fig. 4.1 An example of one-meter core resination

expensive and also precipitates the salt in the core. Gasoline is a better choice but it
also penetrates into the core and changes its properties. With the slabbing process, a
fresh flat surface of the rock is available for macroscopic studies. One third of the
rock is mounted in a shallow plastic tray and fixed using clear epoxy resin with its
flat side down. The length of the tray is one meter. A sufficient amount of epoxy is
poured into the tray to make a basal substrate. Then the epoxy is added again to fix
the sample completely. The liquid should have appropriate viscosity to flow around
the core and fill the gaps between the core parts. The upper circular surface is cut
after epoxy hardening. The core specifications including formation, well, field,
depth, client, and so on are recorded above the tray (Fig. 4.1). The positions of the
plugs are empty. The specifications of the plug such as orientation, depth, and
number are recorded in the plug space. These trays could be used for future geo-
logical studies. They are stored in drawers of special boxes. Further disordering,
breakage, and disaggregation will also be minimized. Putting the core in resin
prevents direct access for sampling or some treatments such as acid reaction. If the
core contains significant amounts of oil, it could react with the epoxy.

4.2 Core Description

After core slabbing, the slabbed surfaces are ready for geological descriptions. The
purpose of macroscopic core description is recognition of macroscopic core prop-
erties such as macroscopic depositional, structural, and diagenetic features of the
whole or slabbed core. The macroscopic features can be recorded after core cleaning
on whole cores or after slabbing on the fresh slabbed core surface, but the main
geological core description process is performed after the slabbing. At this stage, a
flat, fresh, and clear surface of the rock is available for the geologist. The purpose of
the first stage of core description is to help RCAL sample selection and to obtain more
effective petrographical descriptions. The recorded data are integrated with other
sources and provide the basis for further reservoir studies such as reservoir quality
evaluation and sequence stratigraphy. A core description provides a permanent,
easily accessible record of core features. All macroscopic features of the cores are
recorded during this stage. The core description should begin from the bottom of the
core upward. This is consistent with the sedimentation pattern. Thus the geologist can
see the changes of rocks according to the time order of deposition. A logging page or
sheet is recommended for systematic data collection (Fig. 4.2, e.g.). It is better to
4.2 Core Description 55

Fig. 4.2 An example of a core description sheet. Items are variable based on core properties

record each parameter with a predefined symbol, in order that more time in the lab can
be saved during the process. There is no standard set of symbols but Bebout and
Loucks (1984) provide a good example. Some others are also illustrated in Fig. 4.3.
Certain simple equipment is necessary in this process. A magnifier, spray bottle of
water, a sponge, hydrochloric acid, ruler, protractor, pencil, waterproof whiteboard
marker, eraser, paper and paper holder, and plastic bags for probable sampling are
necessary (Fig. 4.4). A photo scale should be used in each photo (Fig. 4.5).
Various software has been developed for the core description process and
installed on laptops or even tablets. Although all parameters can be recorded on
prepared sheets or software, a paper notebook is recommended as there are always
some parameters beyond the predictions. At least two people should be involved in
a core description job. Before starting the analysis:
56 4 Macroscopic Studies

Fig. 4.3 Some useful symbols for macroscopic core description


4.2 Core Description 57

Fig. 4.4 Some simple equipment for macroscopic core analysis

Fig. 4.5 Photo scale for macroscopic core photography. Scale is applicable when the width of the
rectangle box is 15 mm
58 4 Macroscopic Studies

• Lay out the core on the table. The height of the table and the environmental light
are two important factors. Sunlight should be used if possible or a light projector
with both white and yellow color lamps.
• Check the length of the cores with the coring report and microscopic records. Be
sure that no core was missed during core handling or sampling.
• Check the core order against depth. They should be arranged from bottom to
top.
• Check the top and bottom of each box. This is routinely forgotten during the
core description process.
• Hang the previously prepared logs, such as microscopic descriptions, evaluated
wire line, or even the wire line log data on the wall of the laboratory.
• Assess the entire cored interval before starting the description. Review the major
boundaries such as formations or reservoir units.
The parameters of interest are different for siliciclastic and carbonates.
Generally, the following parameters are recorded.
• Missed layers. In a routine plugging, plugs are prepared every 30 cm. Therefore
some variations are missed between the plugs. These features such as shale or
anhydrite layers can have important effects on the final reservoir studies. They
should be recorded carefully and applied in future reservoir evaluation studies.
• Color of the core. There are two primary factors for any color including hue and
the quality of lightness (light and dark colors). When we name a color, actually
we are naming its hue. Color naming of the cores should kept as simple as
possible and have general acceptance. Standard color charts can be used for this
purpose. When a secondary color is also evident, a combination of two colors is
applied. Greyish-green, greyish-gray, and milky-white are some examples.
Combining the hue and lightness gives a wide range of names. The core color
could be an indicator of the paleoenvironmental conditions such as oxidation
state. It also gives clues to the composition of the rock (such as red color for iron
in oxidizing environments).
• Sedimentary structures such as laminations, cross-beddings, graded beddings,
soft-sediment deformations, flaser and lenticular beddings, mud cracks, and load
casts. These structures should be recorded accurately because they are one of the
most important keys for paleoenvironmental interpretations.
• Compaction-related features such as stylolites and solution seams. They are
good indicators of the paleostress of the samples. The stylolites and solution
seams also act as reservoir barriers (or even carriers) in some cases.
• Beddings and stratal surfaces. Such surfaces are the result of changes in envi-
ronmental conditions, such as depositional energy. Stratal surfaces are very
important in sequence stratigraphic interpretations.
• Intraclasts and rip-up clasts. They are mostly derived from the lower bed and
indicate a reactivation of the environmental energy. Carbonate intraclasts are
autochthonous in most cases.
4.2 Core Description 59

• Visual porosities. Porosities are also recorded in microscopic studies but some
large pores, such as large vugs, cannot be detected in microscopic studies.
• Trace fossils. They are fossilized structures produced in previous substrates
(including sediments, sedimentary rocks, or organic materials) by the organisms
(Knaust 2017). Biotic sedimentary structures (Frey 1973) including bioturba-
tion, bioerosion, biodeposition, and biostratification could also be recorded and
interpreted. Examples are given in Knaust (2017).
• Fractures. The fracture presentation of the core is described in Sect. 4.3.
• Fining or coarsening upward sequences. They obviously show the change in
depositional energy or some special events (such as turbidity currents).
• Oil staining. Record the level of oil staining present in the samples. The 1–4
scale is applied here too.
It is recommended to transfer all recorded parameters to a final log. The core
logs show the features against depth and are very useful for later interpretations (see
Sect. 4.4). Various software has been developed to draw a geological log. A core
description log has basic parameters gathered from microscopic or wire line logging
(such as lithology or natural gamma ray) plus all macroscopic features recorded at
the time of core description.

4.2.1 Core Photography

Photos provide a visual record of the whole, slabbed, or some special parts of the
core. They are used to visualize core features easily, minimize core handling, and
digitally archive the sample characters. Both natural (white) and ultraviolet
(UV) light (fluorescence mode) are used for core photography. The natural light
shows core features and allows us to examine some macroscopic core records after
the description. Ultraviolet light photography is used to highlight hydrocarbon
stained zones, oil-bearing parts, and fluid contacts. Heavy oils show orange-brown
whereas light hydrocarbons show bright yellow fluorescence. Rocks with no
hydrocarbon are dark yellow to purple under UV light (Fig. 4.6). UV images
should be taken as soon as possible following slabbing because light hydrocarbons
are vaporized from the sample. Note that UV imaging is not necessary for
nonassociated gas reservoirs (such as South Pars and North Dome fields), as all
hydrocarbons dissipate just after core opening. Care should be taken to differentiate
oil-based mud filtrate from the natural hydrocarbons. Core images can be used in
petrophysical evaluation packages to compare with wire line analysis and geo-
logical information.
The whole core photos are not perfect for later recognition of all core features.
Therefore during macroscopic core description, close-up photos are also taken of
some important and outstanding features (Fig. 4.7). Some features are more vivid
when the rock is wet, and a water spray is very useful in these cases.
60 4 Macroscopic Studies

Fig. 4.6 Three meters of cores stained with oil in some places under UV and normal lights

Fig. 4.7 An example of core close-up photos. Intraclasts (a), shale lamination (b), cross-bedding
in carbonates (c), and moldic porosities with a facies change surface (d). Scale is the same for all
four photos
4.3 Fracture Presentation 61

4.3 Fracture Presentation

The first step in fracture study of a reservoir is image log interpretations, if there are
any, but the first step on the cores is whole core CT-scanning. CT images show the
fractures and some of their properties before core opening and layout (see Sect. 2.3).
The fractures of the cores are studied simultaneously with core description. There are
two stages of fracture studies. The first phase is after core cleaning. Oriented coring
is possible, but most core samples have rotated during the coring process and their
direction is unknown. Considering the well inclination, the dip angle of the fracture
is calculated. If there are any image logs from the coring interval, some cores can be
oriented using the dip direction of the fractures on the image logs. Fracture properties
are also transferred to a fracture log and the final result is illustrated beside the other
rock properties (such as porosity, permeability, and lithology). If there are any
fractures parallel to the slabbing surface, it may be possible that they were missed in
the core fracture study. Thus an integration of all data sources should be used.
A predefined sheet is used for recording fracture properties from the cores (Fig. 4.8).
General information is recorded from the core (such as core and box number).
The depth of the fracture should be known as with any other parameters of the
reservoir. The depth of the fracture midpoint is recorded. Visible length of the
fracture shows the core and fracture intersection. Fractures can be systematic, which

Fig. 4.8 An example of the fracture study sheet and related parameters that are recorded from the
cores (Courtesy of M. Eliassi)
62 4 Macroscopic Studies

are formed by tectonic forces, or nonsystematic and not related to the tectonic
regime of the area (see Nelson 2001). Good examples are shrinkage cracks. Induced
fractures are not related to any natural forces and are created at the drilling or
transfer stages by human interference (see Lorenz and Cooper 2018). The opening
size of the fracture surface is recorded as the aperture size. Fractures can be closed,
open, or filled by cement. This opening has remained empty or filled by mineral
cement. The depths of photos taken from the fractures are also recorded on the
sheet. The fracture frequency per meter represents the fracture intensity in each
meter of the cores.

4.4 Core Log Preparation

Core logging consists of recording core parameters side by side with regard to a
reference. This reference is the depth in most cases. Various parameters are
recorded regarding the purpose of logging. There is no strict rule for parameter
selection or ordering in a log. Some companies have standard log formats that must
be followed in the core log preparation. Using a standard has many advantages. The
output can be easily understood by others and more effective scientific communi-
cation is reached. Anyway, the final result is a multicolumned graphical paper
presentation in a predefined scale. Regarding the core length, different scales could
be selected but 1/200 is appropriate in most cases. The paper log is about 2 m long
for a 400-m cored well in this scale. Selecting an optimum scale is very important
because either very long or very short logs have some disadvantages. A very long
log with 1/20 scale, for example, is not applicable in reservoir studies because the
geologist cannot have a general look at the cored interval. Actually, the interpreter
just sees heterogeneities. On the other hand, a lot of data are missed in a very short
log. If data are in a digital format, preparing a log with two different scales is also
possible.
There are unlimited combinations of parameters for building a log that depends
on the purpose of the study. For example, if the theory is that fossils change with
the environment and facies, they should be illustrated beside the biostratigraphical
log. It is very useful to have a kick-off meeting with the client’s experts before
starting the core description and logging because they are the end-users of the data.
Various logs are prepared for various parts of the core study. Routine core logs
include microscopic, macroscopic (core description), biostratigraphy, fracture, and
reservoir properties. Some parameters are similar for all logs. Routinely all prepared
logs have depth, formation track, reservoir unit, and age. Plotting some wire line
logs beside the other parameters on a core log is very useful. A GR log, for
example, represents the shaleness variation and provides a very good indicator for
stratigraphic correlation between wells.
It is recommended to input data into computer software before illustrating it
graphically. A spreadsheet is appropriate for this purpose. Many of the core
description parameters are defined qualitatively using alphabets instead of numbers.
4.4 Core Log Preparation 63

The author’s recommendation is to try to be more quantitative in any part of a core


analysis job. Having data in digital format has many advantages. Data may be easily
copied, transferred, compared, and gathered in just one database. The numbers are
used for core log preparation, which is not possible using alphabets. At least, a
Boolean file could be prepared. Each item has value “1” if the character has been
recorded at that depth. The other depths have “0” value. Software can use these
numbers for illustrating a core description log that shows the presence or absence of
core macroscopic features.
There are also various methods to depict data on a graphical paper log. These
include symbols, lines, dots, horizontal bars, presence/absence, color, and alphabets
(descriptive) (Fig. 4.9). Selecting the type of illustration for each datum also
depends on the analyzer but there are some accepted customs. Symbols are used for
lithology, sedimentary structures, and uncommon objects. All of them could be
plotted in just one column. Curves are used for continuous numeric data such as
wire line logs. Dots are used for frequent discontinuous data or comparing two
datatypes. In such cases, the second datatype is illustrated by dots. Comparing a
neutron porosity log with RCAL porosity is a good example. Horizontal bars are
suitable for integer or dispersed numeric characters such as the amount of each pore
type, percentage of allochems, numeric facies code, rock types, and so on. Using
color without any other indicator is not recommended as many publications are in
black and white mode. Combining color and horizontal bars is appropriate in many
cases such as sedimentary and electrofacies, sedimentary environments, and
reservoir rock types. Presence/absence (Boolean) drawing is used for most of the
core description parameters such as presence of stylolites, intraclasts, brecciation,
bioturbation, or anhydrite nodules. If none of these methods is applicable, alphabets
are used for descriptive data. Core color, vuggy porosities, sequence stratigraphic
surfaces, and fissility are some examples.
The order of data columns from left to right depends on the datatype and
importance. Lithology affects many other parameters in a reservoir and wire line
logs are unique in a core log. They are also important because extending geological
core properties to the uncored wells is based on core–wire line log comparison. The
first and most important column is the depth track. Any data in a reservoir are
meaningless without depth.
Petrographical features are recorded on a microscopic log (see Sects. 3.2 and 3.3
for these parameters). It is recommended to include porosity and permeability in all
logs for comparison with reservoir properties. As the amount of porosity in thin
section study is also illustrated beside the routine porosity and permeability, rec-
ognizing micro and intercrystalline porosities is also possible.
The macroscopic log is composed of features recorded in the core description
process on slabbed or whole cores. Close-up core photos are taken at the same
time. Some microscopic features are also plotted in the core description log.
Microscopic lithology, for example, helps a lot in comparing the microscopic and
macroscopic results.
Fracture properties are depicted on a fracture log. The log routinely includes
depth, lithology, porosity, and permeability, various types of surfaces (such as
64 4 Macroscopic Studies

Fig. 4.9 Schematic picture of some datatypes and the final core description log

bedding or sequence stratigraphic boundaries), CT-derived fractures, image log


fractures, stylolites and solution seams, slabbed core fractures, and some properties
from wire line logs (such as sonic velocity).
The appearance and extinction of fossils, actually the first and last appearance on
the cored intervals, are illustrated on a fossil log. The frequency of each fossil
(semiquantitative) could also be depicted.
A combination of reservoir properties and their geological proofs is plotted on a
reservoir geology log. The log includes lithology, some wire line data, core-derived
4.4 Core Log Preparation 65

porosity and permeability, pore types, rock types, and any geological parameter that
affects reservoir properties. This log, in conjunction with various charts (see
Sect. 6.3), is used to interpret the relationships between geological characteristics
and reservoir properties.
The log should be plotted on a continuous paper. This facilitates comparison of
two cores, formations, reservoir parts, and so on. In some cases, the log should be
kept as simple as possible. You don’t have to plot all of your data! Sometimes this
confuses the client or end-user in application or interpretation. Try to draw the log
simply and applied, as much as possible.

4.5 Sequence Stratigraphy and Reservoir Zonation

After microscopic and macroscopic data gathering and interpretations, this infor-
mation is used to reconstruct the interwell reservoir properties. An appropriate
framework is needed for this purpose. Such a framework should honor the geo-
logical, chronostratigraphical, and reservoir boundaries. The science of sequence
stratigraphy is used as this basis. Sequence stratigraphy considers the genetically
related strata in a chronostratigraphic framework. Genetically related strata have
been deposited in the same physical, chemical, and biological conditions and thus
have the same facies properties and probable diagenetic history. This branch of
science was invented based on the need for determining reservoir boundaries. Prior
to this invention, petroleum geologists had realized that lithological and biostrati-
graphical boundaries are not exactly applicable for reservoir zonation. They thought
of new chronostratigraphic boundaries that bound the rocks and facies with the
same reservoir properties. Rocks with the same sea-level conditions have the same
genetic situations and the same primary characteristics. Similar facies characteris-
tics cause the same flow behavior and therefore they also have the same diagenetic
processes. Some diagenetic features such as fractures are exceptional.
Regarding the nature of sequence stratigraphic concepts, sequence stratigraphy is
the most effective method for reservoir zonation (van Buchem et al. 1996; Martin
et al. 1997; Angulo and Buatois 2012; Enayati-Bidgoli and Rahimpour-Bonab
2016; Tavakoli 2017). In such zonation, a whole studied formation is divided into
zones or layers with respect to its reservoir properties, mainly porosity and per-
meability. In a reservoir zone, reservoir behavior for fluid flow and storage is
uniform. As the sequence stratigraphic units are genetically related and have similar
sedimentological properties, they have similar reservoir properties too. Therefore
they are good indicators of reservoir behavior. Some deviations can be seen due to
the fabric-retentive diagenesis.
Relative sea-level change is the result of interaction between sediment supply
(S) and change in accommodation space (A). Although increasing accommodation
causes landward migration of facies (retrogradation), increasing sediment supply
(siliciclastics) or sediment creation (carbonates) results in seaward migration of the
facies (progradation). The rate of S and A can be the same, resulting in in situ
66 4 Macroscopic Studies

sediment accumulation (aggradation). Sediments deposited in one stage of inter-


action between S and A are called the stacking pattern. Generally, retrogradation
forms a fining upward cycle. In contrast, the result of progradation is a coarsening
upward sequence.
Discussion about the sequence stratigraphic concepts is beyond the scope of this
book but some basic concepts are discussed for better understanding of the subject.
There are three main sequence stratigraphic models routinely used in the petroleum
industry. The first one, introduced by Embry and Johannessen (1992), has two
systems tracts. Systems tracts are genetically related strata deposited at a certain
stage of sea-level change (Posamentier et al. 1988). In their definition, a complete
sequence is composed of a regressive (RST) and a transgressive (TST) systems tract
(ST). It is known as a T–R sequence model. A regressive ST is formed when the
effect of S is more than A in the depositional environment and thus sediments are
prograded into the sea. Actually, a coarsening upward cycle is formed. A sequence
boundary lies between the shallowing and deepening cycles. The shallowest facies
indicate the sequence boundary (SB) and the deepest facies represent the maximum
landward position of the facies (maximum flooding surface, MFS). Determination
of the sequence boundary is almost simple in this method because two different
cycles with different stacking patterns are compared. The second model was
developed by Exxon company researchers (e.g., van Wagoner et al. 1988, 1990;
Posamentier et al. 1988), and is called the Exxon model. In this model, the
regressive part of the T–R model is divided into two STs including lowstand
(LST) and highstand (HST). The LST includes deposits that accumulate at the start
of sea-level rise and it lies on the nonconformities or their correlative conformities.
These surfaces show the lowest state of the sea-level and are considered as the SB.
In this model, the LST is capped by the TST and overlies the HST of the previous
sequence. The HST is bounded below by the MFS and finishes with the SB. It
should be noted that the stacking pattern of both the LST and HST is progradational
but distinguishing a nonconformity is very helpful in separating the sequences. In
the third model developed by Hunt and Tucker (1992), the HST of the Exxon model
is divided into two systems tracts including HST and FRST (forced regressive
systems tract) or FSST (falling stage systems tract). The FRST is composed of
sediments accumulated after the onset of a relative sea-level fall and before the start
of the next sea-level rise. Indeed, there is an event of erosion in many cases at this
stage. Again, the LST, HST, and FRST have the same stacking patterns. Inasmuch
as no sediment is deposited in the FRST stage in most basin locations, distin-
guishing HST from FRST is not simple in the third model (Fig. 4.10).
The key question is which model is the best for zonation of a specific hydro-
carbon reservoir. It completely depends on the case. If, for example, you are
working on a deep marine stratum, maybe Embry is the best model because there is
no complete evidence for separating other ST. If you have high-resolution data such
as seismic, cores, and wire line logs from a carbonate ramp, probably the Hunt and
4.5 Sequence Stratigraphy and Reservoir Zonation 67

Fig. 4.10 Comparison of three sequence stratigraphic models described in the text

Tucker model is the best. In the first case, you have not enough facies change or
even reservoir quality variations to separate four ST but in the second example
karstification of the carbonate sediments and rocks is very important, which is
developed in the FRST stage. In most cases, the T–R or Exxon models are
preferable.

4.6 Macroscopic Uncertainties

The quality of a macroscopic description completely depends on the geologist’s


experience. A trained eye can distinguish many parameters correctly but they are
not clear for a beginner. Recording data on a sheet and transferring them to the
software and the final log may cause some errors. The final interpretation also
depends on the geologist’s knowledge. Fining upward sequences, for example, are
deposited in various environments with different factors. Compiling microscopic
and macroscopic data for a reasonable sequence stratigraphic reservoir zonation is
not a simple problem. As macroscopic description is routinely the final stage of
geological core observations, all data should be integrated for understanding the
geological parameters that specify the reservoir behavior.
68 4 Macroscopic Studies

References

Angulo S, Buatois LA (2012) Integrating depositional models, ichnology, and sequence


stratigraphy in reservoir characterization: the middle member of the Devonian-Carboniferous
Bakken formation of subsurface southeastern Saskatchewan revisited. AAPG Bull
96:1017–1043
Bebout DG, Loucks RG (1984) Handbook for logging carbonate rocks. Bureau of Economic
Geology, Texas
Embry AF, Johannessen EP (1992) T–R sequence stratigraphy, facies analysis and reservoir
distribution in the uppermost Triassic–Lower Jurassic succession, Western Sverdrup Basin,
Arctic Canada. In: Vorren TO, Bergsager E, Dahl-Stamnes OA, Holter E, Johansen B, Lie E,
Lund TB (eds) Arctic geology and petroleum potential. Special Publication 2, Norwegian
Petroleum Society, pp 121–146
Enayati-Bidgoli AH, Rahimpour-Bonab H (2016) A geological based reservoir zonation scheme in
a sequence stratigraphic framework: a case study from the Permo-Triassic gas reservoirs,
offshore Iran. Mar Petrol Geol 73:36–58
Frey RW (1973) Concepts in the study of biogenic sedimentary structures. J Sediment Petrol
43:6–19
Hunt D, Tucker ME (1992) Stranded parasequences and the forced regressive wedge systems tract:
deposition during base-level fall. Sed Geol 81:1–9
Knaust D (2017) Atlas of trace fossils in well core-appearance, taxonomy and interpretation.
Springer International Publishing, Cham
Lorenz JC, Cooper SP (2018) Atlas of natural and induced fractures in core. Wiley, Hoboken
Martin AJ, Solomon ST, Hartmann DJ (1997) Characterization of petrophysical flow units in
carbonate reservoirs. AAPG Bull 81:734–759
Nelson RA (2001) Geological analysis of naturally fractured reservoirs. Gulf Professional
Publishing, Boston
Posamentier HW, Jervey MT, Vail PR (1988) Eustatic controls on clastic deposition, I. Conceptual
framework. In: Wilgus CK, Hastings BS, Kendall C, Posamentier HW, Ross CA, Van
Wagoner JC (eds) Sea-level changes—an integrated approach. Society of Economic
Paleontologists and Mineralogists Special Publication, vol 42, pp 109–124
Tavakoli V (2017) Application of gamma deviation log (GDL) in sequence stratigraphy of
carbonate strata, an example from offshore Persian Gulf Iran. J Petrol Sci Eng 156:868–876
van Buchem FSP, Razin F, Homewood PW, Philip JM, Eberli GP, Platel JP, Roger J, Eschard R,
Desaubliaux GMJ, Boisseau T, Leduc JP, Labourdette R, Cantaloube S (1996) High resolution
sequence stratigraphy of the Natih formation (Cenomanian/Turonian) in Northern Oman:
distribution of source rocks and reservoir facies. GeoArabia 1:65–91
van Wagoner JC, Posamentier HW, Mitchum RM, Vail PR, Sarg JF, Loutit TS, Hardenbol J
(1988) An overview of the fundamentals of sequence stratigraphy and key definitions. In:
Wilgus CK, Hastings BS, Kendall C, Posamentier HW, Ross CA, Van Wagoner JC
(eds) Sea-level changes—an integrated approach. Society of Economic Paleontologists and
Mineralogists Special Publication, vol 42, pp 39–45
van Wagoner JC, Mitchum RM, Campion KM, Rahmanian VD (1990) Siliciclastic sequence
stratigraphy in well logs, cores and outcrops. American Association of Petroleum Geologists
Methods in Exploration Series, vol 7, 45 pp
Chapter 5
Geochemical Analysis

Abstract Inorganic geochemistry is used indirectly for reservoir rock analysis.


Carbon and oxygen stable isotopes, strontium isotopes, and elemental concentra-
tions are used for this purpose. In isotope analysis, the ratio of the heavier to the
lighter isotope type is measured. This ratio is compared to a standard. The differ-
ence is positive if the sample contains heavier isotopes and is negative if it is
reached in light isotopes. The fractionation has a major role in isotope values of
different samples. Vital effects, for example, cause negative excursion in organisms.
Both carbon and oxygen isotope ratios are used for sequence stratigraphy and
reservoir zonation, recognition of nonconformities, and hiatuses and mass extinc-
tions. Paleotemperature can be calculated by oxygen isotope ratios. The balance
between continental and mantle Sr input to the oceans determines the variations of
this isotope. The result is used for absolute age dating and understanding sea-level
fluctuations. Elemental analysis of the rocks also provides some important proxies
for interpreting paleoenvironmental conditions, stratigraphic correlations, facies
classifying, provenance studies, and the rate of weathering. Uranium geochemistry
has attracted more attention in recent years because it is available from many
reservoirs through spectral gamma logging. Rate of erosion and redox conditions,
as well as original mineralogy are inferred from uranium distribution in a studied
formation. Sample selection is very important in geochemical analysis because the
final results and interpretations strongly depend on sample type, distance from each
other, and final quality control. Studies of some of these aspects and their appli-
cations are just at their beginning stages.

5.1 Sample Selection

Geochemical sample selection strategy controls the final data quality. Sample
quantity is dictated by the project budget in most cases. The number of candidate
samples for geochemical studies is considerably less than the number of plugs or
thin sections. The majority of samples must be representative of the studied interval,

© The Author(s) 2018 69


V. Tavakoli, Geological Core Analysis, SpringerBriefs in Petroleum
Geoscience & Engineering, https://doi.org/10.1007/978-3-319-78027-6_5
70 5 Geochemical Analysis

not special features. Therefore try to select the sample from the dominant facies and
lithology. If a special feature or interval is of interest, select at least three samples
from it or its counterparts in the studied interval. Selecting just one sample increases
the uncertainties because you will not be sure about the value!
The other important factors are time and material availability. Isotopic analysis
does not need a huge volume of samples in many cases but the number of wells is
an important issue. A sampling interval of 1–2 m is recommended. Regular sam-
pling is not mandatory. The sampling could be denser near the intervals of interest
such as formation boundaries or nonconformities. For example, if a Permian–
Triassic section is studied, the sampling interval reduces near the boundary. The
important point about isotope analysis is that the sample must be autogenic. This is
not a problem for most of the carbonates inasmuch as siliciclastics are terrigenous in
most cases. The isotope composition of a detrital grain only shows the formation
conditions and has no application for environmental interpretations Therefore
sampling from clasts of conglomerates and breccias must be avoided. Secondary
processes change the isotopic composition of rocks, especially in carbonates.
Bulk rock or a selective analysis of some special parts is possible. As a small
quantity is analyzed in most cases, heterogeneity of the sample cannot be detected
by bulk rock analysis. A drilling sample with a dental drill or using a laser
microprobe is preferable. The isotope zoning within a single shell or cement can be
detected using a laser microprobe. These advanced techniques make it possible to
select for analysis the more stable minerals that represent the original depositional
conditions.
Always provide petrographical and macroscopic properties of the samples for
the laboratory. These properties are also very useful for future interpretations. Label
your samples with a unique code including core number, plug number, and ori-
entation and depth. The samples should be taken from the central intact part of the
cores. Coring and cleaning fluids as well as laboratory treatments (such as plug-
ging) have less effect on this part.
Even after all care and treatments, there are some spikes in the data in most
cases, especially for heterogeneous formations. This could be for different reasons.
First, it is always possible that the wrong data change the general trends. They
could be ignored with some considerations. The geological properties of the sam-
pling point must be considered carefully. After that, they are compared with the
other time equivalents in the same field, other fields or sections, and other sections
of the world. Second, based on the author’s experience, the geochemical changes
have obvious trends. In other words, sharp changes are rare. Therefore compare the
spike with the nearest datapoints. They should also show a comparative change.
Third, the data trends are interpreted in most cases, and are not a unique value. They
show the change in depositional conditions more clearly and meaningfully.
5.2 Isotopes 71

5.2 Isotopes

Atoms of an element with the same atomic number but different atomic mass are
called isotopes. The difference in their mass is due to the different number of neu-
trons in their nuclei. “Isotope” is a Greek word that means equal places and denotes
their same position in the periodic table. The nuclei of stable isotopes do not
spontaneously decay and are stable compared to the age of the Earth. Each element
has different types of isotopes relating to the number of neutrons in its nucleus. For
example, carbon has two natural stable isotopes including 12C and 13C. At 98.93%,
12
C is the most frequent carbon isotope in nature (Rosman and Taylor 1998).
Isotopes separate naturally between two different substances or two phases of one
material. This process is called isotope fractionation and plays an important role in
geochemical studies. Generally, lighter isotopes are more mobile because of their
lower mass. It is obvious that they move more easily than their heavier counterparts.
The lighter isotopes also possess higher vibrational frequency, and therefore their
bonds break more easily with lower energy. This is why the organisms tend to break
the bond of the lighter isotopes and use them in their vital activities.
Measuring the absolute abundance of each isotope is very difficult. Therefore,
the isotope ratio of the heavier to the lighter isotope is determined. The result is
compared with a reference material. Isotope abundance is reported as the relative
difference between the ratios of that isotope in the material to an international
standard. The d sign is used to show this difference (Eq. 5.1).

dA ¼ ðRa =Rst Þ  1 ð5:1Þ

where Ra is the ratio of the heavier to the lighter isotope in the studied material and
Rst is this ratio in a standard. Accordingly, the d13C is the difference between the ratio
of 13C to 12C in material and this ratio in a standard substance. The d is the ratio of
two similar quantities and is a dimensionless value. This is multiplied by 103 and
expressed as per mil (‰). As the heavier isotope is the numerator, the d is positive
when the sample has a heavier isotope and is negative when the sample is obtained
from the lighter isotopes. The latter is frequently seen in the body of organisms.
The isotopic composition of seawater and the resulting autogenic sediments have
previously been determined for various geological times (e.g., Veizer 1989; Veizer
et al. 1999). Comparison of each isotope datum with its time equivalent reveals the
diagenetic trends of the rocks.

5.2.1 Carbon Isotope

The isotope ratios are measured by an isotope ratio mass spectrometer (IRMS). For
carbon isotope analysis, CO2 or CO gas is used. It is obvious that the sample must
contain a carbon element for isotope ratio measurement. Carbon isotope analysis is
72 5 Geochemical Analysis

not possible on a pure quartz arenite containing just silicon and oxygen. Anyway,
there are some other constituents containing carbon in most cases. The carbonates
are reacted with pure phosphoric acid and released CO2 is used for carbon ratio
isotope measurement.
All measured isotope ratios should be comparable to each other and thus an
international standard is needed. Vienna Pee Dee belemnite (VPDB) is the most
famous standard for carbon isotope measurements. This is the isotopic ratio of
Cretaceous Pee Dee formation belemnite used at Chicago University in the early
1950s.
There are two main carbon reservoirs in the world including organic matter and
sedimentary carbonates. Two main fractionation mechanisms determine the final
carbon isotope of these two sources. The first is the vital effect of the organisms. As
mentioned, the light carbon isotopes possess higher vibrational frequency and
therefore form weak bonds with the other elements. Such bonds do not take a large
amount of energy to break. Living organisms prefer to break the lighter isotope
bond to get carbon elements for their metabolism. Therefore, the organic materials
contain lighter isotopes and have more negative d value. The reaction between
atmospheric CO2, dissolved bicarbonates, and solid carbonate concentrate the
heavier carbon in rocks. Temperature is another factor that reduces isotope frac-
tionation. It is believed that the carbon fractionation is about 0.035‰ per degree
Celsius (Grossman 1984; Emrich et al. 1970). Therefore temperature has a negli-
gible effect on carbon isotope fractionation. The range of carbon isotope values has
been determined in various carbonate reservoirs before (Fig. 5.1).

Fig. 5.1 Carbon isotope ratio (d13C) in some important carbon reservoirs (modified from Hoefs
2009)
5.2 Isotopes 73

Generally, the d13C of carbonate shells varies according to the water depth
(Kroopnick 1985; Berger and Vincent 1986). Biological activity decreases with
increasing depth. In the photic zone of shallow waters, there is a high rate of organic
productivity that reduces the light 12C isotope of the waters. The remaining shells in
the shallow waters have higher 13C isotopes. This effect reduces with increasing the
water depth. Many studies attempted to calculate the water depth and the ther-
mocline thickness based on such a gradient (e.g., Srinivasan and Sinha 1998; Tiwari
et al. 2015).
Carbon isotope excursions are used in sequence stratigraphy and reservoir
zonation of the reservoirs. In transgressive systems tracts (TST), most parts of the
continental shelf are covered by seawater. The photic zone area increases effectively
and biological activities rise accordingly. A huge volume of carbonate shells is
buried in transgressive deposits and the seawater is enriched in the 13C isotope (e.g.,
Holmden et al. 1998; Immenhauser et al. 2003; Swart and Eberli 2005; Huck et al.
2013). In regressive cycles, the previously deposited carbonate tests are eroded or at
least not produced abundantly. Thus the precipitated sediments have lighter carbon
isotopes and a negative shift in d13C. The Permian–Triassic boundary is a good
example. There are still considerable debates about the cause of extinction and
sea-level change but many studies believe that both sea-level regression and
extinction events cause a negative carbon isotope excursion in this boundary
(Tavakoli et al. 2018 and references therein). A mass extinction causes death of
organisms with light carbon isotope composition. Regression eroded the previously
precipitated sediments with negative isotope values and reinforced the negative
carbon excursion. The effect of atmospheric CO2 must be considered in all cases as
it also changes the isotopic equilibrium of other carbon reservoirs.
The effect of diagenesis must be taken into account in all studies. The diagenetic
processes change the isotopic composition of the rocks after deposition and
therefore the final isotopic ratio is not a good indicator of the depositional envi-
ronments in these cases. Carbonate stability is another important factor. Carbonates
originally are deposited as aragonite or high Mg–calcite and then converted to low
Mg–calcite during diagenesis. Such a conversion is accompanied by a considerable
change in the isotopic composition of the rocks. This effect is negligible for carbon
isotopes in most cases because the main diagenetic factor is water which has no or
small amounts of the carbon element. These systems are rock-buffered and thus do
not change the carbon isotope composition of the sediments. The exception is
subaerial exposures where decay of organic materials produces a large volume of
CO2 with very low carbon isotope ratios. The CO2 is combined with water and
interacts with previously deposited carbonates. The result is low Mg–calcite with
very low d13C (Fig. 5.2). Such an excursion accompanied by a negative oxygen
shift is used for recognition of sea-level fall when enough facies and paleontology
data are not available.
74 5 Geochemical Analysis

Fig. 5.2 Carbon and oxygen isotope ratios from an exposed Permian–Triassic transition in
western Tethys show a negative shift of both isotopes (Tavakoli et al. 2018)

5.2.2 Oxygen Isotope

The main rules are the same for all isotopes, especially for oxygen and carbon. They
are studied together in most cases. Both carbon and oxygen isotopes can be ana-
lyzed by the derived CO2 gas from the sample. Oxygen is the most common
element on Earth. It is found in almost all carbonates and siliciclastic rocks.
Therefore, it is one of the most interesting elements in geological isotopic analysis.
There are three naturally occurring oxygen isotopes including 16O (99.757% fre-
quency), 17O (0.038%), and 18O (0.205%) (Rosman and Taylor 1998). The lightest
isotope, 16O, and the heavier one, 18O, are more frequent and thus their ratio is
normally determined. They also have more mass difference and are easier to
measure.
Oxygen isotopes of silicates and oxides are liberated through fluorination.
Decomposition at high temperature using carbon reduction is also used for some
compounds. The process for isotope ratio measurement in carbonates is the same as
for carbon isotopes, as described in Sect. 5.2.2. Two standards are used for cal-
culating d18O including VPDB and VSMOV (Vienna standard mean ocean water).
These two standards are converted to each other using the following formulas
(Hoefs 2009).
5.2 Isotopes 75

d18 O ðVSMOWÞ ¼ 1:03091d18 O ðPDBÞ þ 30:91 ð5:2Þ

and

d18 O ðPDBÞ ¼ 0:97002d18 O ðVSMOWÞ  29:98 ð5:3Þ

The oxygen isotope fractionation follows the general rules of isotope fraction-
ation. Fractionation of isotopes between liquid water and vapor is essential for
interpreting many characteristics of oxygen isotope change in sedimentary rocks.
The oxygen isotopic composition of meteoric waters is controlled mainly by
Rayleigh fractionation. Water is vaporized from the equator and precipitated in
different latitudes. The water vapor contains more light isotopes. Vapor conden-
sation and precipitation also enrich the remaining vapor from the light isotope
(16O). It is obvious that if the SMOW standard is used, any meteoric water derived
from equatorial waters has negative d18O values. This negative excursion increases
with increasing latitude. The decayed organic materials in soils have a negative
ratio due to the vital effect of organisms. Therefore meteoric waters passing through
soils and sediments have a strong negative d value.
The main reservoirs of oxygen are rocks and waters in nature (Fig. 5.3). Isotope
fractionation between carbonates and seawater is determined by phase change and
temperature. McCrea (1950) established the first empirical equation between d18O
and paleotemperature of the depositional environment for inorganic calcite. Epstein
et al. (1953) introduced such an equation in organic calcite based on mollusk shells.
Further studies refined the equation several times. Erez and Luz (1983) introduced
the following equation for temperature calculation in carbonates.

T ¼ 17  4:52 ðd18 Oc  d18 OwÞ þ 0:03 ðd18 Oc  d18 OwÞ2 ð5:4Þ

where T is temperature (degree centigrade), d18Oc is the isotopic ratio of carbon-


ates, and d18Ow is the isotopic ratio of water (both compared to the standard).

Fig. 5.3 Oxygen isotope


ratio (d18O) in some
important oxygen reservoirs
(modified from Hoefs 2009)
76 5 Geochemical Analysis

Various considerations should be taken into account before temperature calculation.


Diagenesis changes the isotopic composition of metastable carbonates including
aragonite and high Mg–calcite. Water is a main reservoir of oxygen isotopes and
effectively changes the isotopic composition of the rock. The diagenetic environ-
ment (marine, meteoric, or burial) can be inferred by combining carbon and oxygen
isotope changes of the sample. Vital effect is another major problem.
Shell-secreting organisms must be in equilibrium with the seawater from which
they have precipitated. Therefore tests with negligible or no vital effect on oxygen
isotopes (such as brachiopods) must be selected. The vital effect of most organisms
has been calculated previously (e.g., Wefer and Berger 1991).
The isotopic composition also changes in icehouse and greenhouse periods. In
cold icehouse stages, the ice caps expand and light oxygen isotopes lock in these
ices, and the isotopic composition of the remaining waters becomes heavier.
Simultaneously, the sea-level falls. Change in oxygen isotope ratios can also be
used for deducing the global sea-level fluctuations caused by paleoclimate change.

5.2.3 Strontium Isotope

The strontium element has four isotopes including 84Sr (0.56% frequency), 86Sr
(9.87%), 87Sr (7.04%), and 88Sr (82.53%). The 87Sr is radiogenic and produced by
87
Rb decay. The 87Sr/86Sr ratio is used in low-temperature isotope geochemistry.
As 87Sr is the product of 87Rb decay, it has increased continuously since the
creation of Earth. The Sr isotopic composition of seawaters of the various geo-
logical times is determined by analyzing this ratio in autogenic marine carbonates.
This ratio has been determined accurately in the Phanerozoic eon (e.g., McArthur
et al. 2001). The 87Rb is continuously separating from the mantle and oceanic crust
and enriches the continental crust. This Rb and its product, 87Sr, inputs to the
oceans from three sources including weathering and transportation from the con-
tinental crust, input from the mid-ocean ridges, and dissolution and precipitation of
marine carbonates. As the mantle and crust ratios are two end-members (Fig. 5.4),
the Sr ratio could be used as a good proxy for the tectonic evolution of the Earth.
The Sr has a long residence time in the oceans (2–5  106 years). Ocean water
masses mix every 103 years and thus the Sr ratio in oceanic waters in a relatively
short time is controlled by the Sr input to the waters (e.g., Veizer 1989; McArthur
2007; McArthur et al. 2012). In such periods, Sr is used as the weathering index.
The average strontium ratio in the mantle is 0.7035 and it is 0.7119 for the crust
(Palmer and Elderfield 1985). Increasing the ratio represents the continental erosion
and riverine input to the oceans. This erosion would be the result of climatic or
tectonic activities. High tectonic activities cause more erosion and a higher Sr
isotope ratio. High humidity and temperature and increasing rainfall have the same
effect. The dissolution and recrystallization of previously deposited carbonates also
change the ratio in oceanic waters.
5.2 Isotopes 77

Fig. 5.4 Strontium isotope ratio of various sources for oceanic waters. Dotted lines show Sr ratios
of seawater. Dashed line shows the modern diagenetic Sr input to the oceans (from Wierzbowski
2013)

The 87Sr/86Sr of carbonates is used for absolute age dating (strontium isotope
stratigraphy, SIS). This ratio is 0.709241 in present-day seawaters (Elderfield 1986)
and has been changed through geological time. The Sr incorporates into the car-
bonate minerals, replacing the calcium especially in aragonite. Phosphates and
evaporites also contain variable amounts of Sr. This ratio in a well-preserved
sample is compared with the calibration curve and the absolute age is determined
(see McArthur et al. 2012 for more discussion on the method). The accuracy of the
resulting age depends on the sample quality, analytical accuracy, and the slope of
the calibration curve. The selected sample must be diagenetic free as much as
possible because secondary processes change the isotopic composition of the rocks
and the resulting age does not represent the real sample age. Mud-dominated car-
bonates routinely have the least diagenetic overprints and are the most suitable
samples for bulk rock analysis. If microdrilling or laser microprobe is possible, low
Mg–calcite would be the best selection. The slope of the calibration line, which is
the rate of change in Sr isotope ratios, is also important. Best results may be
obtained from the more accurate calibration curve with sufficient slope such as
Early Jurassic, Late Cretaceous, and Late Paleogene to recent intervals. The best
accuracy for Sr age dating is ±0.1 Ma (McArthur et al. 2012). The isotopic ratio of
Sr also is used for chemostratigraphy and correlating time-equivalent strata
throughout the world. In addition, SIS is used for estimating the duration of
stratigraphic gaps and calculating the sedimentation rate of strata.
The Sr ratio of rocks changes during meteoric diagenesis. As discussed, conti-
nental rocks have a higher Sr ratio. Thus meteoric waters increase this ratio as a
78 5 Geochemical Analysis

result of water–rock interactions. Therefore, prior to any measurement, the effect of


diagenesis must be determined using petrographic observations, geochemical
analysis, cathodoluminescence considerations, or staining techniques. Low Mg–
calcite shells such as brachiopods and oysters are suitable for Sr isotope mea-
surements. Comparison of the bulk rock Sr isotope analysis with coeval samples
has been successfully used for validating the bulk rock data. Diagenetic changes
that are inferred from such comparisons can be used for tracing secondary alter-
ations of the rocks and their related diagenetic environment. The vital effect of
fractionation by organisms has less effect on Sr isotopes and thus they are reliable
proxies for tracing diagenetic changes. 87Sr/86Sr is also used in deducing the time
and model of dolomitization, origin of diagenetic cements, separating marine and
nonmarine sediments, the origin and timing of mineral deposits, and groundwater
studies.

5.3 Elemental Analysis

The characteristics of the studied rock are the most important point in sample
selection, tests, and later interpretations in elemental analysis. Carbonates and
siliciclastic are completely different from this point of view. The chemical com-
position of terrigenous rocks is inherited from their original source. Carbonates are
created in the environment and are greatly influenced by diagenetic processes.
Therefore, their composition reflects these two environments. This point must be
seriously considered before any analysis or interpretations.
There are various methods for elemental analysis of a geological sample such as
inductively coupled plasma-optical emission spectrometry (ICP-OES), inductively
coupled plasma-mass spectrometry (ICP-MS), X-ray fluorescence (XRF), atomic
absorption spectrophotometry (AAS), spark source mass spectrometry (SSMS),
neutron activation analysis (INAA and RNAA), laser-induced breakdown spec-
troscopy (LIBS), electron microprobe analysis, and geochemical logging (see
Craigie 2018 for more details). All of these methods determine the type and fre-
quency of elements in the samples. The data quality control is mandatory before
starting the interpretations. Preparing a chemostratigraphic log is very useful.
Elemental concentrations are plotted alongside the lithological, facies, rock types,
or any other required data. It is important to establish a relationship between ele-
mental variations and other rock properties, if there is any. Grain size distribution
controls these changes in some cases (Craigie 2018). The mineral density controls
the distribution of heavy and light minerals and therefore the final geochemical
properties for bulk rock samples (e.g., Lahijani and Tavakoli 2012). Organic
materials and clays transport some special elements. For example, high uranium
content is associated with increasing organic materials and clay content (e.g.,
Spirakis 1996; Fiet and Gorin 2000). Calcium is replaced by strontium in car-
bonates especially aragonite. Diagenesis strongly affects the elemental composition
of the rocks, especially carbonates. Dolomitization increases the Mg content of
5.3 Elemental Analysis 79

sedimentary rocks. This change does not represent the depositional conditions. Any
interpretation based on such variations is unacceptable. Plotting each parameter
(e.g., organic, sand and clay contents, dolomite percentage, or facies code) versus
the element of interest characterizes their relationship (e.g., Tavakoli and
Rahimpour-Bonab 2012). Plotting two elements against each other reveals the
cause of changes in many cases (Fig. 5.5).

Fig. 5.5 Pairwise comparison of various parameters of the Late Permian–Early Triassic carbonate
samples of the Persian Gulf. Thorium, potassium, and shale volume (Vsh) show strong covariation
(a–c) whereas uranium has no relationship to them (d–f). Change in anhydrite content and
total organic carbon (TOC) also has no effect on uranium (g, h) (from Tavakoli and
Rahimpour-Bonab 2012)
80 5 Geochemical Analysis

Elemental analysis data provide a significant tool for stratigraphical correlations


(Ramkumar 2015). The method is very helpful when paleontological data are
limited, such as the deep ocean sediments. Changes in elemental concentrations as
well as their trend are used beside other data extracted from core analysis to
correlate precisely the formations, reservoir boundaries, sequence stratigraphic
surfaces, and nonconformities. The correlated strata are used for reservoir zonation
and 3D modeling.
Elemental concentrations have been used for classifying the facies of the rocks.
For example, Blatt et al. (1972) classified the sandstones based on a ternary diagram
of Na, Fe + Mg, and K vertices. Lithology or facies classification based on ele-
mental concentration is not a routine task in geochemistry of the cores because
petrographical data are cheaper and more available.
Various ratios of the elements are used for recognition of the sediment’s
provenance. This is more applicable to clastic rocks because carbonates are created
in the environment. Na, K, Zr, Th, Ti, and Nb are used frequently. Using any ratio
or absolute values depends completely on the rock properties, probable provenance,
and previous successful experiences and studies. For example, the amount of Ca,
Na, and K and their ratio is used for recognition of an albite–anorthite–orthoclase
series. This is applicable when a possible igneous rock is the provenance of the
sample. This collection has a very different meaning in carbonates. The Na and K
are evidence of evaporation in a carbonate–evaporite formation. The Ca, Na, and K
evaporites are deposited with an increase in the rate of evaporation, respectively.
Elemental concentration is also used for deducing paleoenvironmental condi-
tions. The redox potential and pH are two important parameters of the environment.
High concentrations of Mo, Cu, Co, Ni, Zn, Cr, U, and V are often used to prove
the anoxic situations at the time of deposition (Wedepohl 1971; McLennan 2001).
An anoxic environment is also associated with dark color, presence of anoxic
minerals such as pyrite, and high total organic carbon that must be considered
carefully (Craigie 2018).
The rate of weathering represents the paleoclimate at the time of deposition. It is
also associated with nonconformities in most cases. The alkali metals (Na, Ca, K,
Mg) as well as Al are routinely used as the weathering indices (Craigie 2018).
Various indices have been developed to quantify the effect of source rock weath-
ering based on the objectives of the study. Obviously the chemical aspect of
weathering is discussed here. The concept was proposed for the first time with the
work of an early researcher, Reiche (1943). He compared the mobile cations with
the hydroxyl water and the amount of silica. Miura (1973) proposed another index
based on the mobility of ferric and ferrous iron. Nesbitt and Young (1982) intro-
duced the chemical index of alteration (CIA) as

CIA ¼ ½Al/ðAl þ Ca þ Na þ KÞ ð5:5Þ

where Ca* is the calcium content in the siliciclastic fraction. The Ca, Na, and K are
released from the feldspar minerals during chemical weathering resulting in high Al
5.3 Elemental Analysis 81

concentration and CIA increase. The weathering index (WI) (Retallack 1997;
Pearce et al. 2005) is calculated as

WI ¼ Al=ðCa þ Mg þ Na þ KÞ ð5:6Þ

Mg and Ca are derived from siliciclastic minerals. Both indices show the
weathering intensity. The effect of grain size, provenance, and special minerals
must be considered in all cases.
The leaching of some elements, U, for example, is another good indicator of
subaerial exposure and the resulting weathering. Uranium is mobile in oxic situa-
tions and is easily removed from the exposed surfaces concentrated in the lower
layers (see Sect. 5.4 for more explanations).

5.4 Uranium Geochemistry

Uranium geochemistry has attracted more attention in recent years. This is because
of the wide use of a spectral GR tool in hydrocarbon wells. The amount of thorium
(Th), potassium (K), and uranium (U) is measured by a natural gamma detector.
The resolution is about 15 cm and thus a very good proxy is available for further
studies. Among these, U has unique geochemical properties and behavior.
Handheld spectral gamma counters are also used in geological field works. The
measuring mechanism and results are the same. The resulting log is a very useful
tool for stratigraphical correlations. K and Th are derived from the land in most
cases and represent the continental input but U has a more complicated story. In
siliciclastics, U is commonly associated with shale or heavy minerals, particularly
zircon, apatite, and monazite. In carbonates, different conditions including oxida-
tion state, leaching, organic material (OM), and also clay content determine the
frequency of this element. Type and abundance of carbonate minerals as well as
brine concentration also control the U distribution in space and time (Yang et al.
2015). The U6+ is more easily mobilized than another U state, U4+. In oxidizing
environments, U4+ tends to oxidize to U6+ and remains in solution as UO22+
(uranyl). Soil surfaces, organic materials, clays, carbonate minerals, and metal
oxides adsorb dissolved U6+ and form uranium complexes. In contrast, Th is not
affected by redox conditions. The ratio of these two elements is very useful in
paleoenvironmental interpretations.
Th and K are derived from the weathering of continental rocks. They are
enriched in fine-grain siliciclastic rocks. There is a direct relationship between K,
Th, and the volume of shale in most cases. The chemistry of K has been discussed
before.
In anoxic conditions, U6+ is reduced to U4+ and precipitated as uraninte (UO2)
(Wignall and Twitchett 1996). Therefore anoxic sediments contain high amounts of
uranium. In oxic situations, the mobile U6+ remains in solution. Change from Late
Permian anoxic to an oxygenated Early Triassic environment has been studied
82 5 Geochemical Analysis

before (Ehrenberg et al. 2008; Tavakoli and Rahimpour-Bonab 2012). Anoxic


environments are suitable for source rock deposition. It is also favored for forma-
tion of some unconventional reservoirs such as oil or gas shales.
U remobilization occurs if oxygen or oxygenated waters penetrate into the
previously deposited sediments (e.g., Naderi-Khujin et al. 2016). The element is
leached and concentrated in the lower strata. The result is a U peak in the rock or
sediment column. This peak is a useful indicator for recognizing nonconformities
especially where other data are not sufficiently available.
The original minerals of carbonates are mostly aragonite and calcite. The ionic
radii of uranyl and Sr2+ exceed that of Ca2+ thus U and Sr concentrate in a
metastable aragonite mineral. Therefore, the original mineralogy of carbonates can
be understood using the amount of Sr and U. The shift from aragonite to calcite seas
have been studied before using Sr variations in carbonates (Heydari et al. 2013;
Tavakoli 2016). U can also be used for this purpose, which is in its beginning steps.

References

Berger WH, Vincent E (1986) Deep-sea carbonates: reading the carbon isotope signal. Geol
Rundsch 75:249–269
Blatt H, Middleton G, Murray R (1972) Origin of sedimentary rocks. Prentice Hall, New Jersey
Craigie N (2018) Principles of elemental chemostratigraphy. Springer, Cham
Ehrenberg SN, Svana TA, Swart PK (2008) Uranium depletion across the Permian-Triassic
boundary in Middle East carbonates: signature of oceanic anoxia. AAPG Bull 92:691–707
Elderfield H (1986) Strontium isotope stratigraphy. Palaeogeogr Palaeoclimatol Palaeoecol
57:71–90
Emrich K, Ehhalt DH, Vogel JC (1970) Carbon isotope fractionation during the precipitation of
calcium carbonate. Earth Planet Sci Lett 8:363–371
Epstein S, Buchsbaum R, Lowenstam H, Urey H (1953) Revised carbonate-water isotopic
temperature scale. Geol Soc Am Bull 64:1315–1326
Erez J, Luz B (1983) Experimental paleotemperature equation for planktonic foraminifera.
Geochim Cosmochim Acta 47:1025–1031
Fiet N, Gorin GE (2000) Gamma-ray spectrometry as a tool for stratigraphic correlations in the
carbonate-dominated, organic rich, pelagic Albian sediments in central Italy. Eclogae Geol
Helv 93:175–181
Grossman EL (1984) Carbon isotopic fractionation in live benthic foraminifera—comparison with
inorganic precipitate studies. Geochim Cosmochim Acta 48:1505–1512
Heydari E, Arzani N, Safaei M, Hassanzadeh J (2013) Ocean’s response to a changing climate:
clues from variations in carbonate mineralogy across the Permian-Triassic boundary of the
Shareza Section, Iran. Global Planet Change 105:79–90
Hoefs J (2009) Stable isotope geochemistry. Springer, Berlin
Holmden C, Creaser RA, Muehlenbachs K, Leslie SA, Bergström SM (1998) Isotopic evidence for
geochemical decoupling between ancient epeiric seas and bordering oceans: implications for
secular curves. Geology 26:567–570
Huck S, Heimhofer U, Immenhauser A, Weissert H (2013) Carbon-isotope stratigraphy of Early
Cretaceous (urgonian) shoal-water deposits: diachronous changes in carbonate-platform
production in the north-western Tethys. Sed Geol 290:157–174
References 83

Immenhauser A, Della Porta G, Kenter JAM, Bahamonde JR (2003) An alternative model for
positive shifts in shallow-marine carbonate d13C and d18O. Sedimentology 50:953–959
Kroopnick P (1985) The distribution of 13C of RCO2 in the world oceans. Deep Sea Res 32:57–84
Lahijani H, Tavakoli V (2012) Identifying provenance of South Caspian coastal sediments using
mineral distribution pattern. Quatern Int 261:128–137
McArthur J (2007) Recent trends in strontium isotope stratigraphy. Terra Nova 6:331–358
McArthur JM, Howarth RJ, Bailey TR (2001) Strontium isotope stratigraphy: LOWESS Version
3: best fit to the marine Sr-isotope curve for 0–509 Ma and accompanying look-up table for
deriving numerical age. J Geol 109:155–157
McArthur JM, Howarth RJ, Shields GA (2012) Strontium isotope stratigraphy. In: Felix M,
Gradstein FM, Ogg JG, Schmitz M, Ogg G (eds) The geologic time scale 2012, pp 127–144
McCrea JM (1950) On the isotopic chemistry of carbonates and a paleotemperature scale. J Chem
Phys 18:849–857
McLennan SM (2001) Relationships between the trace element composition of sedimentary rocks
and upper continental crust. Geochem Geophy Geosy 2:2000GC000109
Miura K (1973) Weathering in plutonic rocks. Part I weathering during the late-Pliocene of Gotsu
plutonic rock. J Soc Eng Geol Jpn 14(3)
Naderi-Khujin M, Seyrafian A, Vaziri-Moghaddam H, Tavakoli V (2016) Characterization of the
late Aptian top-Dariyan disconformity surface offshore SW Iran: a multiproxy approach.
J Petrol Geol 39:269–286
Nesbitt HW, Young GM (1982) Early Proterozoic climates and plate motions inferred from major
element chemistry of lutites. Nature 299:715–717
Palmer MR, Elderfield H (1985) Sr isotope composition of sea water over the past 75 Myr. Nature
314:526–528
Pearce TJ, Wray DS, Ratcliffe KT, Wright DK, Moscariella A (2005) Chemostratigraphy of the
upper carboniferous schooner formation, southern North Sea. In Colinson JD, Evans DJ,
Holiday DW, Jones NS (eds) Carboniferous hydrocarbon geology: the southern North Sea and
surrounding onshore areas Yorkshire Geological Society, Occasional Publication Series, vol 7.
Yorkshire, pp 147–164
Ramkumar Mu (ed) (2015) Chemostratigraphy: concepts, techniques and application. Elsevier,
Amsterdam
Reiche P (1943) Graphic representation of chemical weathering. J Sediment Petrol 13:58–68
Retallack GJ (1997) A colour guide to paleosols. Wiley, Chichester
Rosman JR, Taylor PD (1998) Isotopic compositions of the elements (technical report):
commission on atomic weights and isotopic abundances. Pure Appl Chem 70:217–235
Spirakis CS (1996) The roles of organic matter in the formation of uranium deposits in
sedimentary rocks. Ore Geol Rev 11:53–69
Srinivasan MS, Sinha DK (1998) Early Pliocene closing of the Indonesian Seaway: evidence from
northeast Indian Ocean and southwest Pacific deep sea cores. J Southe Asian Earth 16:29–44
Swart PK, Eberli G (2005) The nature of the d13C of periplatform sediments: implications for
stratigraphy and the global carbon cycle. Sed Geol 175:115–129
Tavakoli V (2016) Ocean chemistry revealed by mineralogical and geochemical evidence at the
Permian-Triassic mass extinction, offshore the Persian Gulf, Iran. Acta Geol Sin 90:1852–1864
Tavakoli V, Rahimpour-Bonab H (2012) Uranium depletion across Permian-Triassic boundary in
Persian Gulf and its implications for paleooceanic conditions. Palaeogeogr Palaeoclimatol
Palaeoecol 350:101–113
Tavakoli V, Naderi-Khujin M, Seyedmehdi Z (2018) The end-Permian regression in the western
Tethys: sedimentological and geochemical evidence from offshore the Persian Gulf, Iran.
Geo-Mar Lett 38:179–192
Tiwari M, Singh AK, Sinha DK (2015) Stable isotopes: tools for understanding past climatic
conditions and their applications in chemostratigraphy. In: Ramkumar Mu
(ed) Chemostratigraphy: concepts, techniques and application. Elsevier, Amsterdam, pp 65–92
Veizer J (1989) Strontium isotopes in seawater through time. Annu Rev Earth Planet Sci 17:141–
167
84 5 Geochemical Analysis

Veizer J, Ala D, Azmy K, Bruckschen P, Buhl P, Bruhn F, Carden GAF, Diener A, Ebneth S,
Godderis Y, Jasper T, Korte C, Pawellek F, Podlaha OG, Strauss H (1999) 87Sr/86Sr, d13C
and d18O evolution of Phanerozoic seawater. Chem Geol 161:59–88
Wedepohl KH (1971) Environmental influences on the chemical composition of shales and clays.
In: Ahrens LH, Press F, Runcorn SK, Urey HC (eds) Physics and chemistry of the Earth.
Pergamon, Oxford
Wefer G, Berger WH (1991) Isotope paleontology: growth and composition of extant calcareous
species. Mar Geol 100:207–248
Wierzbowski H (2013) Strontium isotope composition of sedimentary rocks and its application to
chemostratigraphy and palaeoenvironmental reconstructions. Ann Phys 68:23–37
Wignall PB, Twitchett RJ (1996) Oceanic anoxia and the end Permian mass extinction. Science
272:1155–1158
Yang Y, Fang X, Li M, Galy A, Koutsodendris A, Zhang W (2015) Paleoenvironmental
implications of uranium concentrations in lacustrine calcareous clastic-evaporite deposits in the
western Qaidam Basin. Palaeogeogr Palaeoclimatol Palaeoecol 417:422–431
Chapter 6
Rock Typing

Abstract Rock typing is the process of assigning reservoir properties to geological


facies. The samples in an ideal rock type have the same geological and reservoir
properties. Such a unit is considered as a reservoir building block. The properties of
interest are fluid storage and flow and thus it is possible that two facies are grouped
in one rock type or one facies is divided into two rock types. There are three main
categories for this process including geology, reservoir (static properties), and
petrophysics. Geological rock types are defined by integrating facies and diagenesis
characteristics of the samples in a porosity–permeability framework. Reservoir
methods use porosity, permeability, pore throat size, and their relationships for
dividing the samples into various rock types. Logs are routinely grouped by cluster
analysis into electrofacies. The final result is an ideal unit that contains the same
geological, reservoir, and wire line characteristics. As wire line data are available
from almost all wells and reservoir intervals, this ideal unit could be distributed in
3D space, even with limited core data. These units are perfect for both static and
dynamic modeling. On a regional scale, the concept of flow unit is used to divide
the reservoir into compartments with the same reservoir quality. Sequence strati-
graphic concepts are also used for correlating the defined units in the reservoir
scale. All methods try to reduce the reservoir heterogeneity to understand reservoir
behavior on microscopic and macroscopic scales.

6.1 Geological Rock Typing

A perfect rock type has the same geological, petrophysical, and reservoir properties.
Rocks with the same depositional characteristics and diagenetic history have the
same reservoir behavior. But it is very normal that two different samples with
different diagenetic processes show the same fluid flow and storage capacity. In the
same way, two samples with similar primary (facies) characteristics with a different
diagenetic history have different reservoir properties. Such changes complicate the
process of geological rock typing. Inasmuch as the rock types must have the same
range of porosity and permeability, routinely geological rock typing is done after

© The Author(s) 2018 85


V. Tavakoli, Geological Core Analysis, SpringerBriefs in Petroleum
Geoscience & Engineering, https://doi.org/10.1007/978-3-319-78027-6_6
86 6 Rock Typing

finalizing the geological and routine studies. At this stage, facies and diagenetic
properties of the samples are known and their porosity and permeability have been
measured.
Routinely after geological studies, the geologist has a good sense of the sample’s
nature. For example, the grain-dominated samples may have more porosity or loose
sandstones contain a considerable amount of intergranular empty spaces. Plotting
various charts to understand the effect of geological parameters on reservoir
properties is very useful. Porosity–permeability plots of various lithologies, facies,
and diagenetic processes (see Sect. 6.3) clearly show their effect on reservoir
behavior. These parameters can also be plotted against each other. For example,
plotting the dolomite percentage versus the amount of porosity represents their
relationship. It should be noted that an integration of all charts, diagrams, and
interpretations should be involved in the final geological rock typing. It is recom-
mended to check out the porosity–permeability crossplot of each facies. Comparing
these plots clearly shows the reservoir properties of each facies. At this stage, two
facies with the same reservoir properties will be integrated. Also, one facies with
two types of fluid behavior is divided into two rock types (Figs. 6.1 and 6.2).
As pore types play an important role in reservoir properties, samples with the
same pore type are differentiated in most cases. For example, an ooid grainstone
with disconnected moldic porosities has high porosity with low permeability. This
is the same for a mud-dominated carbonate sample with a considerable amount of
microporosity, but both of them are not routinely present in a reservoir. In the same
way, an ooid grainstone with high interparticle porosity has the same behavior with
a dolomitized packstone from the fluid storage and flow point of view. This time,
both of them can be present in a reservoir and they form one rock type.
Regarding the various facies, diagenetic processes, and pore types, it seems that
many geological rock types could be defined in a reservoir. This is not true because
the depositional characteristics and diagenetic processes of one reservoir are lim-
ited. For example, in a carbonate ramp environment, facies and environments have
low diversity. In most cases, one or two parts of the ramp (inner, middle, and outer)
are present. Even if facies of all three parts were included, they would be grouped in
limited classes. Diagenetic processes are also similar in most cases. In an arid
climate, the main early diagenetic process is early dolomitization of the samples.
Another point is nonreservoir units. Below the cutoff level of porosity or per-
meability, samples have no reservoir properties. Caution should be taken using the
cutoff value. Both types of samples with porosity or permeability lower than the
cutoff value are nonreservoir and can be classified as nonreservoir rock types.
The following example is rock typing in a Permian–Triassic carbonate ramp
deposited in an arid climate in the Persian Gulf. The main lithologies are limestone,
dolomite, and anhydrite. Facies change from mud to grain-dominated and the main
diagenetic processes are dissolution and dolomitization. Rock types are:
• RT1: Anhydrite and mud/wackestone with anhydrite cement (nonreservoir)
• RT2: Marine cemented wackestone/packstone
• RT3: Dolomitic wackestone/packstone
6.1 Geological Rock Typing 87

Fig. 6.1 Various facies deposited in different environments with diagenetic processes lead to the
various rock types
88 6 Rock Typing

Fig. 6.2 A geological rock type can be composed of two or more facies or a facies can be divided
into two rock types

• RT4: Dissolved ooid/skeletal grainstone


• RT5: Marine cemented/dolomitized ooid/skeletal grainstone
The first group (RT1) is composed of nonreservoir samples with very low
porosity and permeability. Anhydrite and anhydrite cemented mud-dominated
samples are included. In RT2 samples have high micrite content but they also have
marine cements caused by preservation of primary interparticle porosity. Therefore
they have low to moderate porosity and permeability. Dolomitization in RT3 causes
moderate permeability but samples have low porosity due to their muddy nature. In
RT4, allochems of grain-dominated samples have been dissolved selectively in
diagenetic environments and thus samples have high porosity with low perme-
ability. In the last group, samples have the highest porosity and permeability.
Grain-dominated samples of this group have been cemented in marine environ-
ments or dolomitized in the later stages. Both processes resulted in high porosity
and permeability.
6.1 Geological Rock Typing 89

Another example is presented in Fig. 6.2 which integrates both facies and dia-
genetic properties to build final rock types. It can be seen in this figure that various
facies are integrated into one rock type and the same facies are classified in different
rock types. The defined rock types are as follows.
• GRT1 is ooid/bioclast packstone to grainstone with fabric-destructive dolomi-
tization or a crystalline carbonate. Both facies and lithology are different but the
final product of facies and diagenesis combination is one rock type.
• GRT2 has the same facies as GRT1 with fabric-retentive dissolution. It has high
porosity but they are not connected with each other.
• GRT3 has the same facies again deposited in near-shore environments and thus
is highly anhydritic cemented. It has no considerable porosity or permeability
and so is classified as a nonreservoir rock type.
• GRT4 changes from mudstone to bioclast wackestone with microporosity. It has
high porosity with very low permeability.
LØnØy (2006) tried to introduce a comprehensive system of pore type classi-
fication and related it to porosity–permeability distribution (Table 6.1).
He integrated both Choquette and Pray (1970) and Lucia’s (1995) terms of
porosity and added two other criteria including pore size and pore distribution to
their classifications. His study showed that there is very good correlation between
porosity and permeability in the final pore fabrics. LØnØy criteria are very effective
but also very time consuming. Classifying the pores into micro, meso, and macro
and then determining their distribution needs many thin sections, mercury injection
data, or SEM analysis and a lot of time which are not available in most cases. Also,
there are 19 pore fabrics in his classification. Such a number is too high for a
reasonable rock typing and to use them as a basis for special core analysis (SCAL)
sample selection or 3D reservoir modeling.
Winland related the pore throat size in the 35th percentile of mercury saturation
in a mercury injection test to the core porosity and permeability of the samples. His
equation was published by Kolodzie (1980) (Eq. 6.1) as

Log R35 ¼ 0:732 þ 0:588 Log Kair  0:864 Log Ucore ð6:1Þ

where R35 is the pore aperture radius in the 35th percentile of mercury saturation in
a mercury saturation test, Kair is the air permeability (in mD), and U is porosity (in
%). It is obvious that the three constants in this equation are different for various
hydrocarbon fields and can be calculated based on mercury injection tests of some
samples, preferentially from various rock types. Later studies showed that this
relationship is also applicable in various mercury percentiles (e.g., Rezaee et al.
2006) depending on the reservoir characteristics. The various pore throat sizes in
R35 could be used for rock type classification. In a standard Winland chart, these
include 0.2, 0.5, 1, 2, 5, 15, and 60 lm.
Lucia (2007) related the grain size to porosity–permeability distribution in
carbonate reservoirs. He concluded that in limy reservoirs, these parameters are
related to each other with some equations. The results were petrophysical classes
90 6 Rock Typing

Table 6.1 LØnØy (2006) classification of pore system


Pore type Pore size Pore Pore fabric R2
distribution
Interparticle Micropores Uniform Interparticle, uniform 0.88
(10–50 lm) micropores
Patchy Interparticle, patchy 0.79
micropores
Mesopores Uniform Interparticle, uniform 0.86
(50–100 lm) mesopores
Patchy Interparticle, patchy 0.85
mesopores
Macropores Uniform Interparticle, uniform 0.88
(>100 lm) macropores
Patchy Interparticle, patchy 0.87
macropores
Intercrystalline Micropores Uniform Intercrystalline, uniform 0.92
(10–20 lm) micropores
Patchy Intercrystalline, patchy 0.79
micropores
Mesopores Uniform Intercrystalline, uniform 0.94
(20–60 lm) mesopores
Patchy Intercrystalline, patchy 0.92
mesopores
Macropores Uniform Intercrystalline, uniform 0.80
(>60 lm) macropores
Patchy Intercrystalline, patchy
macropores
Intraparticle Intraparticle 0.86
Moldic Micropores Moldic micropores 0.86
(<10–20 lm)
Macropores Moldic macropores 0.90
(>20–30 lm)
Vuggy Vuggy 0.50
Mudstone Micropores Tertiary chalk 0.80
microporosity (<10 lm) Cretaceous chalk 0.81
Uniform Chalky micropores, uniform 0.96
Patchy Chalky micropores, patchy

and rock fabric numbers. He extended this concept to the dolostones and concluded
that a similar relationship exists between the porosity–permeability distribution and
dolomite crystal size. Again, the constant coefficients are different for different case
studies which can be calculated with other known data (see Lucia 2007 for more
explanation). Many studies tried to correlate various methods of rock typing (e.g.,
Chehrazi et al. 2011; Al-Tooqi et al. 2014; Rahimpour-Bonab et al. 2012, 2014;
Moradi et al. 2017; Riazi 2018).
6.2 Hydraulic Flow Units (HFU) 91

6.2 Hydraulic Flow Units (HFU)

The concept of flow unit is very close to the rock type but there are also some
differences in their definition. Bear (1972) defined the flow units as a volume of
reservoir rock where the geological and petrophysical properties are similar. Ebanks
(1987) added the mappable concept to the definition. It means that flow units should
be traceable on a regional scale. Hearn and his coworkers (1984) defined the flow
unit as a reservoir zone that is laterally and vertically continuous. Gunter and his
coworkers (1997) believed that a flow unit has similar reservoir properties, is
stratigraphically continuous, and honors the geological framework. Tiab and
Donaldson (2015) defined four characteristics for a reservoir unit including specific
reservoir quality lithology, correlative and mappable, recognizable on a wire line
log, and communication with other units. From all these definitions, it is concluded
that the concepts of the same reservoir and geological properties are similar for both
flow units and rock types but the difference is in their scale. Although flow unit is a
continuous traceable unit, the rock type has higher dispersion and heterogeneity in
the rock column.
There is also some research that formulated the concept of flow unit. Amaefule
and his coworkers (1993) showed in their experiments that the ratio of permeability
to effective porosity is a good indicator of reservoir hydraulic behavior. Their two
defined parameters included the reservoir quality index (RQI) and flow zone
indicator (FZI) which are related to each other (Eq. 6.3).

RQI ðlmÞ ¼ 0:0324 ðK=UÞ1=2 ð6:2Þ

FZI ¼ RQI=Uz ð6:3Þ

Uz ¼ U=ð1  UÞ ð6:4Þ

where K is permeability (mD), U is porosity (fraction), and Uz is the ratio of pore


volume to grain volume (fraction).
Various flow units are defined with respect to their different FZI. The limits are
mostly 0.1, 0.2, 0.5, 1, and 2. The final result is routinely presented by plotting
porosity versus permeability for various rock types or flow units (Fig. 6.3).
The stratigraphic modified Lorenz plot (SMLP) is another way to determine the
HFUs of the reservoir. The Lorenz technique was developed for measuring
heterogeneity of wealth across a population (Lorenz 1905). Later, the technique was
used in petroleum geology by plotting cumulative flow capacity, which is perme-
ability multiplied by thickness of the zone against cumulative thickness (Schmalz
and Rahme 1950). Generally, plotting the cumulative of the property of interest
against cumulative depth represents the amount of heterogeneity. It is obvious that
if the values increase by a constant coefficient, the resulting plot will be a straight
line. This line is known as the line of perfect quality. Any heterogeneity causes
departure from the perfect line. The Lorenz coefficient (Lc) is expressed as twice the
92 6 Rock Typing

Fig. 6.3 Plot of porosity–permeability for Winland method (a), plot of Uz versus RQI (b),
porosity–permeability plot for FZI method (c), and Lucia rock classes (d) for a Permian–Triassic
carbonate reservoir in the Persian Gulf. Colors indicate the different values of the calculated
parameter (Courtesy of M. Nazemi)

area between that line, known as the Lorenz curve, and the perfect quality line. In
the SMLP method, each porosity and permeability datum measured on core plugs is
multiplied by the thickness of the measurement. The results are called storage
capacity and flow capacity for porosity and permeability, respectively. A SMLP is a
plot of cumulative storage capacity versus cumulative flow capacity, both nor-
malized to 100%. The best result is achieved when enough data are available within
a reasonable distance. Any abrupt change in the line slope reflects the shift in
reservoir characteristics. The slope of the line represents the gradient of the change
in reservoir properties.
6.3 Reservoir Data Integration 93

6.3 Reservoir Data Integration

Integrating porosity–permeability data with geological variations represents the


effect of various geological parameters on reservoir properties. Such relationships
explain the reservoir data variations and facilitate the prediction of 3D reservoir
properties distribution. For example, distribution of reservoir properties such as
porosity, permeability, pore size, pore type, rock type, and pore throat size distri-
bution can be considered in each facies. This clearly shows the effect of facies’
characteristics on reservoir quality and also makes it possible to trace the facies and
assign reservoir properties within the reservoir. Plotting various data versus each
other is useful. The porosity–permeability plot is recommended (Fig. 6.4). Plotting
various data against depth (preparing a log) is another suitable method.

Fig. 6.4 Integrating porosity and permeability data with lithology (a), facies (b), sedimentary
environment (c), and geological rock type (d). These diagrams represent the effective geological
parameters of reservoir characteristics. MF1: anhydrite, MF2: mudstone, MF3: bioclast
wackestone, MF4: crystalline carbonate, RT1: nonreservoir, RT2: marine cemented wacke/
packstone, RT3: dolomitic wackestone/packstone
94 6 Rock Typing

6.4 Defining Electrofacies

Every parameter that is defined in the reservoir studies should be correlated to the
wire line logs because these data are available from many reservoirs. Wire line data
are a continuous record of rock and fluid properties of the borehole wall. Following
the main concept one specific rock type resulting from the wire line data should
have the same log characteristics. Similar rocks in terms of their log properties are
called electrofacies (EF) (Serra and Abbott 1980).
Data quality control is the first step in any scientific data analysis research.
Therefore checking the difference between the caliper and bit size is strongly rec-
ommended before any other analysis. If the difference is more than a specific size
(1.5 in. in most cases), the data are not usable. The DRHO log indicates the
washout of the borehole wall and thus needs to be checked carefully. If DRHO is
more than 0.1 g/cm3, the density (and maybe some other logs) is failed. The
environmental corrections are also necessary before starting any wire line log
analysis. The depth-matching process between core and log is very important
because rock types from log data are compared with the core analysis results. If data
are relatively old, despiking is also necessary.
We are dealing with a multivariate datatype when we analyze the wire line logs.
This means that there are some values for one observation. For downhole data, the
observation is depth and the values are wire line logs. Each point or observation
consists of many log values. For example, a depth point in the well has GR, NPHI,
resistivity, DT, and RHOB logs. Calculating the similarity between each pair of
points means that the user should calculate how close they are to each other or what
the distance between these two points is. When the distance is known, pairs with
minimum distances form the first group. We can realize the distance in a 2D or 3D
space, but what about more than three values?
Before starting the data grouping in a multidimensional space, it is important to
consider data relationships. In a log grouping task, the relationships between all
data should be considered first. This is possible through single or multiple
regression analysis. In a single regression, there is just one dependent and one
independent variable (Eq. 6.5), whereas in multiple regression one log is selected as
dependent and the others as independent variables (Eq. 6.6):

y ¼ a0 þ a1 x ð6:5Þ

y ¼ a 0 þ a1 x 1 þ a2 x 2 þ    þ an x n ð6:6Þ

The coefficients are determined with the specified mathematical formula.


Relevant software is used in most cases. The data predicted by the equation are
plotted against the real values. The slope, intercept, and coefficient of determination
(R2) are used to infer the response variable variation that is explained by a linear
model. It is obvious that the slope should be as close as possible to one, the
intercept to zero, and R2 to one (Fig. 6.5).
6.4 Defining Electrofacies 95

Fig. 6.5 Predicted bulk density (a) and computed gamma ray (b) with other logs using multiple
regression analysis in the Asmari Oligo-Miocene reservoir in Iran

Now we go back to the question of distance calculation. Let’s start from just two
logs. For these two data, a scatterplot will be a good indicator of the distances. The
graphical distribution is obvious and the user can divide the data into various
clusters (Fig. 6.6a). The data distance can be calculated from various methods; the
well-known one is the Euclidean formula. This is the straight line distance between
two points p and q with coordinates p (p1, p2) and q (q1, q2) calculated by the
following formula in a 2D space.
 1=2
d ðp; qÞ ¼ d ðq; pÞ ¼ ðq1  p1 Þ2 þ ðq2  p2 Þ2 ð6:7Þ

Fig. 6.6 The concept of clustering hierarchy in a scatterplot (a) and dendrogram (b)
96 6 Rock Typing

Note that p and q are two points (equivalent to two depth or log readings), p1 and
p2 are two different logs in point p, and q1 and q2 are values of the same logs in
point q. The Euclidian distance is the most well-known method. Many other for-
mulas have been developed for various cases. Each of these formulas can be
extended to a multidimensional space. For example, the Euclidian distance for these
two points with the coordinates p (p1, p2, p3, …, pn) and q (q1, q2, q3, …, qn) is
written as
 1=2
d ðp; qÞ ¼ ðq1  p1 Þ2 þ ðq2  p2 Þ2 þ    þ ðqn  pn Þ2 ð6:8Þ

Again note that p1 to pn are the values of different logs in point p and q1 to qn are
the values of the same logs in point q. The distance between each pair of data is
calculated by the appropriate method and pairs with less distance are grouped
together. The result is a matrix with one row including all pair distances (Eq. 6.9).
This is called a distance matrix (dm).

dm ¼ ½d ð1; 2Þ; d ð1; 3Þ; d ð1; 4Þ; . . .; d ð1; nÞ; d ð2; 3Þ; d ð2; 4Þ; . . .; d ðn  1; nÞ ð6:9Þ

where numbers are points.


Now the groups of pairs should be integrated to build the bigger groups. There
are also different formulas for calculating the distance between two groups or
clusters. This process continues until there is just one big cluster including all data
(Fig. 6.6a, b). The user may stop this process at a desired distance. Such a distance
is called the cutoff level and completely depends on the project objectives. Actually,
this distance determines the final number of clusters or electrofacies. This process is
called cluster analysis and is shown graphically by a dendrogram (Fig. 6.6b).
Cluster analysis is the most accepted method for electrofacies determination in
reservoir studies (e.g., Gill et al. 1993; Ye and Rabiller 2000; Tavakoli and Amini
2006). There are also other methods for defining electrofacies including drawing 2D
plots of logs, composite logs, radar diagrams, and artificial intelligence (such as
neural networks) but the statistical methods have some advantages. Statistical
methods do not need any training prior to analysis whereas most of the artificial
intelligence systems must be trained with known results. The results are not always
available before analysis. The process of calculations and clustering is clear in
statistical methods whereas the internal layer of some intelligent methods is not
understandable to the human mind. The final results of statistical cluster analysis are
flexible and the user can modify them in many ways. Other methods are not
applicable to huge data volume.
The final problem is the reservoir properties of each electrofacies. The electro-
facies is just a number prior to this step. Such numbers show nothing about the
nature of the samples. For solving this problem, the results of log data analysis are
useful. Many parameters are derived from logs using various methods and software.
These results are correlated with each electrofacies. For example, the EF number
one has a specific dominant lithology and specified porosity value. As the fluid
6.4 Defining Electrofacies 97

Fig. 6.7 Graphical representation of the relationship between various methods of rock typing

content of the rock has a major effect on some wire line data, the resulting facies
also reflect the water saturation of the samples in most cases. Therefore the Sw is
also attributed to an EF. The other more reliable method is using core data, when
available. Various parameters result from the core analysis in routine core analysis
(RCAL), SCAL, geology, or even geochemical analysis. Once established, elec-
trofacies are correlated with core-derived rock types if the logged interval has been
cored. These final electrofacies are stratigraphically correlated in the field in a
sequence stratigraphic framework. Figure 6.7 shows the relationship between all
methods of rock typing.

6.5 Comparison and Final Results

The final rock types from geology, RCAL, SCAL, and wire line log data should be
integrated into one rock typing system. The integration of RCAl, geology, and wire
line log data were considered earlier. The SCAL results are also integrated with
previous rock types. It should be mentioned that if the previous steps have been
done accurately, the SCAL results should be fitted in the determined rock types.
Anyway, some modifications may be necessary.
The important point is that some SCAL tests, such as relative permeability, are
related to the fluid properties, not the rock characteristics. In the relative perme-
ability test, the results depend on the type and saturation of the fluids and are not an
intrinsic property of the rock. Such characteristics cannot be used for static reservoir
rock typing. Actually, the dynamic characters are simply assigned to the previously
determined rock types. They do not change the rock type of each sample or the rock
type distribution in the well or field.
98 6 Rock Typing

References

Al-Tooqi S, Ehrenberg SN, Al-Habsi N, Al-Shukaili M (2014) Reservoir rock typing of Upper
Shu’aiba limestones, northwestern Oman. Petrol Geosci 20:339–352
Amaefule JO, Altunbay MH, Tiab D, Kersey DG, Keelan DK (1993) Enhanced reservoir
description using core and log data to identify hydraulic (flow) units and predict permeability in
uncored intervals/wells. In: Paper presented at SPE annual technical conference and exhibition,
Houston, Texas, 3–6 Oct 1993
Bear J (1972) Dynamics of fluids in porous media. Elsevier, New York
Chehrazi A, Rezaee R, Rahimpour-Bonab H (2011) Pore-facies as a tool for incorporation of
small-scale dynamic information in integrated reservoir studies. J Geophys Eng 8:202–224
Choquette PW, Pray LC (1970) Geological nomenclature and classification of porosity in
sedimentary carbonates. AAPG Bull 54:207–250
Ebanks WJ (1987) The flow unit concept—an integrated approach to reservoir description for
engineering projects. Am Assoc Geol Annu Conv 71:551–552
Gill D, Shomrony A, Fligelman H (1993) Numerical zonation of log suites and logfacies
recognition by multivariate clustering. AAPG Bull 77:1781–1791
Gunter GW, Finneran JM, Hartman DJ, Miller JD (1997) Early determination of reservoir flow
units using an integrated petrophysical method. In: Paper presented at SSPE annual technical
conference and exhibition, San Antonio, 5–8 Oct 1997
Hearn CL, Ebanks WJ, Tye RS, Ranganatha V (1984) Geological factors influencing reservoir
performance of the Hartzog draw field, Wyoming. J Petrol Technol 36:1335–1344
Kolodzie S (1980) Analysis of pore throat size and use of the Waxman-Smits equation to
determine OOIP in Spindle Field, Colorado. In: Paper presented at 55th annual fall technical
conference of society of petroleum engineers, Dallas, Texas, 21–24 Sept 1980
LØnØy A (2006) Making sense of carbonate pore systems. AAPG Bull 90:1381–1405
Lorenz MO (1905) Methods of measuring concentration of wealth. Am Stat Assoc 970:209–219
Lucia FJ (1995) Rock fabric/petrophysical classification of carbonate pore space for reservoir
characterization. AAPG Bull 79:1275–1300
Lucia FJ (2007) Carbonate reservoir characterization. Springer, Berlin
Moradi M, Moussavi-Harami R, Mahboubi A, Khanehbad M, Ghabeishavi A (2017) Rock typing
using geological and petrophysical data in the Asmari reservoir, Aghajari Oilfield, SW Iran.
J Petrol Sci Eng 152:523–537
Rahimpour-Bonab H, Mehrabi H, Navidtalab A, Izadi-Mazidi E (2012) Flow unit distribution and
reservoir modelling in Cretaceous carbonates of the Sarvak Formation, Abteymour Oilfield,
Dezful Embayment, SW Iran. J Pet Geol 35:213–236
Rahimpour-Bonab H, Enayati-Bidgoli AH, Navidtalab A, Mehrabi H (2014) Appraisal of intra
reservoir barriers in the Permo-Triassic successions of the Central Persian Gulf, Offshore Iran.
Geol Acta 12:87–107
Rezaee MR, Jafari A, Kazemzadeh E (2006) Relationships between permeability, porosity and
pore throat size in carbonate rocks using regression analysis and neural networks. J Geophys
Eng 3:370–376
Riazi Z (2018) Application of integrated rock typing and flow units identification methods for an
Iranian carbonate reservoir. J Petrol Sci Eng 160:483–497
Schmalz JP, Rahme HS (1950) The variations in water flood performance with variation in
permeability profile. Prod Mon 15:9–12
Serra O, Abbott HT (1980) The contribution of logging data to sedimentology and stratigraphy. In:
SPE9270 presented at the 1980 annual technical conference and exhibition, Dallas, TX
Tavakoli V, Amini A (2006) Application of multivariate cluster analysis in logfacies determination
and reservoir zonation, case study of Marun Field, South of Iran. J Sci Univ Teheran 32:69–75
References 99

Tiab D, Donaldson EC (2015) Petrophysics, theory and practice of measuring reservoir rock and
fluid transport properties. Gulf Professional Publishing, Houston
Ye SJ, Rabiller P (2000) A new tool for electro-facies analysis: multi-resolution graph-based
clustering. In: Paper presented at SPWLA 41st annual logging symposium, society of
petrophysicists and well-log analysts. Dallas, Texas, 4–7 June 2000

You might also like